Quantitative mineralogical mapping of hydrated low pH concrete

Quantitative mineralogical mapping of hydrated low pH concrete

Accepted Manuscript Quantitative mineralogical mapping of hydrated low pH concrete Stéphane Gaboreau, Dimitri Prêt, Valérie Montouillout, Pierre Henoc...

13MB Sizes 0 Downloads 36 Views

Accepted Manuscript Quantitative mineralogical mapping of hydrated low pH concrete Stéphane Gaboreau, Dimitri Prêt, Valérie Montouillout, Pierre Henocq, Jean-Charles Robinet, Christophe Tournassat PII:

S0958-9465(16)30747-8

DOI:

10.1016/j.cemconcomp.2017.08.003

Reference:

CECO 2881

To appear in:

Cement and Concrete Composites

Received Date: 18 November 2016 Revised Date:

7 July 2017

Accepted Date: 7 August 2017

Please cite this article as: Sté. Gaboreau, D. Prêt, Valé. Montouillout, P. Henocq, J.-C. Robinet, C. Tournassat, Quantitative mineralogical mapping of hydrated low pH concrete, Cement and Concrete Composites (2017), doi: 10.1016/j.cemconcomp.2017.08.003. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Quantitative Mineralogical mapping of hydrated low pH concrete

1 2 3

RI PT

6 7 8 9 10 11 12 13

Stéphane Gaboreaua,*, Dimitri Prêtb, Valérie Montouilloutc, Pierre Henocqd, JeanCharles Robinetd, Christophe Tournassata,e a

BRGM, Environment and Process Division, 3, avenue Claude Guillemin, F-45060 Orléans Cedex 2, France b UMR CNRS 7285 IC2MP, Université de Poitiers, Equipe HydrASA, rue Albert Turpain, Bat B8, 86022 Poitiers, France c CNRS-CEMHTI UPR 3079, 1 Avenue de la Recherche Scientifique, 45071 Orléans, cedex 2 France d Andra, 1/7 rue Jean Monnet, Parc de la Croix Blanche, 92298 Châtenay-Malabry Cedex, France * Corresponding author: [email protected] e Earth and Environmental Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, USA

SC

4 5

M AN U

14

Abstract

16

Concrete materials are made of various minerals and phases, whose spatial

17

heterogeneous distributions impact the overall physical and chemical properties of the

18

materials. We have investigated the heterogeneous distribution of minerals and phases

19

in two types of concrete using quantitative X-ray intensity maps coupled with an

20

innovative data treatment method based on image segmentation. This method provided

21

quantitative data on spatial distribution, modal content and associated calculated

22

formulas for each identified mineral and phase in the binder with micrometer resolution.

23

We also obtained quantitative information on the porosity associated with the phases,

24

making it possible to differentiate poorly hydrated cement phases (initial clinker

25

hydration reaction) from highly hydrated phases (final cement product) despite their

26

similar chemical composition, when expressed in terms of cationic formulas. We

27

quantified the mineralogical and phase contents, independent of crystal size or

AC C

EP

TE D

15

1

ACCEPTED MANUSCRIPT crystallinity considerations. We report spatial resolution in the pozzolan hydration

29

process over different observation scales for the two investigated concretes.

30

Keywords: concrete, quantitative mapping, mineralogy, phases, porosity

AC C

EP

TE D

M AN U

SC

RI PT

28

2

ACCEPTED MANUSCRIPT

1. Introduction

32

Concretes are complex finely divided materials made of components of various natures

33

(cement, aggregates, water, and more) and with variable sizes. They are dynamic

34

materials that pass from a liquid to a solid state when prepared. While in the solid state,

35

hydration processes with slow kinetics lead to changes in concrete microstructure and

36

phase composition. Hydration results in the formation of various calcium silicate and

37

calcium aluminate hydrates, among other hydrates. Amounts, composition and

38

distribution are sensitive to the water/cement (w/c) ratio, the relative humidity and the

39

reaction kinetics; the chemistry stays the same with a redistribution of elements through

40

transition and transformation reactions like dissolution/precipitation. Many different

41

formulations may be used depending on the industrial application. The cement

42

component of concretes contains mainly portland cement clinkers, but can also contain

43

supplementary cementitious materials (SCM) such as fly ash (FA), metakaolin (MK),

44

blast furnace slag (BFS) and silica fume (SF) in relative amounts that depend on the

45

cement formulation. Using these formulations mitigates detrimental physical and

46

chemical effects caused by the environment surrounding the concrete such as sulfate

47

attack [1, 2], chloride attack [3], alkali-silica reaction [4] or carbonation [5]. Obviously,

48

hydration of cementitious materials involves complex chemical reactions. Their service-

49

life depends on the initial formulation and interactions with the environment over time.

50

Concrete mineralogy and the associated pore solution composition evolve according to

51

the initial cement composition, the kind and amount of SCM (SF, MK, FA, BFS) and

52

their respective reactivity. To assess the locally uneven distribution and reactivity of the

53

materials and the associated complex hydrate distribution, imaging techniques are

54

necessary. To tackle this type of problem, the method proposed (e.g. chemical X-ray

55

maps obtained with electron probe microanalysis) allow computing mineral/phase maps

AC C

EP

TE D

M AN U

SC

RI PT

31

3

ACCEPTED MANUSCRIPT based on procedures of chemical segmentation based on ternary scatterplot projections.

57

This method, initially developed by [6, 7] for clay materials, was improved and adapted

58

to the case of concrete materials in order to map the mineralogy with µm resolution on

59

millimeter-scale areas, in which the hydration processes were resolved by quantifying

60

the anhydrous residual components (i.e., clinker phases and SCMs) and analyzing the

61

spatial chemical evolutions resulting from hydration. To illustrate the numerous

62

advantages of this mineralogical mapping method, we studied two concrete

63

formulations with various substitutions with SF, BFS and MK and curing conditions

64

(hydric conditions). The chemistry of the hydraulic binders was characterized and

65

quantified, considering the initial formulation.

66

2. Materials & Methods

67

2.1. Concrete formulations

68

We studied two different low-pH concretes (Concretes I and II), with the Portland

69

cement partially substituted by SCMs, as silica fume, blast furnace slag and metakaolin.

70

Both samples were provided by the French National Radioactive Waste Management

71

Agency (ANDRA) as part of the Cigéo Waste Disposal Centre project. The detailed

72

compositions of both concretes are given in table 1.

SC

M AN U

TE D

EP

Table 1 concrete compositions

AC C

73

RI PT

56

Components Particle size Cement CEM I 52.5N <100 µm Cement additive SF <10 µm BFS <100 µm MK <20 µm Aggregates Carbonates 4-12 mm sand 0-4 mm

Concrete I

wt%

wt% (normalized to cement)

kg m-3

Concrete II

wt%

wt% (normalized to cement)

kg m-3

79

3.8

22.0

450

21.4

86.1

128 188

4.7 8.6

27.6 50.4

50

1.6

6.5

50

1.8

7.4

736 900

75.1

1035 813

82.9

4

ACCEPTED MANUSCRIPT Water Organic additives (Plasticizer) 74

181 11

161 14

Low-pH concrete I was a ternary mixture of CEM I 52.5 N CE PM ES CP2 NF

76

(Lafarge, Val d’Azergues), Slag (Orcem) and SF (CONDENSIL® S95 DM) with 80%

77

of substitution of the clinker by the SCMs (Table 1). Concrete I was drilled in the

78

LSMHM Underground Research Laboratory (URL, Meuse/Haute Marne, France),

79

where it had been in contact with the Callovian-Oxfordian clay-rock formation in a bore

80

hole located three meters below the floor of an access gallery. The experiment was

81

dismantled one year after its implementation. No ingress of pore water from the

82

surrounding clay-rich rock occurred during this period and the concrete was exposed to

83

atmospheric conditions.

