Subscriber access provided by - Access paid by the | UCSF Library
Article
Dissolution kinetics of cubic tricalcium aluminate measured by digital holographic microscopy Alexander Sebastian Brand, and Jeffery W. Bullard Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.7b02400 • Publication Date (Web): 25 Aug 2017 Downloaded from http://pubs.acs.org on September 4, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 30
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 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
Langmuir
1
Dissolution kinetics of cubic tricalcium aluminate measured by digital
2
holographic microscopy
3 4
Alexander S. Brand1* and Jeffrey W. Bullard1
5 6 7
1
Materials and Structural Systems Division, National Institute of Standards and Technology, Gaithersburg,
Maryland 20899, USA; Electronic addresses:
[email protected] and
[email protected] *
Corresponding author
8 9
ABSTRACT
10
In situ digital holographic microscopy is used to characterize the dissolution flux of
11
polycrystalline cubic tricalcium aluminate (C3A-c). The surface dissolves at rates that vary
12
considerably with time and spatial location. This implies a statistical distribution of fluxes, but an
13
approximately steady-state median rate was obtained by using flowing solutions and by reducing
14
the water activity in the solution. The dissolution flux from highly crystalline C3A-c depends on
15
the water activity raised to an empirically-derived exponent of 5.2, and extrapolates to a median
16
flux of −2.1 µmol m−2 s−1 in pure water with an interquartile range of 3.2 µmol m−2 s−1. The flux
17
from a less crystalline source of C3A-c has an empirical water activity exponent of 4.6 and an
18
extrapolated median flux of only −1.4 µmol m−2 s−1 in pure water with an interquartile range of
19
1.9 µmol m−2 s−1. These data suggest that the bulk dissolution rate of C3A-c can vary by at least
20
30 % from one source to another and that variability in the local rate within a single material is
21
even greater because of the heterogeneous spatial distribution of structural characteristics (i.e.,
22
degree of crystallinity, chemical impurities, and defects).
23
Keywords: digital holographic microscopy, tricalcium aluminate, dissolution, rate spectra,
24
surface reactivity
25 26 27
1. INTRODUCTION The ability to predict the rate of hydration and development of properties in cement
28
binders is a long-standing goal of concrete research. Computational materials science models of
29
cement hydration have demonstrated the ability to provide accurate calculations of engineering
30
properties based on the microstructural state of the binder. However, accurately predicting the
31
rate of change in microstructure has proven to be extremely difficult, in part because the initial 1
ACS Paragon Plus Environment
Langmuir
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 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
32
binder material is inherently variable in its composition and reactivity, but primarily because
33
there are significant gaps in a quantitative understanding of the basic mechanisms that govern the
34
reactions of the solid binder components with water. Any microstructural model rooted in the
35
basic physics and chemistry of materials will require input of the details of the binder’s starting
36
mineralogical composition, particle size distribution, and basic thermodynamic and kinetic
37
properties of the reactions being simulated, including the dissolution rate laws and constants of
38
individual cementitious minerals.1,2
39
Substantial progress has been made in characterizing the fundamental dissolution rates
40
and mechanisms of some minerals, especially calcite and gypsum, which are minority
41
components of portland cement. Much of the recent progress has involved monitoring nanoscale
42
changes in surface topography with vertical scanning interferometry (VSI) and atomic force
43
microscopy (AFM).3 Measurements of the dissolution rates of tricalcium silicate (Ca3SiO5, or
44
C3S)4 and dicalcium silicate (Ca2SiO4, or C2S)4 in various aqueous solutions also have been
45
reported in the last several years using bulk powder methods,5–8 and VSI observations of surface
46
topography changes have characterized the dissolution behavior of C3S9,10 and calcium silicate
47
hydrate (C–S–H).11 AFM has also been used to observe C–S–H growth on alite12 and to
48
characterize cement clinker hydration.13,14 Scanning electron microscopy (SEM) has also been
49
used to infer a dissolution rate of C3S via ex situ measurement of the dissolved surface area.15
50
More recently, holographic interferometry techniques have been used to study the dissolution
51
kinetics of hardened cement paste under acid attack,16 while digital holographic microscopy
52
(DHM) has been used to examine minerals important to cements.17,18
53
Tricalcium aluminate (Ca3Al2O6, or C3A)4 is a key component of portland cement clinker
54
and is commonly found in cubic form (C3A-c).19 It is usually present in much smaller
55
proportions than C3S and C2S, but it nevertheless has an important influence on early hydration
56
reactions, enthalpy release, and strength development. The reactivity of C3A with water is so
57
great that it can induce rapid precipitation of calcium aluminate hydrates that lead to flash setting
58
unless calcium sulfate is added to mitigate the rate of reaction.19 In fact, an important part of
59
concrete mixture design is the determination of the optimum mass fraction of calcium sulfate to
60
include to properly regulate C3A hydration.20 A wide range of techniques have been used to
61
characterize the C3A reactions in water, such as environmental scanning electron microscopy,21
62
X-ray transmission microscopy,22,23 in situ synchrotron X-ray diffraction,24–26 X-ray adsorption 2
ACS Paragon Plus Environment
Page 2 of 30
Page 3 of 30
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 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
Langmuir
63
spectroscopy,27 nuclear magnetic resonance,28 and solution ionic concentration measurement,29,30
64
but no measurements have reported the net dissolution flux from C3A in aqueous solutions.
65
Therefore, the objective of this study is to experimentally characterize the dissolution flux of
66
C3A and to estimate the form of the dissolution rate equation by in situ monitoring of the
67
nanoscale surface topography evolution in real time.
68 69
1.1 Dissolution Kinetics of C3A 18-
70
The crystal structure of C3A-c features closed rings of six aluminate tetrahedra, Al6 O18 ,
71
with the aluminum ions located near the corners of the cubic structure,19,31,32 and each aluminate
72
tetrahedron has two nonbridging oxygens (NBOs).31 The high electronegativity of the aluminum
73
relative to calcium gives a substantial covalent character to the Al−O bond, while the bonds
74
between the nine calcium atoms and the NBOs in the aluminate ring has more significant ionic
75
character.
76
As in most double oxides, like the other portland cement clinker phases, dissolution of
77
C3A involves an initial step of protonation of the apical NBOs of the aluminate tetrahedra.32–34 In
78
neutral or basic solutions, the proton can be donated by a water molecule, leaving behind a
79
hydroxyl ion. Considering a formula unit as three calcium atoms plus one third of the aluminate
80
ring, with the corresponding formula Ca3Al2O6, this elementary protonation step can be written 3Ca Al O + H O → 3Ca Al O OH + .
(1)
81
This step occurs repeatedly to protonate the other NBOs, and, on average, releases one calcium
82
atom for every two hydroxyls produced: 3Ca Al O OH + H O → 2Ca Al O (OH) + Ca + OH .
(2)
83
The fate of the “released” calcium ion is not immediately clear; it likely could be released from
84
the solid and immediately directly complexed by a hydration shell of six water molecules,35,36 Ca + 6 H O → Ca(H O)
,
(3)
85
which could escape into the bulk solution, remain within the electrical double layer near the
86
surface, or be adsorbed as an outer-sphere complex.27,37,38 Alternatively, as proposed by Mishra
87
et al.,32 each released calcium may condense with two liberated hydroxyl ions to form an
88
amorphous calcium hydroxide layer at the surface, Ca + 2 OH → Ca(OH) (am).
3
ACS Paragon Plus Environment
(4)
Langmuir
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 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
Page 4 of 30
89
Any such amorphous layer should be unstable with respect to dissociation, since the adjacent
90
solution is significantly undersaturated with respect to even crystalline calcium hydroxide
91
(portlandite), with saturation corresponding to about 20 mmol L−1 in pure water at 298 K.19
92
However, such a layer might be kinetically favored if the hydroxyls are locally available because
93
the full hydration of calcium requires coordination with six additional water molecules.