84

Concrete II was prepared with the same CEM I and SF (CONDENSIL S95 DM) as

85

concrete I, with added MK (ARGICAL-M 1000) with 20 % of substitution of the

86

clinker by the SCMs (Table 1). Concrete II was cured for one year in a desiccator at

87

equilibrium with a CO2-free atmosphere having relative humidity close to 100 %, to

88

enable the hydration of the cement phases, while avoiding carbonation. The concrete’s

89

coarse and fine aggregates were calcareous aggregates and quartz (Boulonnais, France).

90

Considering the substitution of the clinker by the SCMs in both concretes (80 and 20 %

91

in concrete I and II, respectively), the Ca/Si ratio of the hydraulic binder should be

92

lower in the case of concrete I.

93

2.2. Bulk chemical and mineralogical characterization

94

Si, Al, Ti, Fe (total), Mn, Ca, Mg, K and Na were chemically analyzed using a PW2400

95

sequential X-ray fluorescence (XRF) spectrometer (Philips). The amount of total carbon

96

(TC) and sulfur were determined by infra-red spectroscopy after burning the samples at

AC C

EP

TE D

M AN U

SC

RI PT

75

5

ACCEPTED MANUSCRIPT 97

900°C in an oxygen atmosphere. Carbonate contents were also measured by dissolution

98

in an HCl solution and titrating the CO2 produced using a volumetric method. The bulk

99

chemical analyses of concrete I and II together with the composition of CEMI and the SCMs are given in table 2.

101 102

Table 2 Global chemical composition of the concretes for major elements (in weight percentage)

RI PT

100

M AN U

SC

Na2O K2O CaO MgO SiO2 Al2O3 Fe2O3 Total Organic MnO TiO2 S LOI C C total Concrete I 0.12 0.08 48.7 1.5 9.8 1.4 0.37 9.97 1.05 0.04 0.09 0.11 36.5 Concrete II 0.09 0.10 45.2 0.8 13.8 1.7 0.87 8.67 1.1 0.02 0.11 0.14 36.5 CEMI 52N 0.1 0.6 65.1 0.6 20.9 3.4 4.4 2.7 1.3 SF <0.2 0.3 0.5 <0.2 96.3 <0.2 0.1 0.02 <0.05 2.2 BFS <0.2 0.2 42.7 7.2 36.0 11.6 0.6 0.2 0.5 <0.1 MK 0.8* 0.3** 55.0 40.0 1.4 1.5 0.3 1.0 103 LOI: loss on ignition at 1000°C; * Na2O+K2O; **CaO+MgO Total bulk porosity were measured by kerosene porosity method [8] and calculated from

105

measured grain density (ρgr) (helium pycnometry) and apparent dry density (ρ)

106

(mercury intrusion porosimetry) according to the following equation [8] :

TE D

104

= 1−

The measured and calculated porosity are 16 ± 1 % and 13 ± 1 % and the grain density

108

at 2.61 g cm-3 and 2.64 g cm-3, for concretes I and II respectively. The total porosity

109

includes the capillary and gel porosity. The local silicon and aluminum environment in

110

both concretes was probed by solid-state NMR.

111

NMR spectra were acquired at 59 MHz on a Bruker AVANCE 7.4 T (300 MHz)

112

spectrometer equipped with a 4 mm double-bearing MAS probe-head spinning at 12

113

kHz. About 20,000 scans were accumulated after a 45° pulse, using 10 s recycling

114

delay. This delay was optimized to ensure complete magnetization relaxation.

29

115

chemical shifts were reported relative to tetramethylsilane (TMS) resonance. The

27

AC C

EP

107

29

Si Magic Angle Spinning (MAS)

Si

Al

6

ACCEPTED MANUSCRIPT NMR high-resolution NMR experiments were carried out on a Bruker AVANCE

117

instrument (magnetic field 17.6 T–750 MHz) equipped with high speed MAS

118

probeheads (spinning rates of 30 kHz in aluminum-free zirconia rotors, diameter

119

2.5 mm). The 1D MAS spectra were acquired after a single short pulse (π/10) ensuring

120

quantitative excitation and identification of the 27Al central transition [9]. 29Al chemical

121

shifts were reported relative to Al(NO3)3 1M resonance. All the spectra were

122

deconvoluted using the Dmfit program [10] into individual Gaussian-Lorentzian peaks,

123

whose integration corresponded to the relative amount of the differently coordinated

124

species.

125

2.3. Electron Probe MicroAnalysis (EPMA) and quantitative chemical maps

126

Samples were fully impregnated with methylmethacryle (MMA) using an impregnation

127

procedure already described in the literature [6-8, 11-13]. Before impregnation, a

128

parallelepiped of 8x4x3 (length, width, heigh) was cutted with a diamond wire saw.

129

This impregnation made it possible to prevent physical perturbation while preparing

130

polished section. The samples were the polished with different diamond suspensions (3,

131

1, ¼ µm) for 1 hour.

AC C

EP

TE D

M AN U

SC

RI PT

116

132 133 134 135 136 137

Figure 1 – Mosaic of BSE images of concrete I (A). The white square represents the size of the analyzed area, illustrated with the BSE image (B) and the associated quantitative chemical map of Ca Kα for a dwell time of 100 ms (C). The brightest parts of the maps represent the highest Ca wt%. On the BSE image, the anhydrous phases appear in light gray, the hydraulic binder in dark gray and the macropores in black.

7

ACCEPTED MANUSCRIPT Quantitative X-ray intensity maps and a BSE image (Figure 1) were acquired with a

139

Cameca SX Five EPMA equipped with five wavelength dispersive spectrometer (WDS)

140

and operating at 15 keV and 30 nA. To investigate the distribution of the ten elements

141

identified in the concrete (Table 2), the area was scanned twice, because the number of

142

the simultaneously detected elements was constrained by the number of available

143

spectrometers. We collected Kα peak intensities (for Si, Al, Fe, K, Na, Ca, Mg, Ti, Mn

144

and S) using large Large Thallium Acid Phtalate (LTAP) and Large Pentaerythritol

145

(LPET) and Pentaerythritol (PET), Thallium Acid Phtalate (TAP) and Lithium Fluoride

146

(LIF) monochromator crystals, allowing for a high counting rate and short dwell time

147

(~4 s) for a quantitative point analysis. This dwell time is too time-consuming for a

148

mapping mode, so we used a shorter counting time of 100 ms per pixel, as discussed in

149

[6]. After the acquisition, no evidence of beam damage was identified on the mapped

150

area. To reduce the total acquisition time (two days), we did not measure background

151

with subtraction from X-ray emission peaks as its contribution at short dwell times and

152

for major concentrations is low [14], and as recording it would have doubled the

153

acquisition time. A standard-based PHIRHOZ matrix correction [15] was then applied

154

to provide a weight percentage for each element per pixel. The 512 by 512 pixel

155

elemental maps were recorded by stage rastering using a stationary beam, with spatial

156

resolution of 2 µm per pixel.

157

2.4. Mineralogical mapping from quantitative chemical maps

158

Mineralogical maps were created for both concretes to display the spatial distribution of

159

the minerals over a surface area of 1 x 1 mm² with resolution of 2 µm (Figure 1)

160

following the methodology developed by [6] and using the µMAPphase software [6].

161

This methodology consists in identifying the mineral phases composing the analyzed

162

area from step by step projections of the scanned elemental composition points, and

AC C

EP

TE D

M AN U

SC

RI PT

138

8

ACCEPTED MANUSCRIPT converting them into ternary plots from which we can visualize contrasting chemical

164

compositions for all the mineral phases.

SC

RI PT

163

166

EP

TE D

M AN U

165

Figure 2 – BSE images and Si – Ca – Al3FeMg ternary projections of all the pixels of the mapped areas of concrete I and II. The main aggregates, pozzolan and hydrate end members of a hardened cement material are added on the concrete I and II scatterplots.

170

All the pre-processing steps that convert the initial quantitative X-ray maps given in

171

element weight percent (Ati wt%) into element molar percent (Ati mol%) have been

172

described in [6]. This conversion led to a chemical oxide composition for each pixel of

173

the map. All the pixels of the mapped area were plotted in ternary diagrams, where each

174

axis of the plot represented the concentration of an element or a combination of

175

elements. This method was not affected by porosity variations as normalized axis

AC C

167 168 169

9

ACCEPTED MANUSCRIPT weights were used to generate the ternary scatter plots. We identified several clusters of

177

pixels in the chemical ternary plots. These clusters represented the chemical

178

compositional fields of one mineral. Their stoichiometry was compared directly to the

179

different mineral end-members by adding their theoretical compositions to the

180

projection (Figure 2). For mixtures or solid solutions, clusters were stretched along lines

181

between the end-members. Pixels with similar chemical compositions (i.e. a cluster)

182

were selected with a polygon tool and back-projected on a mineral map using the same

183

color as the selected polygon. This method of segmentation based on ternary diagrams

184

was particularly efficient for locating small crystals and mineral mixtures [6]. For our

185

study, another advantage of this method was the possibility of comparing the chemistry

186

of the two hardened concretes based on the position of the clusters in the projections

187

(Figure 2). For most materials, the projection using only three elements was not

188

sufficient to distinguish all the mineral phases, so we used a succession of ternary

189

diagrams where the axes covered a large range of element combinations. The theoretical

190

compositions of the main clinker phases, pozzolans, hydrates and aggregates were

191

superimposed on each of these diagrams in order to facilitate reading and interpretation

192

of these chemical scatterplots (Figure 2).

193

3. Results

194

3.1. NMR results

195

The 29Si solid-state NMR spectrum of concrete I was complex as seen in Figure 3 (A),

196

but the majority extends over the typical C-S-H chemical shift region ranging from -70

197

ppm to -100 ppm [16]. The main signal was deconvoluted into at least four components

198

corresponding to the Q1, Q2b, Q2 and Q3 species [17]. Q1 refers to an end of chain silica

199

or dimers connected to only one neighbor; Q2b and Q2 refer a bridging tetrahedron and

200

to pairing silica respectively, connected to two neighbors; and Q3 refers to a crosslinked

AC C

EP

TE D

M AN U

SC

RI PT

176

10

ACCEPTED MANUSCRIPT 201

silica connected to three other tetrahedra. The signal between -104 and -120 ppm was

202

probably related to the presence of quartz and amorphous silica. In the range -70/-100

203

ppm,

204

population characterizing long silicon tetrahedron chains in the C-S-H structure. The

205

presence of long tetrahedron chains is related to low C/S ratio C-S-H. According to the

206

equation published by [18], the C/S ratio calculated from the Q1/Q2tot content is at ~ 0.6,

207

where Q2tot is the total population of Q2 and Q2b. We assigned a weak signal at -71.4

208

ppm to clinker phases (alite – belite) [17, 19]. The

209

concrete I (Figure 3A) had two areas: a broad peak between 40 and 80 ppm

210

corresponding to tetracoordinated aluminum associated with C-S-H, and a much

211

narrower peak centered at 6 ppm, corresponding to hexacoordinated aluminum. The

212

aluminum in tetrahedral sites was assigned to Al substituting for Si in the C-S-H

213

structure [20], whereas the octahedrally coordinated Al was assigned to a low

214

crystalline, disordered aluminate, the so-called “third aluminate hydrate” (TAH) [21-

215

23].

216

The 29Si and 27Al NMR spectra acquired for concrete II were quite different from that of

217

concrete I (Figure 3 B). In addition to C-S-H contributions between -75 and -95 ppm,

218

the

219

were assigned to anhydrous clinker phases (C3S – C2S) and quartz respectively [17, 19].

220

This difference is coherent as regards to the proportion of clinker in both concretes. In

221

concrete II, only 20 % of clinker was substituted by SCMs while in concrete I 80 % was

222

substituted. Consequently, the relative intensity of the C-S-H signal compared to the

223

clinker phase signal was weaker than for concrete I. In terms of Q1 and Q2 populations,

224

29

225

Figure 3 (B) clearly displays that Q1 and Q2 populations were quite similar traducing the

29

Al solid state NMR spectrum of

Si spectrum was composed of two narrow signals at -71.4 and -107.2 ppm that

AC C

29

EP

TE D

M AN U

SC

27

RI PT

Si NMR in Figure 3 A showed a predominant Q2 population compared to Q1

Si NMR showed also significant differences between concrete I and concrete II.

11

ACCEPTED MANUSCRIPT 226

greater presence of dimers in the tetrahedron chains in the C-S-H structure. The C/S

227

ratio of C-S-H calculated from the Q1/Q2tot content was ~ 0.9 [18]. The higher C/S ratio

228

of C-S-H in concrete II is consistent with the expected C/S regarding the concrete

229

composition and the chemical composition (Table 1 and 2).

230

exhibited a weak tetrahedral contribution between 50 and 80 ppm, corresponding to Al

231

substituting for Si in the C-S-H structure and a narrow signal in the range of

232

hexacoordinated aluminum. This resonance was composed of at least three

233

contributions obtained by deconvolution, centered at 6, 10 and 13 ppm, and was

234

assigned to many aluminum hydrates such as TAH [21-23], hydrogarnet [22, 24], AFm,

235

hydrotalcite and ettringite [23]. There was too much overlap to assign these signals to a

236

specific phase. The relative intensity of the C-S-H signal compared to that of phases

237

bearing hexacoordinated aluminum was weaker than for concrete I. Another broad

238

NMR shift ranging between 20 and 40 ppm was detected and possibly match the

239

metakaolin signal [25].

27

Al spectrum

AC C

EP

TE D

M AN U

SC

RI PT

The

240 241 242

Figure 3 – 29Si and 27Al NMR spectra of concrete I (A) and concrete II (B). The dotted lines are the result of the spectral decomposition. 12

ACCEPTED MANUSCRIPT 3.2. Image segmentation step 1: macroporosity and anhydrous phases

M AN U

SC

RI PT

243

244

Figure 4 – First processing step to threshold the macroporosity and the poorly hydrated cement phases of the hardened concrete I. The frequency histogram of the sum of atomic wt% was used to threshold the low hydrated phases of the whole map by limits. The thresholded pixels were plotted in a Si-Ca-Al3FeMg ternary diagram to discriminate the different poorly hydrated phases. The pixels selected with the polygon tool are displayed on a mineral map and superimposed on the BSE image.

251

The first step of mineralogical mapping was to segment the pore network, as described

252

in [6]. The macropores, corresponding to pore larger than the image pixel size (e.g., 2

253

µm) were segmented from the BSE images and the associated derivative of the

254

histogram of the gray level frequency (figure 4). All the pixels attributed to the

255

macropore network were back projected onto the BSE image and mapped. Only the

256

remaining pixels were used in the subsequent segmentation steps to identify minerals

257

(Figure 4).