94
When NBOs of the aluminate tetrahedra have been protonated, further attack by the water
95
on the bridging oxygens can decouple one or more of the aluminate tetrahedra simultaneously
96
from the ring structure. For example, two water molecules could attack each of the two bridging
97
oxygens of one aluminate group to produce a fully protonated Al(OH) 4 ion and releasing the
98
remaining calcium atom that is ionically bonded to that part of the ring: Ca Al O (OH) + 2 H O → 2 Al(OH) + Ca .
99
(5)
As suggested by this discussion, the exact sequence of steps involved in C3A dissolution is
100
difficult to determine with certainty, and the ordering of the steps may even be subject to some
101
statistical fluctuation. Identifying a rate-controlling step is also difficult, although the superficial
102
protonation of the apical NBOs likely can be ruled out as being rate controlling. If one of the
103
steps could be identified as rate-controlling, then all the other steps before and after would be
104
near equilibrium, and the rate equation for dissolution could be written exactly according to
105
transition state theory.39,40
106 107
In the absence of more certain knowledge of the reaction sequence and the ratecontrolling step, the overall reaction for C3A still can be written as Ca Al O + 6 H O → 3 Ca + 2 Al(OH) + 4 OH ,
(6)
108
and one must be satisfied with an empirical rate equation for the overall process. Under the
109
experimental conditions used in this paper, where rapidly flowing water over the solid surface
110
removes the product ions as much as possible, the reaction should proceed almost entirely in the
111
forward direction and without precipitation of a secondary or intermediate solid, in which case
112
the empirical rate equation far from equilibrium is written as a power function of the reactant
113
concentrations or activities.39 For C3A the empirical rate equation is *
= !" # $%&' ( ) ,
(7)
114
where JC3A is the dissolution flux (mol m−2 s−1), k′+ is the overall rate constant (mol s−1), CC3A is
115
the concentration of reactive sites on the C3A surface (m−2), aH2O is the activity of water at the
116
surface, and n is the reaction order, which must be determined empirically. 4
ACS Paragon Plus Environment
Page 5 of 30
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 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
Langmuir
117
Writing the rate equation as in (7) presupposes that there is only one type of reactive site
118
at the surface and that its reactivity is the same everywhere it is found on the surface. However,
119
there are several types of reactive sites on a given crystalline surface. In addition, the actual
120
reactivity of a given site depends on its molecular identity, its local structural configuration (i.e.,
121
terrace, ledge, or kink site), and the effects of nearby structural defects that introduce lattice
122
strain energy, such as dislocation outcroppings, stacking faults, grain boundaries, vacancies, or
123
chemical impurities. Therefore, a distribution of site types and reactivity will typically present at
124
a surface.41
125
These atomic-scale structural details usually are not known in advance, nor can the
126
introduction of a single term like CC3A in the overall rate equation capture their influence. As an
127
alternative, one often measures the total surface area of the solid and assumes it to be
128
proportional to the surface concentration of reaction sites. The variability of sites and their
129
reactivities are then contained within a local rate constant. In that case, Eq. (7) is recast as *
= ! $%&' ( ) ,
(8)
130
where now k+ has units of molar flux. In bulk dissolution experiments using rotating disks or
131
dilute granular suspensions, one generally assumes that dissolution occurs uniformly over the
132
entire surface area, so that a globally averaged rate constant in recovered. However, topographic
133
methods like VSI, AFM, and DHM can track the surface retreat from point to point on a surface.
134
When these methods are used, the range of local dissolution fluxes can be measured, both in
135
terms of point to point flux variability on a given surface,17,18,41–48 and in terms of temporal flux
136
variability at a given point on a surface17,18 as defects are annihilated or uncovered by the
137
dissolution process.
138
This paper assumes that Eq. (8) empirically describes the dissolution flux from several
139
polycrystalline C3A surfaces prepared by solid state sintering. Local in situ dissolution rates are
140
tracked in real time by DHM using flowing solutions in which the water activity is
141
systematically varied by ethanol additions. The statistical distribution of local rate constants is
142
measured and related to the presence and evolution of surface defects. Both the spatially
143
averaged rate constant, k+, and the activity exponent, n, are estimated by regression of the flux
144
data to quantitatively determine the empirical rate equation.
145
5
ACS Paragon Plus Environment
Langmuir
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 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
146
2. EXPERIMENTAL SECTION
147
2.1 Materials
148
Two sources of laboratory-synthesized C3A, varying primarily in the abundance of
149
amorphous material, were evaluated in this study. The first source, termed hereafter “C3A #1”, in
150
this study was produced by Mineral Research Processing (Meyzieu, France).49 Elemental
151
analysis by inductively coupled plasma optical emission spectroscopy (ICP-OES) yielded mass
152
fractions (with standard deviations) of 64.6 % ± 1.2 % CaO and 35.4 % ± 0.6 % Al2O3, so the
153
molar ratio CaO:Al2O3 is 3.3. C3A #1 was provided as nodules, which were ground and
154
micronized prior to sample preparation.
155
The second source, termed hereafter “C3A #2”, was produced by Construction
156
Technology Laboratories (Skokie, Illinois, USA).49 This source was provided as a powder and
157
exhibited evidence of some prehydration (i.e., hydrogarnet and portlandite), so the powder was
158
heated to 500 °C to dehydrate prior to micronizing for sample preparation. Elemental analysis by
159
ICP-OES yielded mass fractions of 64.4 % ± 1.1 % CaO and 35.6 % ± 1.0 % Al2O3,
160
corresponding to the molar ratio CaO:Al2O3 = 3.3.
161
In keeping with some other topographic analyses of mineral surfaces,50 the nominally flat
162
sample surface required for DHM examination was produced by solid state sintering and
163
polishing of a cylindrical powder compact. The pellets were 13 mm in diameter and uniaxially
164
compressed at approximately 170 MPa, with a small volume of ethanol mixed in to act as a
165
compaction aid. The samples were sintered by heating at 10 °C min−1 to 1500 °C, holding
166
constant at 1500 °C for 8 h, and finally cooling at 2 °C min−1 to 30 °C. The sintered samples
167
were mounted onto titanium stubs and the surface was polished using silicon carbide (grits of
168
400, 600, 800, and 1200), diamond paste (3 µm and 0.25 µm), and alumina (0.05 µm) with
169
propylene glycol as lubricant.
170
Powder samples before and after sintering were micronized and examined with X-ray
171
diffraction (XRD) using rutile as an internal standard. Phase abundances determined by Rietveld
172
refinement are shown in Table 1. The compositions are consistent with other reports of synthetic
173
C3A,51,52 which typically contains 1 % to 2.5 % free lime and some amorphous content. The
174
calcite content of C3A #2 before sintering could have been produced by initial heat treatment at
175
500 °C to remove its prehydration products, which transformed the free lime content to calcite by
176
reaction with ambient CO2 in the furnace atmosphere. A stoichiometric balance implies that C3A 6
ACS Paragon Plus Environment
Page 6 of 30
Page 7 of 30
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 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
Langmuir
177
#2 had a free lime content of about 2.2 % before sintering. The higher-temperature sintering
178
treatment caused all the calcite to decompose back to free lime. Some mayenite (C12A7) was
179
found in both C3A sources but was either totally or substantially transformed by sintering in C3A
180
#1 and C3A #2, respectively. C3A #2 has more than twice the amorphous content of C3A #1 both
181
before and after sintering. No other crystalline calcium aluminate phases (e.g., orthorhombic
182
C3A, monocalcium aluminate) were detected by XRD.