258

One of the main issues with this method was to differentiate poorly hydrated initial

259

phases (anhydrous cement and additive phases) from highly hydrated phases with

AC C

EP

TE D

245 246 247 248 249 250

13

ACCEPTED MANUSCRIPT similar chemical composition. This issue was related to the kinetics of hydration of

261

clinker phases and slag grains, some of which remained unhydrated at the time of

262

sample analysis. As elements with low atomic numbers (H, C, O) were not analyzed by

263

EPMA, the sum of element weight concentration analyzed per pixel Σ (oxide) wt%

264

(including the stoichiometric O content) varied according to the proportion of H and C

265

composing the mineral (depending on the degree of hydration). For non-porous phases,

266

the theoretical Σ (oxide) wt% was calculated from the theoretical unit formula [7]. As

267

exemplified in Table 3, it was possible to discriminate the anhydrous clinker grains (Σ

268

(oxide) wt% = 100 %) from portlandite (Σ (oxide) wt%=76 %). According to the

269

polymodal distribution in the frequency histogram of Σ (oxide) wt% (Figure 4), the

270

transition between anhydrous phases and hydrates corresponded to a Σ (oxide) wt%

271

value of approximately 82 %. Considering the Gaussian distribution of Σ (oxide) wt% in

272

one phase, the minimum value of Σ (oxide) wt% equal to 82 % was considered for

273

thresholding all the pixels belonging to the unhydrated phases (Figure 4). Note that a

274

few pixels associated with a mix of hydrated and anhydrous phases were thus included

275

in this set of pixels.

276 277

Table 3 Example of Σ (oxide) wt% of unhydrated and hydrated cement phases used in the first segmentation step

AC C

EP

TE D

M AN U

SC

RI PT

260

Phases

C2 S

Formula Ca Si O*

2CaOSiO2 46 16

Portlandite Ca(OH)2 278

(Ati)wt%

54

0

38 22

H2O Σ (oxide) wt% (not analyzed) 100 24

76

*Total calculated stoichiometric oxygen

14

ACCEPTED MANUSCRIPT The pixels of the mapped area with a Σ (oxide) wt% above 82 % were plotted in a

280

ternary Si-Ca-Al3FeMg diagram for concrete I (Figure 4). Pixels related to clinker and

281

two types of slag were superimposed on the BSE images to verify that the segmented

282

pixels corresponded to grains in the BSE image. Some clusters corresponding to a part

283

of the pixels of weakly hydrated phases with Σ (oxide) wt% just above 82 % were

284

detected on the scatterplot but were not thresholded at this step.

AC C

285

EP

TE D

M AN U

SC

RI PT

279

286 287 288 289 290 291 292 293

Figure 5 – First processing step for concrete II to discriminate the hydrated and unhydrated MK. The frequency histogram of the sum of wt% (Σ (oxide) wt%) elements revealed a bimodal composition (pink polygon) in the pixel group belonging to the hydrated and unhydrated MK grains on the Si-Ca-Al3FeMg projection of the whole image (upper left and central figures). The segmentation by limits, considering the Σ (oxide) wt%, discriminated the hydrated MK (blue polygon) from the unreacted MK (purple polygon) with a threshold at 82 %. Their composition was then plotted on the Si-Ca- Al3FeMg diagram and the associated pixels back-projected onto the BSE image.

294

The process of segmentation based on the Σ (oxide) wt% was applied to concrete II as

295

shown in Figure 5, in which the Si-Ca-Al3FeMg ternary diagram illustrates the chemical

15

ACCEPTED MANUSCRIPT composition distribution of the whole pixels in the analyzed area of concrete II. The

297

polygon drawn in pink on the ternary diagram selected a cluster of pixels corresponding

298

to the pure end-members of MK grains and the associated mixture with C-(A-)S-H

299

(pink area on the BSE image). The Σ (oxide)wt% frequency histogram computed from

300

this selected cluster of pixels evinced the presence of a bimodal distribution with

301

Gaussian distributions centered on 75 and 90 wt%. This bimodal distribution for this

302

narrow chemical composition field implied that two phases with different hydration

303

states exist and could be segmented. Only pixels with Σ (oxide) wt% above 82 % were

304

segmented and plotted into a Si-Ca-Al3FeMg ternary diagram (Figure 5). The selected

305

pixels were segmented and displayed on the BSE images as MK (purple). The

306

remaining pixels corresponding to so-called “hydrated MK” were plotted again in the

307

same projection and segmented with the same polygon selection (as MK) in blue. The

308

corresponding pixels were distributed on an external ring of the anhydrous MK grains

309

(Figure 5). The illustration of the pixels on the BSE images confirmed that the selected

310

pixels correspond to grains and not to a mixture with the non-segmented binder pixels.

311

3.3. Image segmentation step 2: aggregates and portlandite

312

This step was identical for both concretes so we only describe it for concrete I (Figure

313

6). The pixels remaining after the first step were plotted into a sequence of ternary

314

diagrams (Figure 6) with different chemical axes to discriminate the pixels with

315

different chemical compositions. This use of sequential projection was well suited for

316

discriminating pixel clusters corresponding to pure end-members of carbonates or

317

quartz. Quartz grains were segmented by considering the pixels included in the yellow

318

polygon at the Si2 coordinate. The carbonates were segmented in different steps to

319

distinguish calcite, dolomite and siderite. This distinction was assessed by extracting all

320

pixels with high calcium content and devoid of Si (black polygon) in the Si2-CaFeMg-

AC C

EP

TE D

M AN U

SC

RI PT

296

16

ACCEPTED MANUSCRIPT Al3S3 plot and plotting them into another scatterplot to differentiate calcium, iron and

322

magnesium content (Figure 6). One of the main drawbacks linked with this aggregate

323

segmentation was the superimposition of theoretical compositional fields of calcite

324

(CaCO3) and portlandite (CaOH2) because among the analyzed elements, they were

325

only composed of Ca. As the C, O and H contents were not analyzed, we could not

326

distinguish then by only considering the measured chemical composition. This issue

327

was solved by considering the Σ (oxide) wt%, as an additional constraint: portlandite

328

has a CaO weight percentage of 75 % whereas calcite has a CaO weight percentage of

329

56 %. The pixels corresponding to the chemical composition of calcite and portlandite

330

were plotted in a new scatterplot Ca – Si – Σ (oxide) wt%. We could not detect any

331

portlandite in either concrete, as all the pixels were only stretched along a line towards

332

C-(A-)S-H composition that corresponds to calcite grains at the boundaries of

333

aggregates (Figure 6). This observation is consistent with the low-pH concrete

334

composition (Table 1).

AC C

EP

TE D

M AN U

SC

RI PT

321

17

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

335

Figure 6 – Segmentation methodology by polygon selection in ternary plots for the pixels remaining after step I for concrete I. The colored polygons outline the selected pixels for each mineral back-projected on the mineral maps.

339

3.4. Image segmentation step 3: S-bearing phases and hydrated phases

340

Due to the small number of pixels corresponding to S-bearing phases, all these phases

341

were differentiated according to their sulfur, calcium and aluminum content. The sulfur

342

frequency histogram (Figure 7) displayed progressive decrease in pixel frequency for

343

values between 0 and 0.11 sulfur atoms for every four oxygen atoms. This feature was

344

assigned to the background level at the position of the S Kα peak and corresponds to

345

pixels devoid of sulfur [6].

AC C

EP

TE D

336 337 338

18

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

346

Figure 7 – Thresholding methodology by polygon selection in ternary plots for the pixels remaining after step II for concrete I. The colored polygons outline the selected pixels for each mineral back-projected on the mineral maps.

350

The zone observed on the histogram having low pixel density with more than 0.1 sulfur

351

atom per 4 oxygen atoms, was assigned to the presence of S-bearing phases mixed with

352

different amounts of the surrounding hydraulic binder, although they could have been

353

related to sulfate in the C-S-H [26]. The selected pixels were then plotted on a S-Ca-Al

354

ternary plot for discriminating and mapping pyrite (FeS2), gypsum (CaSO4, not detected

355

here), monosulfate (Ca4(Al, Fe)2SO10 12 H2O) and AFt (Ca6(Al, Fe)2(SO4)3(OH)12 26

356

H2O) (Figure 7). Next, the remaining pixels were used spread out on a series of ternary

357

scatterplots in order to identify clusters of pixels related to hydrated grains.

358

Hydrogarnets (Ca3(Al, Fe)2(OH)12), zeolites (NaAlSi2O6 H2O) and hydrates formed by

359

the pozzolanic reaction of the silica fume with the calcium hydroxide (CaOH2) were

360

segmented in a Si2–CaFe–Al3 diagram. These minerals were segmented by successive

AC C

EP

TE D

347 348 349

19

ACCEPTED MANUSCRIPT polygons outlining the pixels that can be attributed to pure end-members or mixtures

362

between these pure end-members and C-(A-)S-H. The pixels attributed to the silica

363

fume hydrates spread out between the silica and C-(A-)S-H domain, while the pixels

364

assigned to the hydrogarnet domain spread out between the C-(A-)S-H domain and the

365

theoretical hydrogarnet solid solution domain. The selected pixels were plotted in a Ca-

366

AlSi-Fe diagram for discriminating the aluminous hydrogarnets from the ferrous

367

hydrogarnets. The aluminous hydrogarnets displayed compositions ranging from pure

368

Al-hydrogarnet (Ca3Al2(OH)12), devoid of Si, to katoite (Ca3Al2SiO4(OH)8) end-

369

members along the Ca and Al+Si solid solution line. The iron-rich hydrogarnet cluster

370

position in the Ca-AlSi-Fe diagram (Figure 7) indicated that this phase had an

371

intermediate composition in the composition range of the hydrogrossular –

372

hydroandradite solid solution [27]. The remaining pixels matched well with the

373

theoretical chemical composition domain of the C-(A-)S-H matrix [28] in the Si2–

374

CaFe–Al3 diagram (Figure 7). The associated cluster was isotropic without stretching

375

along a mixture line towards Al3 end-members of the scatterplot where the hydrotalcite

376

was located.

377

The pixels, which were previously identified as the C-(A-)S-H matrix, were plotted in a

378

Ca-Al-Mg6 diagram to inspect the variation of the chemical composition of the matrix

379

(Figure 7). In the Ca-Al-Mg6 plot, the pixels were distributed along a Ca-Mg line spread

380

between C-(A-)S-H and the hydrotalcite end-members. Two clusters were

381

distinguished. The upper gray polygon corresponded to pixels with the chemical

382

composition of C-(A-)S-H. The lower light blue polygon exhibited Mg-rich pixels

383

organized mainly around some of the unhydrated slag grains, where high fractions of

384

Mg were initially present. At first sight, we might interpret such spreading of

385

compositions along a line as a mixture including hydrotalcite and C-(A-)S-H

AC C

EP

TE D

M AN U

SC

RI PT

361

20

ACCEPTED MANUSCRIPT nanometric crystals. However, no mixture line between C-(A-)S-H and hydrotalcite

387

(Mg6Al2CO3OH16) end-member positions was detected in the previous Si2–CaFe–Al3

388

diagram. This attests to the presence of Mg-rich C-(A-)S-H instead of the presence of

389

hydrotalcite mixed with C-(A-)S-H.

390

4. Discussion

391

4.1. Quantitative spatial distribution of the minerals and phases

392

We

393

simultaneously using only one technique. The ternary projection tool (especially the Si-

394

Ca-Al3FeMg scatterplot) provided a direct overview of the chemistry and how it

395

evolved according to hydration state or the type of hardened cement materials (Figure

396

2). The different phases corresponded either to contrasted isotropic clusters or to

397

mixture/solid solution lines when plotted in these scatterplots. Since the pixel frequency

398

related to each cluster of pixels was encoded on a logarithmic color scale, we could also

399

qualitatively deduce the relative modal content of each cluster/mineral from visual

400

inspection of the scatterplot. For example, C2S and hydrogarnet (AlVI bearing mineral)

401

pixel clusters on the ternary projection were clearly less dense for concrete I than for

402

concrete II (Figure 2). Conversely, the cluster intensity corresponding to silica fume

403

hydrates was less intense for concrete II than for concrete I. This pixel distribution

404

related to the proportion of each phase over the ternary projection (Figure 2) is

405

consistent with the initial concrete compositions (Table 1). In the concrete II, only 20 %

406

of the clinker was substituted by SCMs, expecting to have more residual clinker phase.

407

As in the concrete I, 50 % of the clinker was substituted by slag, which are more

408

detected (e.g., higher wt%, table 4) in the hydraulic binder than in the concrete II. The

409

same observation could be done for the hydrated silica fume, the aluminous hydrate, for

mineral

types

and

associated

calculated

structural

formula

AC C

EP

TE D

M AN U

SC

quantified

RI PT

386

21

ACCEPTED MANUSCRIPT which the pixel distribution intensities (Figure 2) are consistent with the cement

411

compositions (Table 1).

412

4.2. Estimating hydrated phase compositions

413

The statistical composition of the C-(A-)S-H binder was obtained from a relatively large

414

sample volume including C/S and A/S ratios. In this study, the C/S of the two concretes

415

was 0.7 and 1.1 and the A/S was 0.06 and 0.2 for concretes I and II, respectively. These

416

C/S ratios were consistent with the information obtained by solid-state NMR data

417

(Figure 3) and the calculated Q2/Q1 ratio measured from the C-S-H chemical shift [16].

418

As expected, the calculated cationic formula showed that alkalis were present initially in

419

the cement and SCMs (slag, SF, MK). In the hardened concrete, alkalis were found in

420

the different hydrates, mainly in the hydrated silica fume and the C-(A-)S-H, which is

421

the main phase of the cement paste. As a consequence, local K/Na concentrations were

422

low and analyzing tiny changes in concentration for these elements was not relevant

423

since X-ray background intensities were not accounted for in the mapping mode (Table

424

4). Moreover, many of the hydrates (e.g. C-(A-)S-H or hydrogarnet) were in fact solid

425

solutions ranging between different end-members. The existence of such solid solutions

426

implied that it would have been difficult to quantify them by conventional methods such

427

as XRD or NMR because of lattice parameters evolving with chemical composition,

428

whereas mineral maps can discriminate the composition of such solid solutions. Two

429

types of C-(A-)S-H and solid solutions between hydrogarnet – katoite – stratlingite

430

domains were identified in concrete I. The chemical composition obtained for each

431

phase gave information on the composition of some poorly characterized phases such as

432

TAH, which was abundant in concrete II, with

433

mineral mapping results supported the idea that TAH could be considered as a solid

434

solution within the solid solution domain of aluminum-rich hydrogarnet – katoite, in

AC C

EP

TE D

M AN U

SC

RI PT

410

27

Al solid-state NMR (figure 3). Our

22

ACCEPTED MANUSCRIPT agreement with previous evidence based on chemical analyses of cement materials [21,

436

23].

437

4.3. Modal content of mineral and phases in concretes I and II

438

The wt(min) % for all the minerals in the hydrated cement paste obtained from the

439

mineral map was calculated from the spatial distribution of the minerals and their

440

measured modal composition (Table 4). This modal composition is only applied to the

441

map area and could be considered representative of the concretes’ compositions only if

442

this area was representative of the average concrete composition. The modal

443

composition is given according the whole composition (including the aggregates) and

444

normalized to the hydraulic binder composition (without considering the aggregates).

445

Obviously, considering the resolution of the map (1024 µm x 1024 µm) and considering

446

the size of the aggregates (> 4 000 µm), the mapped mineralogy (Table 4, mapped area)

447

is not representative of the whole concrete. The map was focused on the hydraulic

448

binder and the constitutive hydrates gel and the residual cement grains. The normalized

449

mass fraction was thus given (Table 4, normalized) without considering the aggregates.

450

15.4 wt% (14.5 wt% of slag) of residual cement grains were detected in the binder of

451

the concrete I, while only 6.4 wt% (3.3 wt% of C2S) were detected in the binder of the

452

concrete II. This proportion of anhydrous cement grains is coherent in view of the

453

percentage of clinker substitution and of the hydration kinetics of the slag.