183 184
Table 1. Mass Fraction Composition (%) of C3A by Rietveld Analysis with an Internal Standard C3A #1 C3A #2 Before Sintering After Sintering Before Sintering After Sintering C3A-c 89.7 ± 0.3 94.2 ± 0.3 74.5 ± 0.4 84.3 ± 0.3 Free Lime 2.2 ± 0.2 2.0 ± 0.2 N/A 1.0 ± 0.2 Calcite N/A N/A 4.0 ± 0.3 N/A Mayenite 4.3 ± 0.3 N/A 5.0 ± 0.3 4.6 ± 0.3 Amorphous Content 3.8 ± 0.7 3.8 ± 0.7 16.5 ± 0.7 10.1 ± 0.6 Note: the uncertainty represents the average standard error of regression from Rietveld refinement Phase
185 186 187
2.2 Digital Holographic Microscopy (DHM) The DHM utilized in this study was a Lyncée Tec Model R-2203 (Lausanne,
188
Switzerland).49 Additional instrument details are reported elsewhere,17,53 but briefly: A
189
monochromatic coherent light source (665.5651 nm wavelength) is split into two beams known
190
as the reference and object beams. The object beam is sent through an objective lens to the
191
specimen surface, reflects into the instrument, and is recombined with the reference beam. The
192
interference pattern of the recombined beams is recorded as a hologram for the entire field of
193
view. The holograms are collected at frame rates up to 25 s−1, thus allowing full-field, in situ data
194
to be collected in real time. Numerical reconstruction of each hologram yields amplitude and
195
phase data at every pixel.54–56 The reconstructed phase data are converted to surface elevation at
196
each image pixel, so this configuration and procedure enable the quantitative tracking of in situ
197
surface topography evolution. Similar approaches have already been reported in studies of
198
mineral dissolution17,18 and metal electrodeposition.57 Local dissolution (or growth) fluxes can be
199
computed by tracking the change in pixel height with time. The surface normal flux, ks (mol m−2
200
s−1), is determined3,58 by the quotient of the surface normal velocity, vs (m s−1), which is the per
201
pixel measured change in height over change in time, ∆h/∆t, and the known molar volume of the
202
sample, Vm (m3 mol−1):
7
ACS Paragon Plus Environment
Langmuir
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 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
!+ =
,+ ∆ℎ 1 = . -. ∆1 -.
Page 8 of 30
(9)
203
The surface normal flux, ks, in Eq. (9) is measured at each pixel and can then be equated to the
204
net dissolution flux, JC3A, in Eq. (8), thus linking the measured surface topography changes to the
205
unknown k+ and n. For sign convention, dissolution will be denoted as negative vs and negative
206
ks, while precipitation will be regarded as positive vs and positive ks. The molar volume of C3A-c
207
is 8.91×10−5 m3 mol−1.59 For this configuration, uncertainty in the phase measurement results in
208
approximately 0.41 nm (mean value) uncertainty in the height for data averaged from 20
209
holograms and a water flow rate of 33 mL min−1.53
210
DHM measurements of absolute height changes require a flat and inert surface reference
211
plane that remains at constant elevation during the experiment. For this purpose, a thin film of
212
chromium was introduced by physical vapor deposition to a typical thickness of about 70 nm to
213
100 nm, as shown in Figure 1a. The figure shows some residual porosity at the C3A surface that
214
survived the sintering treatment. The depth and poor reflectivity of the larger pores produce
215
locally noisy phase data, so they are digitally removed during DHM post-processing (Figure 1b)
216
using an image gradient thresholding. Note that only the large pores are removed by this method;
217
the smaller, shallower pores provide sufficient signal to permit their numerical reconstruction. In
218
general, sintered C3A #2 had lower surface porosity than C3A #1, which might be attributed to
219
the mayenite phase, which melts at 1400 °C and therefore could act as a liquid phase sintering
220
aid.
221
8
ACS Paragon Plus Environment
Page 9 of 30
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 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
Langmuir
(a)
(b)
222
Figure 1. (a) Greyscale phase map of a polished C3A #1 surface showing the selected area for
223
reference surface (red rectangle) and measurement area (blue rectangle). The pores appear as
224
sudden phase jumps and noisy data. (b) In post-processing, the pores are digitally removed from
225
the measurement area.
226 227
The reaction of C3A in pure water, either static or rapidly flowing across the surface,
228
caused extremely rapid and fine-scale surface roughening. Within seconds the surface became so
229
rough that light no longer reflected into the lens with sufficient intensity to enable a
230
measurement. The rate of these surface changes at each pixel was reduced by decreasing the
231
water activity by mixing deionized water (> 0.18 MΩ m) with 200 proof ethanol. The water
232
activity in the ethanol-water mixture was determined using the non-random two-liquid (NRTL)
233
model,60 which provides the activity coefficient (γ, mole fraction basis) as a function of the mole
234
fraction of each liquid phase (x, mol mol−1). The interaction energy parameters and the non-
235
randomness parameter in the NRTL model were estimated as recommended by Gmehling et al.61
236
The activity, % = 45, was then calculated for use in Eq. (8). Binary mixtures with ethanol
237
volume fractions of 80 % (aH2O = 0.689), 70 % (aH2O = 0.777), 60 % (aH2O = 0.824), and 50 %
238
(aH2O = 0.857) were used for this purpose.
239 240
All experiments were performed by affixing the polished C3A specimen to a titanium stub, which was locked into a chemically inert flow-through reaction cell that has been described 9
ACS Paragon Plus Environment
Langmuir
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 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
241
previously.17,18,53 A 20× immersion lens was used for all experiments. Unless otherwise noted,
242
every 10 s to 30 s sets of 15 to 20 holograms were collected at a rate of 12.5 s−1. The sets of 15 to
243
20 holograms were averaged to be representative of that time step in order to reduce the effects
244
of noise.53,62 The temperature of the solution at the inlet was (23 ± 1) °C.
245
Flowing solutions were used to promote, as much as possible, conditions in which the
246
dissolution rate is controlled by a surface reaction step and not by transport of ions away from
247
the surface. Flowing solutions also help maintain a high degree of undersaturation near the solid
248
surface, with respect to C3A and to any calcium aluminate hydrates, such as C3AH6, C2AH8, or
249
C4AH13, that might otherwise precipitate during the experiment. Dissolution rate measurements
250
were conducted at progressively increasing flow rates to determine the minimum flow rate above
251
which the average dissolution flux remained constant. Figure 2 indicates that (33 ± 0.5) mL
252
min−1 is a suitable minimum flow rate at the highest water activities (40 % water solutions). The
253
error bars overlap in Figure 2, but the low flow rates (0 mL min−1 and 15 mL min−1) still have a
254
lower median dissolution flux than the higher flow rates (33 mL min−1 to 67 mL min−1). This
255
suggests that a minimum of 33 mL min−1 is acceptable for preventing mass transport from being
256
the rate-controlling mechanism of C3A dissolution. Furthermore, as discussed further in Section
257
3 and the Supporting Information, no evidence of precipitated phases was detected when using a
258
flow rate of 33 mL min−1, although precipitated phases were detected at 0 mL min−1. Dissolution
259
fluxes at lower water activities are lower, implying that dissolution is even more likely to be
260
controlled by surface reaction at those lower activities. Therefore, and because noise and other
261
uncertainties in DHM measurements become increasingly significant at higher flow rates,53 33
262
mL min-1 was used for all experiments involving flowing solution.
10
ACS Paragon Plus Environment
Page 10 of 30
Page 11 of 30
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 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
Langmuir
263
264 265
Figure 2. Effect of flow rate on the median dissolution velocity in solutions with 60 % ethanol.
266
Error bars represent the interquartile range.