454 455

Table 4 Weight (mineral) % of the whole analyzed area for concretes I and II and associated calculated cationic formula obtained from the mineral maps

AC C

EP

TE D

M AN U

SC

RI PT

435

Mineral

bind er

Aggregates

Concrete I Calcite Dolomite Siderite Quartz Belite Slag

Wt (min)%

Wt (min)%

Mapped area

Normalized

34.9 2.0 0.1 1.0 0.5 2.5

0.9 4.1

Calculated cationic formula CaCO3 Ca0.6Mg0.4CO3 FeCO3 SiO2 Ca2SiAl0.05O4 Na0.01K0.02Mg0.43Ca1.28Si1.1Al0.52O5

23

Na0.01K0.02Mg0.25Ca1.48Si1.18Al0.44O5 FeS2 Ca4Al2SO10:12H2O Ca3Al1Fe0.5Si0.5O6 :6H2O Ca3Al1.2Mg0.2Si0.5O8 :4H2O CaAl2Si4O12: 5.8H2O K0.01Ca0.34Si1.0Al0.02O2.4:1.4H2O Na0.01K0.02Ca0.56Mg0.15Si0.86Al0.16O2.8:2H2O

46.0

Na0.01K0.02Ca0.65Mg0.03Si0.94Al0.06O2.8:2.1H2O

100.0

CaCO3 Ca0.6Mg0.4CO3 FeCO3 SiO2 Ca2Si0.9Al0.05O4 Na0.01K0.02Mg0.25Ca1.5Si1.12Al0.36O5 K0.06Ca0.36Al1.5Si2.1O7 Ca6Al1.5S0.6O18:31H2O Ca3Al1.2Fe0.2Si0.2O6 :6H2O K0.03Ca0.58Si0.8Al0.2Fe0.03O2.8:1.8H2O K0.01Ca0.44Si0.95Al0.05O2.6:1.3H2O K0.02Ca0.94 Mg0.01Si0.84Al0.16O3.1:2.3H2O

SC

Normalized

RI PT

10.4 0.1 0.2 2.2 1.5 0.6 16.8 17.2

3.3 0.3 2.8 0.1 17.6 5.4 3.6 66.9

M AN U

Slag 6.4 Pyrite 0.1 AFm/Aft 0.1 Fe-Hydrogarnet 1.4 Al-Hydrogarnet 0.9 Ca zeolite 0.4 Hydrated silica fume 10.4 Mg-rich C-(A)-S-H 10.7 (light blue) C-(A)-S-H (gray) 28.6 Sum 100.0 Sum hydraulic binder 62 Mapped area Concrete II Calcite 40.4 Dolomite 2.6 Siderite 0.5 Quartz 5.2 Belite 1.7 Slag 0.2 Metakaolin 1.4 AFm/Aft 0.07 Al-Hydrogarnet 9.0 Hydrated metakaolin 2.8 Hydrated silica fume 1.8 C-(A)-S-H 34.3 Sum 100.0 Sum hydraulic binder 51.3

TE D

binder

Aggregates

ACCEPTED MANUSCRIPT

100.0

In particular, the C/S ratio of C-(A-)S-H determined for both concretes by quantitative

457

mapping in Table 4 is remarkably in accordance with the C/S ratio determined by 29Si

458

NMR as seen in section 3.1. That confirms the expected C/S ratio of C-(A-)S-H for

459

concrete

460

Moreover the relative proportion of hydrated silica fume and aluminous hydrates is also

461

coherent with the clinker and SCMs proportion given in the table 1 and the chemical

462

composition given in the table 2.

463

4.4. Spatial distribution of porosity

464

The nanometer scale microstructure of cement materials depends on the nanometer

465

scale porosity of the C-(A-)S-H gel phases [29]. Spatial resolution of BSE and X-ray

466

maps (2 µm in the present work) are not sufficient to resolve the nanometer pore

and II regarding their formulation and binder chemical composition.

AC C

I

EP

456

24

ACCEPTED MANUSCRIPT domain of this gel porosity. Nonetheless, the C-(A-)S-H gel volume should be corrected

468

for the presence of a porosity with a range of gel pore size. As the capillary porosity

469

(2.6 % and 5.9 % for concretes I and II, respectively), detected on the mapped area, was

470

smaller than the total bulk porosity of samples (16 ± 1 % and 13 ± 1 % for concrete I

471

and II, respectively); the undetected pore volume was probably associated to the

472

hydration products gel [29].

473

The gel porosity, ε, associated to the C-(A-)S-H gel phase was thus calculated from the

474

method developed by [7], from the difference between the theoretical sum of oxides

475

weight concentration Σtheo (oxides)wt% constitutive of C-(A-)S-H gel phases (~95 %,

476

table 5) and the value measured by EPMA that was actually measured (~70 %) by

477

applying the following formula for each of the pixels corresponding to the C-(A-)S-H

478

matrix:

SC

(

%) × ( %)

M AN U

=

with

(

%) =

TE D

479

RI PT

467

∑ "#$( ∑ '"((

!)%&% !)%&%

× 100

where ρr is the density of the resin that had polymerized in the pores, and ρm is the

481

mineral grain density (table 5). This method allows calculating the phase weight

482

proportion m(wt%) in the X-ray emission volume and to calculate the mean total

483

porosity (ε) within a X-ray volume and according to the spatial resolution (2 µm). This

484

method was thus adapted for cement materials to calculate the gel porosity.

485 486

Table 5 – Example of grain density and theoretical sum of oxide weight percentages Σtheo (oxides)wt% used in the porosity calculation.

AC C

EP

480

Mineral/phase Calcite Dolomite Quartz/Am. Si C-S-H0.6

Chemical composition CaCO3 CaMg(CO3)2 SiO2 Ca0.6SiO2.53(OH)0.55*

Σtheo (oxides) wt% 56 52 100 95

Grain density (g/cm3) 2.71 2.84 2.65 2.13* 25

ACCEPTED MANUSCRIPT Ca0.8SiO2.58(OH)0.57* Ca1.2SiO3.0(OH)0.73* Ca6Al2(SO4)3(OH)12:26H2O 3CaOSiO2 2CaOSiO2 Ca(OH)2

95 95 87 100 100 75

2.22* 2.42* 1.78 3.28 3.15 2.26

RI PT

487

C-S-H0.8 C-S-H1.2 Ettringite Alite Belite Portlandite * data from [30]

Using this method, we estimated the mean total porosity associated with each pixel, and

489

thus the intrinsic porosity of each phase, providing useful information on the porosity of

490

the domains prone to host diffusive processes [31]. The approach made it possible to

491

show that the largest part of the porosity was associated to C-(A-)S-H gel phases with

492

an intrinsic porosity of ~45 % (Figure 8 and Table 5). This gel porosity of 45 %, needs

493

to be linked to the average total porosity of the concrete (16 and 13 % ± 1), considering

494

the mass fraction of C-(A-)S-H gel phase into the concrete. In the present study, 75-80

495

% of the binder (which represent only 17-25% of the whole concrete composition, table

496

1) are considered to be C-(A-)S-H gel phases, whether 12-20% of C-(A-)S-H in a

497

concrete. Thus, porosity could be calculated considering the gel porosity weighted to

498

the mass fraction of gel phases, highlighting the need to consider the presence of

499

heterogeneous domains for the simulation of transport and reactivity processes in these

500

materials.

AC C

EP

TE D

M AN U

SC

488

501

26

ACCEPTED MANUSCRIPT Figure 8 – BSE images of the mapped area (A) and porosity map computed from the sum of oxide of each pixel (B).