267 268
When C3A is dissolved in water, surface heights primarily should decrease as material is
269
removed from the solid (i.e., dissolution). However, the height of any point on the surface may
270
infrequently and temporarily appear to increase. Localized surface advances also may be
271
produced by inherent noise in the data, thermal effects, vibration effects, or errors in the
272
unwrapping algorithm.17 In previous studies, noise and unwrapping errors were minimized by
273
averaging the heights of localized groups of pixels,17,18 somewhat like a median filter in image
274
processing. However, for C3A dissolution, height increases could be caused by actual nucleation
275
and growth of a solid hydrate phase, especially in static solutions that may become progressively
276
saturated with respect to such phases. Therefore, spatial averaging techniques could not be used
277
in this study. Instead, the surface normal velocity was evaluated at each pixel to separate positive
278
and negative surface normal velocities. This pixel-by-pixel analysis was repeated several times
279
during a given experiment to evaluate if the measured flux changes with time. Evaluating the
280
velocities at each pixel yields a single histogram (Figure 3a), from which the dissolution and
281
apparent precipitation velocities must be deconvoluted. In addition, the data must be processed to
282
remove pixels with unchanging height as well as pixels that have a velocity too low to be
283
distinguished from noise (e.g., see Ref. 53). This deconvolution was performed using the Pearson
284
product-moment correlation coefficient, which is a statistical test to determine the strength of 11
ACS Paragon Plus Environment
Langmuir
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 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
Page 12 of 30
285
correlation between two variables. The Pearson coefficient can range from -1 (perfectly linear
286
negative correlation) to +1 (perfectly linear positive correlation). In this instance, the Pearson
287
coefficient was evaluated for height versus time at each pixel. Only velocities that had a Pearson
288
coefficient of at least 0.95 were retained, while all other values were disregarded (Figure 3b).
289
The analysis revealed portions of the surface at which this criterion suggested statistically
290
significant surface height increases, but ultimately no physical evidence of precipitation was
291
found, as discussed in Section 3 and in Supporting Information Section S.2. Therefore, the
292
surface height increases in Figure 3b are likely artefacts of the data processing and have been
293
subsequently disregarded. A more thorough discussion of the validity of retaining only the
294
identified (deconvoluted) “dissolution” rates is given in Supporting Information Section S.3.
295
Unless otherwise noted, all statistical descriptions and graphical representations of the data (i.e.,
296
median, mean, interquartile range, and standard deviation) in this study are made after the
297
deconvolution, such as in Figure 3b.
298
(a)
(b)
299
Figure 3. As-measured surface normal velocity (a) per pixel of the entire measurement area and
300
(b) deconvoluted as “dissolution” and “precipitation” velocities. These data are from a C3A #1
301
experiment with 60 % ethanol solution at a flow rate of 33 mL min−1.
302 303 304
2.3 Surface Analysis The surfaces of the C3A specimens were examined by SEM and Raman spectroscopy
305
after the dissolution experiments to search for evidence of solid hydrate precipitation. Field
306
emission scanning electron microscopy (SEM) with secondary and backscattered electron 12
ACS Paragon Plus Environment
Page 13 of 30
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 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
Langmuir
307
imaging was used to search for signs of precipitates on the surface. Raman spectroscopy was
308
performed with a Raman microscope using a 20× objective lens and a 532 nm laser with output
309
power between 10 mW and 20 mW. The Raman microscope had a spectral range of 4400 cm-1 to
310
50 cm-1 and a nominal resolution of about 12 cm-1. Raman spectroscopy was used because the
311
vibration bands of hydrated C3A solids can potentially be distinguished from those of anhydrous
312
C3A.63–66 A band around 3650 cm−1 is attributable to O-H vibration in hydrated C3A phases;63–66
313
this band is not present in unhydrated C3A.
314 315
3. RESULTS AND DISCUSSION
316
3.1 General Observations
317
In static and flowing solutions with water activity less than 0.83, the primary (initial)
318
topographical features that forms during dissolution are etch pits (Figure 4). The etch pits were 3
319
µm to 5 µm wide after 60 min in the solutions with water activities of 0.689 or 0.777. Etch pits
320
are commonly reported to form on a wide range of dissolving crystals, including alite.67 Besides
321
the locally enhanced dissolution fluxes at etch pits, Figure 5 shows significant variation in the
322
extent of material removal. At higher water activities for C3A #1, some regions on the order of
323
50 µm in characteristic dimension showed higher overall dissolution, as etch pits formed quickly
324
and were followed by an extensive surface roughening event (Figure 6). In static solutions, this
325
surface roughening began within 15 min to 20 min when aH2O = 0.824, but only took about 2 min
326
to begin when aH2O = 0.857. The widespread roughening behavior was somewhat delayed in
327
flowing solutions, beginning after about 30 min when aH2O = 0.824. Rapid roughening did not
328
happen at lower water activities, at least not during the observation times used in this study.
329
Larger-scale heterogeneous reactivity in water, as evidenced in Figure 5 and Figure 6, has also
330
been reported for tricalcium silicate and portland cement.10,12,67–72
331
At higher water activities for C3A #2, the primary topographical features are again etch
332
pits (Figure 7), although they are more faceted, are typically wider and deeper, and have steeper
333
walls than the etch pits formed on C3A #1 surfaces. The SEM micrograph in Figure 7 reveals an
334
etching morphology on C3A #2 that is reminiscent of scaffold microstructures produced by
335
leaching of a two-phase composite73,74 and suggests that the crystalline fraction in C3A #2 may
336
be preferentially leaching. The rapid surface roughening exhibited by C3A #1 was not observed
337
on C3A #2 during the time scales investigated. These differences in dissolution morphology 13
ACS Paragon Plus Environment
Langmuir
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 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
Page 14 of 30
338
imply that C3A #1 and C3A #2 have differing intrinsic variabilities (e.g., differences in defect
339
densities and types, composition and amorphous content, and initial synthesis process).
340
341
10 min
30 min
45 min
60 min
Figure 4. Formation of etch pits over time in C3A #1 in a static solution of 70 % ethanol.
342
14
ACS Paragon Plus Environment
Page 15 of 30
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 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
Langmuir
343 344
Figure 5. SEM image of C3A#1 after 40 min of exposure to 60 % ethanol solution at 33 mL
345
min−1. In the bottom left, etch pits are seen in a flatter, seemingly less-reacted region with etch
346
pits, while in the top right, the surface has undergone significant dissolution and “roughening.”
347
2 min
8 min
12 min
13 min
15
ACS Paragon Plus Environment
Langmuir
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 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
14 min
Page 16 of 30
18 min
348
Figure 6. Rapid dissolution behavior of C3A #1 in a static solution of 60 % ethanol. Etch pits
349
initially formed within the first 10 min, which was followed by a rapid dissolution of the entire
350
surface, starting at 12 min.
351
(a)
(b)
352
Figure 7. Dissolution morphology (etch pits) of C3A #2 after 40 min of exposure to 60 % ethanol
353
solution at 33 mL min−1: (a) surface topography from DHM and (b) SEM image (note the
354
chromium mask on the right side of the image). See Supporting Information for a time-lapse
355
video of this sample surface.
356 357 358
3.2 Dissolution in Static and Flowing Solutions The dissolution velocity increased with water activity, as expected. The 80 % ethanol and
359
70 % ethanol experiments yielded a somewhat “steady state” flux as etch pits formed and grew.
360
In contrast, dissolution in the 60 % ethanol solution gave rise to two phenomena: stable behavior
361
etch pit formation and dissolution up to 60 min, and rapid dissolution events within the first 20
362
min to 30 min (e.g., Figure 6). In static 60 % ethanol solution, the stable regime exhibited a 16
ACS Paragon Plus Environment
Page 17 of 30
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 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
Langmuir
363
median dissolution velocity of about 0.03 nm s−1 to 0.05 nm s−1, while the fastest velocity during
364
the rapid dissolution event was up to ten times greater (0.32 nm s−1 to 0.37 nm s−1). The 50 %
365
ethanol solutions, with the highest water activity, induced only the rapid dissolution events, with
366
the fastest velocity being about 1.5 nm s−1. Figure 8 compares the results from the various water
367
activities in static solution conditions. As expected, the figure shows that the dissolution velocity
368
increases with increasing water activity (decreasing ethanol proportion). The stable dissolution
369
for the 80 % ethanol, 70 % ethanol, and 60 % ethanol experiments all exhibit similar temporal
370
behaviors, having similar slopes in the log-log format of Figure 8. Similarly, the rapid
371
dissolution events for the 60 % ethanol and 50 % ethanol experiments exhibit similar behavior
372
(e.g., slope) prior to the sudden rapid dissolution event. However, Figure 8 also suggests that the
373
dissolution velocity is decreasing with time, since the ionic saturation of the solution is
374
increasing with time in the static solution and suppressing dissolution. It is therefore necessary to
375
replicate these experiments under flowing solution conditions to provide a nominally
376
undersaturated condition to agree with the assumptions in Eq. (8).