504

4.5. Visualizing and identifying the hydration process

M AN U

SC

RI PT

502 503

505

Figure 9 – Mineral map of the analyzed area of concrete I. Zoom at the bottom of the maps illustrate the hydration process (in green and light blue) of the cement component (slag, SF)

509

From the mineralogical mapping we could discriminate the anhydrous phases and the

510

different hydrates and display the hydration process of some pozzolanic reactions

511

(Figures 9 and 10 for concretes I and II respectively).

512

In Figure 9, areas were magnified as indicated by the black squares to illustrate the

513

hydration process of the pozzolanic reactions. In concrete I, the light blue pixels (i.e.

514

Mg-rich C-(A-)S-H) revealed the heterogeneous distribution of the hydration reaction of

515

slag. The dark green pixels were associated with the pozzolanic reaction of silica fume.

516

The spatial distribution of the minerals detected showed a limited dispersion of hydrated

517

silica fume, which was mainly present in the form of clusters with size ranging from 10

518

µm to 100 µm. In concrete II, metakaolin particles (dark pink clusters in Figure 10),

519

which had the same grey level as the others hydrated phases (Figure 10, BSE image),

520

were identified with their contrasted chemical compositions (Figure 10). The grains

AC C

EP

TE D

506 507 508

27

ACCEPTED MANUSCRIPT were still present, with chemical compositions varying according to their hydration state

522

(dark blue ring in Figure 10).

523

In addition to the advantage of displaying spatial distribution of hydration reactions, the

524

segmentation by limits, based on the Σ (oxide) wt%, offered the possibility to quantify

525

the different anhydrous phases (cement components and SCMs) with their associated

526

chemical composition. The reaction pathways and changes in the chemistry of the

527

different anhydrous phases were identified, such as the hydration of the SF, slag and

528

MK.

529

EP

TE D

M AN U

SC

RI PT

521

Figure 10 – Mineral map of the analyzed area of concrete II. Zoom at the bottom of the maps illustrate the hydration process (in dark blue) of the cement component (MK)

532

In the case of concrete I, Mg-rich C-(A-)S-H identified in light blue in the map,

533

originated from the slow hydration of the slag. This phase with a C/S ratio of 0.65 had

534

an intermediate composition between the Mg-rich slag and the C-(A-)S-H identified in

535

gray with a C/S ratio of 0.7 (Figure 9). This evolution occurred with a release of Mg

536

(0.12 mol%) and Al (0.08 mol%), and Ca enrichment (0.11 mol%). For concrete II,

537

hydration of MK towards C-S-H was a stepwise process characterized by an

AC C

530 531

28

ACCEPTED MANUSCRIPT intermediate so-called hydrated MK phase whose composition domain was located in

539

the same region as the Mg-rich C-(A-)S-H of concrete I. Basically, the hydrated MK

540

and Mg-rich C-(A-)S-H compositions differed only by their contrasted Al, Fe and Mg

541

contents (table 4, (Al0.2Fe0.05)/O2.8 and (Al0.15Mg0.15)/O2.8, respectively).

542

The mineral maps indicated that the remaining SF was present in the hydraulic binder in

543

the form of clusters with sizes ranging from 10 to 100 µm (Figure 9). This observation

544

was also verified for the MK in concrete II and illustrated the low SCM dispersions

545

during concrete mixing. Local external layer of hydrated products were detected in the

546

boundaries of the coarser, and still unhydrated, SF and MK grains.

547

5. Conclusions

548

We proposed a method, adapted from [6], that displays quantitative spatial phase

549

distribution and the porosity constituting cement materials. While the spatial resolution

550

of the method (2 µm) could be considered a problem in view of the nanometric size of

551

the hydrated cement phases, we showed that the method offered could provide a full

552

mineralogy with associated phase compositions to supply geochemical modelling and to

553

provide better understanding of the hydration process governing physical and chemical

554

properties in cement materials. It yielded a straightforward discrimination of the

555

anhydrous phases and the various hydrated phases and gave a spatial resolution of the

556

hydration process in pozzolans over different observation scales. Two concretes with

557

different SCMs and environmental conditions (hydration condition) were investigated.

558

The hydration of concrete I, based originally on PC, SF and slag, resulted in the

559

formation of C-(A-)S-H mainly with low C/S and A/S, while for concrete II, based on

560

PC, SF and MK, the C-(A-)S-H exhibited a higher C/S and A/S. C-S-H characterization

561

by the quantitative mineralogical mapping was successfully correlated with

562

measurements in case of both concretes. In addition to the quantitative mineralogy, our

AC C

EP

TE D

M AN U

SC

RI PT

538

29

Si NMR

29

ACCEPTED MANUSCRIPT characterization method made it possible to visualize the spatial phase distribution. In

564

the case of concrete I, we highlighted the presence of clusters of hydrated silica fume.

565

For the pozzolan reactions, we further addressed the hydration rate by quantifying the

566

area corresponding to residual anhydrous grains compared to hydrated grains with the

567

same chemical composition. In a further step, by estimating the reactive surface areas

568

for each phase it should be possible to obtain realistic reactive surface areas for each

569

phase and mineral, following procedures established for other materials [32, 33]. This

570

could provide the kinetic parameters for reactive transport modeling approaches [34-

571

38].

M AN U

SC

RI PT

563

Acknowledgements

573

The results presented in this article were collected during the GL-ESC, MLH and UP-

574

Transfert projects granted by Andra as part of the Andra/BRGM scientific partnership.

575

We gratefully acknowledge contributions from Yannick Linard, Xavier Bourbon and

576

Guillaume Wille.

[1] A. Neville, The confused world of sulfate attack on concrete, Cement and Concrete Research 34(8) (2004) 1275-1296. [2] T. Schmidt, B. Lothenbach, M. Romer, J. Neuenschwander, K. Scrivener, Physical and microstructural aspects of sulfate attack on ordinary and limestone blended Portland cements, Cement and Concrete Research 39(12) (2009) 1111-1121. [3] S. Ahmad, Reinforcement corrosion in concrete structures, its monitoring and service life prediction––a review, Cement and Concrete Composites 25(4–5) (2003) 459-471. [4] T. Chappex, K. Scrivener, Alkali fixation of C–S–H in blended cement pastes and its relation to alkali silica reaction, Cement and Concrete Research 42(8) (2012) 1049-1054. [5] V.G. Papadakis, Effect of supplementary cementing materials on concrete resistance against carbonation and chloride ingress, Cement and Concrete Research 30(2) (2000) 291299. [6] D. Pret, S. Sammartino, D. Beaufort, A. Meunier, M. Fialin, L.J. Michot, A new method for quantitative petrography based on image processing of chemical element maps: Part I. Mineral mapping applied to compacted bentonites, American Mineralogist 95(10) (2010) 1379-1388. [7] D. Pret, S. Sammartino, D. Beaufort, M. Fialin, P. Sardini, P. Cosenza, A. Meunier, A new method for quantitative petrography based on image processing of chemical element maps: Part II. Semi-quantitative porosity maps superimposed on mineral maps., American Mineralogist 95 (2010) 1389-1398.