377
378 379
Figure 8. Median surface normal dissolution velocities for C3A #1 in a static solutions of varying
380
water activity. The rapid dissolution results (60 % ethanol and 50 % ethanol) are from one
381
experiment while the others represent the median values from the combined results of at least
382
four replicate tests.
383
17
ACS Paragon Plus Environment
Langmuir
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 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
384
The dissolution behavior under flowing conditions was similar to the behavior in static
385
conditions, albeit at different time scales. As was also seen in the static solutions, under flowing
386
conditions, the rapid surface roughening event occurred with the 60 % ethanol solution, but not
387
with the 70 % ethanol or 80 % ethanol solutions. This rapid dissolution event manifested around
388
25 min to 35 min in flowing solutions (Figure 9). Dissolution fluxes were comparatively constant
389
in 70 % ethanol or 80 % ethanol solutions (Figure 10) within 5 min to 15 min.
390
391 392
Figure 9. Median surface normal dissolution velocities for replicate experiments of C3A #1 in 33
393
mL min−1 60 % ethanol, indicating time intervals of transient, relatively constant, and rapid
394
dissolution behavior. Each of the four colored lines represents the data from one replicate
395
experiment.
396
18
ACS Paragon Plus Environment
Page 18 of 30
Page 19 of 30
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 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
Langmuir
(a)
(b)
397
Figure 10. Median surface normal dissolution velocities for replicate experiments of C3A #1 in
398
33 mL min−1 solutions with (a) 70 % ethanol and (b) 80 % ethanol. Each of the four colored lines
399
represents the data from one replicate experiment.
400 401 402
3.3 Statistical Analysis of C3A Dissolution Kinetics The data from each experiment were collated from within the relatively constant rate
403
region and yielded the flux data shown in Figure 11. The flux increases with increasing water
404
activity as expected. In addition, Figure 11 indicates that dissolution flux is a spectrum of values,
405
as has been shown for C3S10 and other minerals,17,18,41,43–48 primarily as a result of the intrinsic
406
heterogeneity of crystalline surfaces. The distributions are non-normal and heavy-tailed
407
(skewed), which is similar behavior to the distributions in the literature.
408
Based on Eqs. (8) and (9), the estimated distributions of k+ for water are shown in Figure
409
12, using the method detailed in the Supporting Information Section S.1. Table 2 summarizes the
410
statistics of these distributions, including the mean, median, standard deviation, and interquartile
411
range (IQR). Assuming n is a constant value, from the method detailed in the Supporting
412
Information, n = 5.24 for C3A #1 and n = 4.62 for C3A #2 (see the Supporting Information for a
413
discussion of the uncertainty in n). In Figure 12 the modal value of k+ is −1.94 µmol m−2 s−1 and
414
the median value of k+ is −2.06 µmol m−2 s−1 for C3A #1, both of which agree with the value of
415
−2.07 µmol m−2 s−1 found by linear regression (see Figure S1). This also agrees with the
416
dissolution flux value of dissolution flux is −2.0 µmol m−2 s−1 that was proposed based on
19
ACS Paragon Plus Environment
Langmuir
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 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
Page 20 of 30
417
simulations.75 For C3A #2, the modal k+ is −1.28 µmol m−2 s−1 with a median k+ of −1.42 µmol
418
m−2 s−1.
419
(a)
(b)
420
Figure 11. Histogram of surface normal dissolution flux as a function of ethanol content for (a)
421
C3A #1 and (b) C3A #2.
422
423 424
Figure 12. Empirical distributions of k+ for C3A #1 and C3A #2, assuming n is constant. These
425
data are available in the Supporting Information.
426
20
ACS Paragon Plus Environment
Page 21 of 30
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 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
Langmuir
427
Table 2. Surface Normal Dissolution Flux Statistics Solution Ethanol Water Content Activity 60 % 0.824 70 % 0.777 80 % 0.689 Predicted k+ (assuming constant n)
C3A #1 Flux (µmol m-2 s-1) Standard Mean Median Deviation −1.12 −0.754 1.57 −0.608 −0.473 0.545 −0.363 −0.286 0.295
0.741 0.425 0.254
C3A #2 Flux (µmol m-2 s-1) Standard Mean Median IQR Deviation −0.671 −0.502 0.757 0.465 −0.454 −0.369 0.361 0.287 −0.352 −0.305 0.214 0.216
−3.25
3.15
−1.99
−2.06
4.42
IQR
−1.42
2.45
1.87
428 429
The dissolution behaviors of C3A #1 and C3A #2 are different, as is evident by Figure 12,
430
with C3A #2 dissolving more slowly than C3A #1. This may be a function of the crystallinity and
431
purity of the C3A samples, since C3A #1 was 94 % C3A-c while C3A #2 consisted of only 84 %
432
C3A-c, which may indicate preferential etching of crystalline phases (e.g., Ref. 73). C3A #2
433
might have fewer intrinsic crystalline heterogeneities, owing to its lower crystalline content,
434
which can affect the overall reactivity (e.g., see Ref. 67 and references therein).
435
The regression analysis yielded water activity exponent parameters, n, of 5.2 and 4.6 for
436
C3A #1 and C3A #2, respectively. These values differ from the stoichiometric value of 6 from
437
Eq. (6), which is not surprising since Eq. (6) only applies to elementary reactions and C3A
438
dissolution is almost certainly not elementary. The spatial heterogeneity in dissolution
439
morphology and fluxes suggests that different (i.e., multiple) reaction mechanisms occur,41,47,76,77
440
which are associated with various sources (e.g., defect type and density, crystallographic
441
orientation and exposed surfaces). Furthermore, the value of n computed by this study considers
442
the bulk effect of all morphologies. Further study is needed to link specific surface reaction
443
mechanisms to a given probabilistic dissolution flux (or range of fluxes) and reaction order. That
444
effort is being undertaken for other minerals with the help of kinetic Monte Carlo
445
simulations.35,47,78
446
However, what if n is not assumed to be constant? Kinetically, n can be non-constant for
447
multistep reaction processes,79 and, as already discussed, the dissolution of C3A is a complicated
448
process depending on composition, crystallinity, and structural defects. Performing the least
449
squares optimization (see Supporting Information) for the C3A #1 data under this assumption
450
yields a spectrum of n values with a modal value of 5.3, which is very similar to the value of 5.2
451
obtained if n is assumed constant. The assumption of a non-constant n would imply that n is a
452
different constant value for each specific value of k+, which means that further discretization
453
could yield many values of n. If there are multiple reactions, each with a different reaction order, 21
ACS Paragon Plus Environment
Langmuir
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 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
454
it would be more reasonable to assume that there are a few constant n values, for example, one
455
value of n associated with each defect type. Further research is needed to confirm this, although
456
kinetic Monte Carlo simulations35,47,78 may offer additional insight.
457
Finally, the preceding analysis assumed that the data were obtained from experiments
458
without the formation of precipitated phases. To confirm this, the surface was examined a
459
posteriori by various techniques to search for evidence of any precipitated phases. XRD was
460
unable to detect any precipitated phase. SEM and Raman spectroscopy demonstrated that
461
precipitated phases could form under static solution conditions (see Figure S4 and Figure S6) but
462
showed no evidence of precipitation under flowing solution conditions (see Figure 5, Figure 7b,
463
and Figure S6). Therefore, the analysis is justified in assuming that the data are unadulterated by
464
precipitation in flowing solution experiments.
465 466 467
4. CONCLUSIONS Reflection digital holographic microscopy (DHM) was used to monitor the in situ, real
468
time nanoscale surface evolution of cubic tricalcium aluminate (C3A) in water in nominally
469
undersaturated conditions. The computed dissolution flux was observed to be a function of the
470
rate constant and the water activity raised to an empirically-derived exponent. Experiments were
471
conducted at multiple water activities to estimate these values.