EP

578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596

References

AC C

577

TE D

572

30

ACCEPTED MANUSCRIPT

EP

TE D

M AN U

SC

RI PT

[8] S. Gaboreau, D. Pret, E. Tinseau, F. Claret, D. Pellegrini, D. Stammose, 15 years of in situ cement-argillite interaction from Tournemire URL: Characterisation of the multi-scale spatial heterogeneities of pore space evolution, Applied Geochemistry 26(12) (2011) 2159-2171. [9] D. Massiot, B. Touzo, D. Trumeau, J.P. Coutures, J. Virlet, P. Florian, P.J. Grandinetti, Twodimensional magic-angle spinning isotropic reconstruction sequences for quadrupolar nuclei, Solid State Nuclear Magnetic Resonance 6(1) (1996) 73-83. [10] D. Massiot, F. Fayon, M. Capron, I. King, S. Le Calve, B. Alonso, J.O. Durand, B. Bujoli, Z.H. Gan, G. Hoatson, Modelling one- and two-dimensional solid-state NMR spectra, Magnetic Resonance in Chemistry 40(1) (2002) 70-76. [11] S. Gaboreau, C. Lerouge, S. Dewonck, Y. Linard, X. Bourbon, C.I. Fialips, A. Mazurier, D. Pret, D. Borschneck, V. Montouillout, E.C. Gaucher, F. Claret, In-Situ Interaction of Cement Paste and Shotcrete with Claystones in a Deep Disposal Context, American Journal of Science 312(3) (2012) 314-356. [12] P. Sardini, A. El Albani, D. Pret, S. Gaboreau, M. Siitari-Kauppi, D. Beaufort, Mapping and Quantifying the Clay Aggregate Microporosity in Medium- to Coarse-Grained Sandstones Using the C-14-Pmma Method, Journal of Sedimentary Research 79(7-8) (2009) 584-592. [13] D. Pret, P. Sardini, D. Beaufort, R. Zellagui, S. Sammartino, Porosity distribution in a clay gouge by image processing of C-14-PolyMethylMethAcrylate (C-14-PMMA) autoradiographs: Case study of the fault of St. Julien (Basin of Lodeve, France), Applied Clay Science 27(1-2) (2004) 107-118. [14] J.I. Goldstein, D.E. Newbury, P. Echlin, D.C. Joy, A.D.J. Romig, C.E. Lyman, C. Fiori, E. Lifshin, Scanning Electron Microscopy and X-ray Microanalysis, 2nd Edition Plenum Press, New York, NY, 1992. [15] C. Merlet, An Accurate Computer Correction Program for Quantitative Electron-Probe Microanalysis, Mikrochim Acta 114 (1994) 363-376. [16] X. Cong, R.J. Kirkpatrick, 29Si MAS NMR study of the structure of calcium silicate hydrate, Advanced Cement Based Materials 3(3–4) (1996) 144-156. [17] P. Colombet, A.-R. Grimmer, H. Zanni, P. Sozzani, Nuclear Magnetic Resonance Spectroscopy of Cement Based Materials, Springer-Verlag Berlin Heidelberg1998. [18] A. Nonat, The structure and stoichiometry of C-S-H, Cement and Concrete Research 34(9) (2004) 1521-1528. [19] X. Ke, S.A. Bernal, J.L. Provis, Controlling the reaction kinetics of sodium carbonateactivated slag cements using calcined layered double hydroxides, Cement and Concrete Research 81 (2016) 24-37. [20] I.G. Richardson, A.R. Brough, R. Brydson, G.W. Groves, C.M. Dobson, Location of Aluminum in Substituted Calcium Silicate Hydrate (C-S-H) Gels as Determined by 29Si and 27Al NMR and EELS, Journal of the American Ceramic Society 76(9) (1993) 2285-2288. [21] M.D. Andersen, H.J. Jakobsen, J. Skibsted, A new aluminium-hydrate species in hydrated Portland cements characterized by 27Al and 29Si MAS NMR spectroscopy, Cement and Concrete Research 36(1) (2006) 3-17. [22] B. Lothenbach, G. Le Saout, E. Gallucci, K. Scrivener, Influence of limestone on the hydration of Portland cements, Cement and Concrete Research 38(6) (2008) 848-860. [23] R.J. Myers, S.A. Bernal, R. San Nicolas, J.L. Provis, Generalized Structural Description of Calcium-Sodium Aluminosilicate Hydrate Gels: The Cross-Linked Substituted Tobermorite Model, Langmuir 29(17) (2013) 5294-5306. [24] M.D. Andersen, H.J. Jakobsen, J. Skibsted, Incorporation of aluminium in the calcium silicate hydrate (C-S-H) of hydrated portland cements: a high field 27Al and 29Si MAS NMR investigation, Inorganic Chemistry 42 (2003) 2280-2287. [25] B. Fabbri, S. Gualtieri, C. Leonardi, Modifications induced by the thermal treatment of kaolin and determination of reactivity of metakaolin, Applied Clay Science 73 (2013) 2-10. [26] R. Barbarulo, H. Peycelon, S. Leclercq, Chemical equilibria between C-S-H and ettringite, at 20 and 85 degrees C, Cement and Concrete Research 37(8) (2007) 1176-1181.

AC C

597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648

31

ACCEPTED MANUSCRIPT

RI PT

SC

M AN U

TE D

EP

684 685

[27] B.Z. Dilnesa, B. Lothenbach, G. Renaudin, A. Wichser, D. Kulik, Synthesis and characterization of hydrogarnet Ca3(AlxFe1 − x)2(SiO4)y(OH)4(3 − y), Cement and Concrete Research 59 (2014) 96-111. [28] I.G. Richardson, The calcium silicate hydrates, Cement and Concrete Research 38(2) (2008) 137-158. [29] H.M. Jennings, J.J. Thomas, J.S. Gevrenov, G. Constantinides, F.J. Ulm, A multi-technique investigation of the nanoporosity of cement paste, Cement and Concrete Research 37(3) (2007) 329-336. [30] C. Roosz, S. Gaboreau, S. Grangeon, D. Prêt, V. Montouillout, N. Maubec, S. Ory, P. Blanc, P. Vieillard, P. Henocq, Distribution of water in synthetic calcium silicate hydrates, Langmuir (2016). [31] Y. Elakneswaran, A. Iwasa, T. Nawa, T. Sato, K. Kurumisawa, Ion-cement hydrate interactions govern multi-ionic transport model for cementitious materials, Cement and Concrete Research 40(12) (2010) 1756-1765. [32] C.I. Steefel, L.E. Beckingham, G. Landrot, Micro-Continuum Approaches for Modeling PoreScale Geochemical Processes, Reviews in Mineralogy and Geochemistry 80(1) (2015) 217-246. [33] G. Landrot, J.B. Ajo-Franklin, L. Yang, S. Cabrini, C.I. Steefel, Measurement of accessible reactive surface area in a sandstone, with application to CO2 mineralization, Chemical Geology 318–319 (2012) 113-125. [34] N.C.M. Marty, C. Tournassat, A. Burnol, E. Giffaut, E.C. Gaucher, Influence of reaction kinetics and mesh refinement on the numerical modelling of concrete/clay interactions, Journal of Hydrology 364(1-2) (2009) 58-72. [35] S. Molins, D. Trebotich, L. Yang, J.B. Ajo-Franklin, T.J. Ligocki, C. Shen, C.I. Steefel, PoreScale Controls on Calcite Dissolution Rates from Flow-through Laboratory and Numerical Experiments, Environmental Science & Technology 48(13) (2014) 7453-7460. [36] L.E. Beckingham, E.H. Mitnick, C.I. Steefel, S. Zhang, M. Voltolini, A.M. Swift, L. Yang, D.R. Cole, J.M. Sheets, J.B. Ajo-Franklin, D.J. DePaolo, S. Mito, Z. Xue, Evaluation of mineral reactive surface area estimates for prediction of reactivity of a multi-mineral sediment, Geochimica et Cosmochimica Acta 188 (2016) 310-329. [37] N.C.M. Marty, I. Munier, E.C. Gaucher, C. Tournassat, S. Gaboreau, C.Q. Vong, E. Giffaut, B. Cochepin, F. Claret, Simulation of Cement/Clay Interactions: Feedback on the Increasing Complexity of Modelling Strategies, Transport Porous Med 104(2) (2014) 385-405. [38] N.C.M. Marty, F. Claret, A. Lassin, J. Tremosa, P. Blanc, B. Madé, E. Giffaut, B. Cochepin, C. Tournassat, A database of dissolution and precipitation rates for clay-rocks minerals, Applied Geochemistry 55 (2015) 108-118.

AC C

649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683

32