472
The data and analysis strongly suggest that dissolution is spatially heterogeneous, owing
473
to the intrinsic variability of crystalline surfaces (e.g., defect density and type, crystal orientation,
474
etc.), thereby yielding a probabilistic distribution (histogram) of fluxes. Using least-squares
475
optimization, the shape of the empirical cumulative distribution function was derived to
476
determine the empirical probabilistic distribution of the rate constant. The rate constant
477
distribution was found to have median and interquartile range values of −2.1 µmol m−2 s−1 and
478
3.2 µmol m−2 s−1, respectfully, for highly crystalline C3A (94 % cubic C3A, 4 % amorphous
479
content) with a constant exponent of 5.2 and median and interquartile range values of −1.4 µmol
480
m−2 s−1 and 1.9 µmol m−2 s−1, respectfully, for less crystalline C3A (84 % cubic C3A, 10 %
481
amorphous content) with a constant exponent of 4.6. These data can now be used in
482
computational materials science models (e.g., HydratiCA75) to provide a more accurate
483
prediction of cement hydration.
484 22
ACS Paragon Plus Environment
Page 22 of 30
Page 23 of 30
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 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
Langmuir
485
SUPPORTING INFORMATION
486
Accompanying supporting information is available free of charge via the Internet at
487
http://pubs.acs.org. This material includes an appendix detailing the analysis methodology and
488
surface examination for precipitated phases, a time-lapse video of Figure 7, and the data from
489
Figure 12. The data from Figure 12 can also be acquired at https://doi.org/10.18434/M3SW2G.
490 491
ACKNOWLEDGMENTS
492
The authors would like to thank Paul Stutzman (National Institute of Standards and Technology,
493
NIST) for his expertise and advice on XRD and SEM analyses, Pan Feng (Southeast University)
494
for discussions on experimental methodology, Adam Pintar (NIST) for suggesting the use of the
495
nonparametric least squares optimization utilized in this study, and LaKesha Perry (NIST) for
496
performing the ICP-OES measurements. ASB would like to acknowledge the National Research
497
Council (NRC) for funding through the NRC Postdoctoral Research Associateship program at
498
NIST.
499 500
REFERENCES
501
(1)
Bullard, J. W.; Jennings, H. M.; Livingston, R. A.; Nonat, A.; Scherer, G. W.; Schweitzer,
502
J. S.; Scrivener, K. L.; Thomas, J. J. Mechanisms of Cement Hydration. Cem. Concr. Res.
503
2011, 41 (12), 1208–1223.
504
(2)
Biernacki, J. J.; Bullard, J. W.; Constantiner, D.; Meininger, R. C.; Juenger, M. C. G.;
505
Cheung, J. H.; Hansen, W.; Hooton, R. D.; Lüttge, A.; Thomas, J. J. Paving the Way for a
506
More Sustainable Concrete Infrastructure: A Vision for Developing a Comprehensive
507
Description of Cement Hydration Kinetics, NIST Special Publication 1138; 2013.
508
(3)
509
Luttge, A.; Arvidson, R. S. Reactions at Surfaces: A New Approach Integrating Interferometry and Kinetic Simulations. J. Am. Ceram. Soc. 2010, 93 (11), 3519–3530.
510
(4)
Following conventional cement notation (e.g., C = CaO, S = SiO2, A = Al2O3, H = H2O).
511
(5)
Damidot, D.; Bellmann, F.; Sovoidnich, T.; Möser, B. Measurement and Simulation of the
512
Dissolution Rate at Room Temperature in Conditions close to a Cement Paste: From
513
Gypsum to Tricalcium Silicate. J. Sustain. Cem. Mater. 2012, 1 (3), 94–110.
514 515
(6)
Nicoleau, L.; Nonat, A.; Perrey, D. The Di- and Tricalcium Silicate Dissolutions. Cem. Concr. Res. 2013, 47, 14–30. 23
ACS Paragon Plus Environment
Langmuir
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 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
516
(7)
517 518
Tricalcium Silicate. Cem. Concr. Res. 2014, 59, 118–138. (8)
519 520
Nicoleau, L.; Schreiner, E.; Nonat, A. Ion-Specific Effects Influencing the Dissolution of
Bellmann, F.; Sowoidnich, T.; Ludwig, H.-M.; Damidot, D. Dissolution Rates during the Early Hydration of Tricalcium Silicate. Cem. Concr. Res. 2015, 72, 108–116.
(9)
Kumar, A.; Reed, J.; Sant, G. Vertical Scanning Interferometry: A New Method to
521
Measure the Dissolution Dynamics of Cementitious Minerals. J. Am. Ceram. Soc. 2013,
522
96 (9), 2766–2778.
523
(10)
524 525
Juilland, P.; Gallucci, E. Morpho-Topological Investigation of the Mechanisms and Kinetic Regimes of Alite Dissolution. Cem. Concr. Res. 2015, 76, 180–191.
(11)
Pignatelli, I.; Kumar, A.; Alizadeh, R.; Pape, Y. Le; Bauchy, M.; Sant, G. A Dissolution-
526
Precipitation Mechanism Is at the Origin of Concrete Creep in Moist Environments. J.
527
Chem. Phys. 2016, 145, 54701.
528
(12)
529 530
during Its Early Hydration. Mater. Struct. 2005, 38 (278), 435–442. (13)
531 532
(14)
(15)
(16)
Gay, H.; Meynet, T.; Colombani, J. Local Study of the Corrosion Kinetics of Hardened Portland Cement under Acid Attack. Cem. Concr. Res. 2016, 90, 36–42.
(17)
539 540
Bisschop, J.; Kurlov, A. A Flow-through Method for Measuring the Dissolution Rate of Alite and Portland Cement Clinker. Cem. Concr. Res. 2013, 51, 47–56.
537 538
Mitchell, L. D.; Prica, M.; Birchall, J. D. Aspects of Portland Cement Hydration Studied Using Atomic Force Microscopy. J. Mater. Sci. 1996, 31 (16), 4207–4212.
535 536
Ferrari, L.; Kaufmann, J.; Winnefeld, F.; Plank, J. Reaction of Clinker Surfaces Investigated with Atomic Force Microscopy. Constr. Build. Mater. 2012, 35, 92–96.
533 534
Garrault, S.; Finot, E.; Lesniewska, E.; Nonat, A. Study of C-S-H Growth on C3S Surface
Brand, A. S.; Feng, P.; Bullard, J. W. Calcite Dissolution Rate Spectra Measured by in Situ Digital Holographic Microscopy. Geochim. Cosmochim. Acta 2017, 213, 317–329.
(18)
541
Feng, P.; Brand, A. S.; Chen, L.; Bullard, J. W. In Situ Nanoscale Observations of Gypsum Dissolution by Digital Holographic Microscopy. Chem. Geol. 2017, 460, 25–36.
542
(19)
Taylor, H. F. W. Cement Chemistry, 2nd ed.; Thomas Telford: London, 1997.
543
(20)
Sahachaiyunta, P.; Ponpaisanseree, K.; Bullard, J. W.; Stutzman, P. E.; Garboczi, E. J.;
544 545 546
Vichit-Vadakan, W. Virtual Testing in a Cement Plant. Concr. Int. 2012, 34 (9), 33–39. (21)
Meredith, P.; Donald, A. M.; Meller, N.; Hall, C. Tricalcium Aluminate Hydration: Microstructural Observations by in-Situ Electron Microscopy. J. Mater. Sci. 2004, 39 (3), 24
ACS Paragon Plus Environment
Page 24 of 30
Page 25 of 30
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 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
Langmuir
547 548
997–1005. (22)
549 550
Silva, D. A.; Monteiro, P. J. M. Early Formation of Ettringite in Tricalcium AluminateCalcium Hydroxide-Gypsum Dispersions. J. Am. Ceram. Soc. 2007, 90 (2), 614–617.
(23)
Rheinheimer, V.; Chae, S. R.; Rodríguez, E. D.; Geng, G.; Kirchheim, A. P.; Monteiro, P.
551
J. M. A Scanning Transmission X-Ray Microscopy Study of Cubic and Orthorhombic
552
C3A and Their Hydration Products in the Presence of Gypsum. Materials. 2016, 9 (9),
553
745.
554
(24)
Jupe, A. C.; Turrillas, X.; Barnes, P.; Colston, S. L.; Hall, C.; Häusermann, D.; Hanfland,
555
M. Fast in Situ X-Ray-Diffraction Studies of Chemical Reactions: A Synchrotron View of
556
the Hydration of Tricalcium Aluminate. Phys. Rev. B 1996, 53 (22), R14697–R14700.
557
(25)
Barnes, P.; Turrillas, X.; Jupe, A. C.; Colston, S. L.; O’Connor, D.; Cernik, R. J.; Livesey,
558
P.; Hall, C.; Bates, D.; Dennis, R. Applied Crystallography Solutions to Problems in
559
Industrial Solid-State Chemistry: Case Examples with Ceramics, Cements and Zeolites. J.
560
Chem. Soc. Faraday Trans. 1996, 92 (12), 2187–2196.
561
(26)
Merlini, M.; Artioli, G.; Cerulli, T.; Cella, F.; Bravo, A. Tricalcium Aluminate Hydration
562
in Additivated Systems. A Crystallographic Study by SR-XRPD. Cem. Concr. Res. 2008,
563
38 (4), 477–486.
564
(27)
Myers, R. J.; Geng, G.; Li, J.; Rodríguez, E. D.; Ha, J.; Kidkhunthod, P.; Sposito, G.;
565
Lammers, L. N.; Kirchheim, A. P.; Monteiro, P. J. M. Role of Adsorption Phenomena in
566
Cubic Tricalcium Aluminate Dissolution. Langmuir 2017, 33 (1), 45–55.
567
(28)
Holly, R.; Peemoeller, H.; Zhang, M.; Reardon, E.; Hansson, C. M. Magnetic Resonance
568
in Situ Study of Tricalcium Aluminate Hydration in the Presence of Gypsum. J. Am.
569
Ceram. Soc. 2006, 89 (3), 1022–1027.
570
(29)
Minard, H.; Garrault, S.; Regnaud, L.; Nonat, A. Mechanisms and Parameters Controlling
571
the Tricalcium Aluminate Reactivity in the Presence of Gypsum. Cem. Concr. Res. 2007,
572
37 (10), 1418–1426.
573
(30)
Myers, R. J.; Geng, G.; Rodriguez, E. D.; da Rosa, P.; Kirchheim, A. P.; Monteiro, P. J.
574
M. Solution Chemistry of Cubic and Orthorhombic Tricalcium Aluminate Hydration.
575
Cem. Concr. Res. 2017, 100, 176–185.
576 577
(31)
Mondal, P.; Jeffery, J. W. The Crystal Structure of Tricalcium Aluminate, Ca3Al2O6. Acta Crystallogr. 1975, B31, 689–697. 25
ACS Paragon Plus Environment
Langmuir
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 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
578
(32)
Mishra, R. K.; Fernández-Carrasco, L.; Flatt, R. J.; Heinz, H. A Force Field for Tricalcium
579
Aluminate to Characterize Surface Properties, Initial Hydration, and Organically Modified
580
Interfaces in Atomic Resolution. Dalt. Trans. 2014, 43 (27), 10602–10616.
581
(33)
582 583
Electrochem. Soc. 1966, 113 (10), 1067–1070. (34)
584 585
Danilov, V. V. On Hydration Mechanism in Cement Pastes. In Proceedings of the Sixth International Congress on the Chemistry of Cement; Moscow, 1974; Vol. 2 (Part 1), p 73.
(35)
586 587
Vermilyea, D. A. The Dissolution of Ionic Compounds in Aqueous Media. J.
Kurganskaya, I.; Luttge, A. Kinetic Monte Carlo Approach to Study Carbonate Dissolution. J. Phys. Chem. C 2016, 120 (12), 6482–6492.
(36)
Ikeda, T.; Boero, M.; Terakura, K. Hydration Properties of Magnesium and Calcium Ions
588
from Constrained First Principles Molecular Dynamics. J. Chem. Phys. 2007, 127 (7),
589
74503.
590
(37)
Tadros, M. E.; Jackson, W. Y.; Skalny, J. Study of the Dissolution and Electrokinetic
591
Behavior of Tricalcium Aluminate. In Colloid and Interface Science, Vol. IV; Kerker, M.,
592
Ed.; Academic Press: New York, 1976; pp 211–223.
593
(38)
Stumm, W. Chemistry of the Solid-Water Interface; Wiley, Ed.; New York, 1992.
594
(39)
Lasaga, A. C. Rate Laws of Chemical Reactions. In Kinetics of Geochemical Processes;
595
Lasaga, A. C., Kirkpatrick, R. J., Eds.; Mineralogical Society of America, Reviews in
596
Mineralogy 8, 1981; pp 1–68.
597
(40)
Lasaga, A. C. Transition State Theory. In Kinetics of Geochemical Processes; Lasaga, A.
598
C., Kirkpatrick, R. J., Eds.; Mineralogical Society of America, Reviews in Mineralogy 8,
599
1981; pp 135–169.
600
(41)
601 602
Reactivity: What Do We Know? Appl. Geochemistry 2014, 43, 132–157. (42)
603 604
(43)
Fischer, C.; Arvidson, R. S.; Lüttge, A. How Predictable Are Dissolution Rates of Crystalline Material? Geochim. Cosmochim. Acta 2012, 98, 177–185.
(44)
607 608
Arvidson, R. S.; Ertan, I. E.; Amonette, J. E.; Luttge, A. Variation in Calcite Dissolution Rates: A Fundamental Problem? Geochim. Cosmochim. Acta 2003, 67 (9), 1623–1634.
605 606
Fischer, C.; Kurganskaya, I.; Schäfer, T.; Lüttge, A. Variability of Crystal Surface
Lüttge, A.; Arvidson, R. S.; Fischer, C. A Stochastic Treatment of Crystal Dissolution Kinetics. Elements 2013, 9 (3), 183–188.
(45)
Emmanuel, S. Mechanisms Influencing Micron and Nanometer-Scale Reaction Rate 26
ACS Paragon Plus Environment
Page 26 of 30
Page 27 of 30
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 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
Langmuir
609 610
Patterns during Dolostone Dissolution. Chem. Geol. 2014, 363, 262–269. (46)
Fischer, C.; Finkeldei, S.; Brandt, F.; Bosbach, D.; Luttge, A. Direct Measurement of
611
Surface Dissolution Rates in Potential Nuclear Waste Forms: The Example of Pyrochlore.
612
ACS Appl. Mater. Interfaces 2015, 7 (32), 17857–17865.
613
(47)
614 615
Fischer, C.; Luttge, A. Beyond the Conventional Understanding of Water-Rock Reactivity. Earth Planet. Sci. Lett. 2017, 457, 100–105.
(48)
Saldi, G. D.; Voltolini, M.; Knauss, K. G. Effects of Surface Orientation, Fluid Chemistry
616
and Mechanical Polishing on the Variability of Dolomite Dissolution Rates. Geochim.
617
Cosmochim. Acta 2017, 206, 94–111.
618
(49)
Certain commercial equipment, instruments or materials are identified in this paper to
619
foster understanding. Such identification does not imply recommendation or endorsement
620
by the National Institute of Standards and Technology, nor does it imply that the materials
621
or equipment identified are necessarily the best available for the purpose.
622
(50)
Velez, K.; Maximilien, S.; Damidot, D.; Fantozzi, G.; Sorrentino, F. Determination by
623
Nanoindentation of Elastic Modulus and Hardness of Pure Constituents of Portland
624
Cement Clinker. Cem. Concr. Res. 2001, 31 (4), 555–561.
625
(51)
626 627
Calcium Silicate, Dissertation, École Polytechnique Fédérale de Lausanne, 2011. (52)
628 629
De la Torre, A. G.; Aranda, M. A. G. Accuracy in Rietveld Quantitative Phase Analysis of Portland Cements. J. Appl. Crystallogr. 2003, 36 (5), 1169–1176.
(53)
630 631
Quennoz, A. Hydration of C3A with Calcium Sulfate Alone and in the Presence of
Brand, A. S. Phase Uncertainty in Digital Holographic Microscopy Measurements in the Presence of Solution Flow Conditions. J. Res. Natl. Inst. Stand. Technol. 2017, 122.
(54)
Cuche, E.; Marquet, P.; Depeursinge, C. Simultaneous Amplitude-Contrast and
632
Quantitative Phase-Contrast Microscopy by Numerical Reconstruction of Fresnel off-Axis
633
Holograms. Appl. Opt. 1999, 38 (34), 6994–7001.
634
(55)
635
Schnars, U.; Jüptner, W. P. O. Digital Recording and Numerical Reconstruction of Holograms. Meas. Sci. Technol. 2002, 13 (9), R85–R101.
636
(56)
Kim, M. K. Digital Holographic Microscopy; Springer: New York, 2011.
637
(57)
Abbott, A. P.; Azam, M.; Ryder, K. S.; Saleem, S. In Situ Electrochemical Digital
638
Holographic Microscopy; A Study of Metal Electrodeposition in Deep Eutectic Solvents.
639
Anal. Chem. 2013, 85 (14), 6653–6660. 27
ACS Paragon Plus Environment
Langmuir
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 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
640
(58)
641 642
Lüttge, A.; Bolton, E. W.; Lasaga, A. C. An Interferometric Study of the Dissolution Kinetics of Anorthite: The Role of Reactive Surface Area. Am. J. Sci. 1999, 299, 652–678.
(59)
Bentz, D. P.; Coveney, P. V.; Garboczi, E. J.; Kleyn, M. F.; Stutzman, P. E. Cellular
643
Automaton Simulations of Cement Hydration and Microstructure Development. Model.
644
Simul. Mater. Sci. Eng. 1994, 2 (4), 783–808.
645
(60)
646 647
Liquid Mixtures. AIChE J. 1968, 14 (1), 135–144. (61)
648 649
Renon, H.; Prausnitz, J. M. Local Compositions in Thermodynamic Excess Functions for
Gmehling, J.; Onken, U.; Arlt, W. Vapor-Liquid Equilibrium Data Collection; Deutsche Gesellschaft für Chemisches Apparatewesen: Frankfurt, Vol. 1, Part 1a, 1991.
(62)
Kühn, J.; Charrière, F.; Colomb, T.; Cuche, E.; Montfort, F.; Emery, Y.; Marquet, P.;
650
Depeursinge, C. Axial Sub-Nanometer Accuracy in Digital Holographic Microscopy.
651
Meas. Sci. Technol. 2008, 19 (7), 74007.
652
(63)
Torréns-Martín, D.; Fernández-Carrasco, L.; Martínez-Ramírez, S. Hydration of Calcium
653
Aluminates and Calcium Sulfoaluminate Studied by Raman Spectroscopy. Cem. Concr.
654
Res. 2013, 47, 43–50.
655
(64)
Martínez-Ramírez, S.; Gutierrez-Contreras, R.; Husillos-Rodriguez, N.; Fernández-
656
Carrasco, L. In-Situ Reaction of the Very Early Hydration of C3A-Gypsum-Sucrose
657
System by Micro-Raman Spectroscopy. Cem. Concr. Compos. 2016, 73, 251–256.
658
(65)
Black, L.; Breen, C.; Yarwood, J.; Deng, C. S.; Phipps, J.; Maitland, G. Hydration of
659
Tricalcium Aluminate (C3A) in the Presence and Absence of Gypsum - Studied by Raman
660
Spectroscopy and X-Ray Diffraction. J. Mater. Chem. 2006, 16 (13), 1263–1272.
661
(66)
Torréns-Martín, D.; Fernández-Carrasco, L.; Martínez-Ramírez, S.; Ibáñez, J.; Artús, L.;
662
Matschei, T. Raman Spectroscopy of Anhydrous and Hydrated Calcium Aluminates and
663
Sulfoaluminates. J. Am. Ceram. Soc. 2013, 96 (11), 3589–3595.
664
(67)
665 666
Juilland, P.; Gallucci, E.; Flatt, R.; Scrivener, K. Dissolution Theory Applied to the Induction Period in Alite Hydration. Cem. Concr. Res. 2010, 40 (6), 831–844.
(68)
Hu, Q.; Aboustait, M.; Kim, T.; Ley, M. T.; Bullard, J. W.; Scherer, G.; Hanan, J. C.;
667
Rose, V.; Winarski, R.; Gelb, J. Direct Measurements of 3d Structure, Chemistry and
668
Mass Density during the Induction Period of C3S Hydration. Cem. Concr. Res. 2016, 89,
669
14–26.
670
(69)
Gallucci, E.; Mathur, P.; Scrivener, K. Microstructural Development of Early Age 28
ACS Paragon Plus Environment
Page 28 of 30
Page 29 of 30
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 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
Langmuir
671 672
Hydration Shells around Cement Grains. Cem. Concr. Res. 2010, 40, 4–13. (70)
673 674
Ménétrier, D.; Jawed, I.; Sun, T. S.; Skalny, J. ESCA and SEM Studies on Early C3S Hydration. Cem. Concr. Res. 1979, 9, 473–482.
(71)
Makar, J. M.; Chan, G. W. End of Induction Period in Ordinary Portland Cement as
675
Examined by High-Resolution Scanning Electron Microscopy. J. Am. Ceram. Soc. 2008,
676
91 (4), 1292–1299.
677
(72)
678 679
Topography. J. Am. Ceram. Soc. 2016, 99 (3), 773–786. (73)
680 681
Nicoleau, L.; Bertolim, M. A. Analytical Model for the Alite (C3S) Dissolution Toberer, E. S.; Seshadri, R. Template-Free Routes to Porous Inorganic Materials. Chem. Commun. 2006, No. 30, 3159–3165.
(74)
Souza, M. T.; Tansaz, S.; Zanotto, E. D.; Boccaccini, A. R. Bioactive Glass Fiber-
682
Reinforced PGS Matrix Composites for Cartilage Regeneration. Materials. 2017, 10 (1),
683
83.
684
(75)
Bullard, J. W.; Enjolras, E.; George, W. L.; Satterfield, S. G.; Terrill, J. E. A Parallel
685
Reaction-Transport Model Applied to Cement Hydration and Microstructure
686
Development. Model. Simul. Mater. Sci. Eng. 2010, 18 (2), 25007.
687
(76)
Lüttge, A.; Arvidson, R. S.; Fischer, C. Fundamental Controls of Dissolution Rate
688
Spectra: Comparisons of Model and Experimental Results. Procedia Earth Planet. Sci.
689
2013, 7, 537–540.
690
(77)
Pollet-Villard, M.; Daval, D.; Ackerer, P.; Saldi, G. D.; Wild, B.; Knauss, K. G.; Fritz, B.
691
Does Crystallographic Anisotropy Prevent the Conventional Treatment of Aqueous
692
Mineral Reactivity? A Case Study Based on K-Feldspar Dissolution Kinetics. Geochim.
693
Cosmochim. Acta 2016, 190, 294–308.
694
(78)
695 696 697
Kurganskaya, I.; Luttge, A. Kinetic Monte Carlo Simulations of Silicate Dissolution: Model Complexity and Parametrization. J. Phys. Chem. C 2013, 117 (47), 24894–24906.
(79)
Helfferich, F. G. Kinetics of Homogeneous Multistep Reactions; Compton, R. G., Hancock, G., Eds.; Elsevier: Amsterdam, 2001; Vol. 38.
698 699 700 701 29
ACS Paragon Plus Environment
Langmuir
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 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
702 703
For Table of Contents Only
704
705 706
30
ACS Paragon Plus Environment
Page 30 of 30