You are on page 1of 12

Construction and Building Materials 243 (2020) 118214

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Elucidating the nano-mechanical behavior of multi-component binders


for ultra-high performance concrete
Emily Ford, Aashay Arora, Barzin Mobasher, Christian G. Hoover, Narayanan Neithalath ⇑
School of Sustainable Engineering and the Built Environment, Arizona State University, Tempe AZ 85287, USA

h i g h l i g h t s

 Nanomechanical response of highly heterogeneous ultra-high performance cement pastes.


 Elucidates the influence of very low w/b and high volume of cement replacement materials on the nanoscale response.
 Explains the reasoning for an ultra-high stiffness phase in UHPC pastes.
 Models to upscale elastic modulus from nanoscale measurements and their applicability.

a r t i c l e i n f o a b s t r a c t

Article history: The nanomechanical signature of highly heterogeneous ultra-high performance (UHP) cement pastes are
Received 6 November 2019 explored in this paper. The UHP pastes are proportioned using 30% or 50% (by mass) of commonly avail-
Received in revised form 10 January 2020 able cement replacement materials including fly ash, microsilica, and fine limestone. Nanoindentation
Accepted 17 January 2020
experiments coupled with a Bayesian information criterion-based statistical approach is used to develop
Available online 31 January 2020
modulus-hardness (MH) clusters for the UHP pastes. While typical low-density (LD) and high-density
(HD) C-S-H phases are present in early-age UHP pastes, it is shown that an ultra-high stiffness (UHS)
Keywords:
phase, which is a composite of HD C-S-H and nanoscale CH, is predominant at later ages.
Nanoindentation
Ultra-high performance concrete
Nanoindentation data points to the presence of significantly higher proportions of mixed phases in the
C-S-H UHP pastes, comprising of cement hydrates/pozzolanic reaction products and unreacted phases including
Heterogeneity fine limestone and microsilica acting as micro-aggregates to enhance the stiffness of the paste. The pres-
Homogenization ence of such mixed phases complicates upscaling of the elastic modulus using multi-scale homogeniza-
tion models, which is to be carefully accounted for in such highly heterogeneous systems.
Ó 2020 Elsevier Ltd. All rights reserved.

1. Introduction plementary cementing materials such as microsilica and metakao-


lin, and expensive fine fillers like ground quartz) [2–5], thereby
Ultra-high performance concrete (UHPC) is becoming more increasing the cost of UHPC and impacting large-scale adoption.
prominent in infrastructural construction because of its wide- Several recent efforts have focused on tailoring the UHPC
ranging benefits: high compressive and flexural strengths result microstructure with commonly available cement replacement
in smaller cross-sections and material savings; improved ductility materials and fine fillers in order to make it more cost-effective,
enables efficient structural design; and enhanced durability pro- sustainable, and durable [6–10]. The use of coarse aggregates is
vides beneficial attributes to service-life and sustainability [1]. not common in UHPC, but recent studies have reported the devel-
UHPC binders employ a very low water-to-binder ratio (w/b) and opment of UHPC with coarse aggregates to counter issues relating
a high volume of reactive powders to ensure efficient space filling to volumetric stability, given the high paste content in UHPC
and porosity reduction [1]. The attributes that contribute to a den- [9,11,12]. The understanding that physical particle packing effects
ser and stiffer microstructure are responsible for its enhanced of binding materials and fillers/aggregates are as important (or
strength and durability. The demand for a denser microstructure even more important) as the ultimate degrees of reactivity has
has traditionally resulted in the use of very high amounts of resulted in scientific binder selection methods that comprehen-
cementing materials and fine fillers (including highly reactive sup- sively account for microstructural packing using a stochastic pack-
ing model with periodic boundary conditions [1]. This ensures a
⇑ Corresponding author. dense microstructure even at a low water content that does not
E-mail address: Narayanan.Neithalath@asu.edu (N. Neithalath). favor optimal reaction product formation. The microstructure of

https://doi.org/10.1016/j.conbuildmat.2020.118214
0950-0618/Ó 2020 Elsevier Ltd. All rights reserved.
2 E. Ford et al. / Construction and Building Materials 243 (2020) 118214

UHPCs thus formed are highly heterogeneous, with several reac-


tion products and unreacted particles. The UHP binders that are
discussed in this study, which have been shown to be significantly
cheaper compared to proprietary mixtures and other available
alternatives [9], utilize cement replacement levels of 30% or 50%
(by mass), with the replacement materials being fly ash, microsil-
ica, and fine limestone powder. This leads to a highly heteroge-
neous, multi-phase microstructure.
Statistical nanoindentation is a well-accepted technique to
probe microvolumes of cementitious materials to determine the
intrinsic mechanical properties of the different phases present
and their volume fractions [6,13,14], including UHPCs [15]. The
use of micromechanics-based models on nanoindentation data
coupled with colloid-based models for morphological arrangement
of C-S-H have also been used to identify the local packing densities
(and thus the porosities) of the different C-S-H phases [16,17].
Nanoindentation coupled with chemical mapping has also been
used to relate the chemical constitution of the phases to their Fig. 1. Particle size distribution (PSD) of the starting materials. The median size in
mechanical properties [18–20]. The use of nanoindentation to microns is shown in parentheses. The PSD of microsilica is not shown, but the
median size is < 0:5lm in a well-dispersed state.
determine the influence of cement replacement materials such as
fly ash, slag, and silica fume on the nanoscale response of pastes
has also been reported [21,13,22–25]. In this paper, statistical contents in the UHP pastes were based on previous work of the
nanoindentation is employed on two UHPC pastes with high vol- authors [1,9] that proportioned a family of UHPCs based on optimal
ume cement replacement by multiple, commonly available materi- particle packing, rheology, and substantial clinker factor reduction.
als. The presence of multiple/mixed phases attributable to different One quaternary binder with 17.5% fly ash, 7.5% silica fume, and 5%
starting materials with differing reactivities render the matrix limestone (all mass-based) replacing OPC for an overall cement
highly heterogeneous and complex to analyze. Thus, a comprehen- replacement level of 30% (termed as F17.5M7.5L5), and a ternary bin-
sive understanding of the micro-/nano-structure of such UHPCs der with 20% microsilica and 30% limestone (mass-based) replac-
require elucidation of the following aspects, which is carried out ing cement (termed as M20L30), for an overall cement
in detail in this paper: (i) the influence of a combination of very replacement level of 50%, were adopted. Paste samples were pre-
low w/b and high volume of cement replacement materials (in- pared and cured under moist conditions until their respective test-
cluding reactive fillers) on the nanomechanical properties of the ing durations.
constituent phases and their evolution, (ii) the presence of multiple
mixed phases including an ultra-high stiffness (UHS) phase that is 2.2. Sample preparation for nanomechanical studies
observed to be dominant in low w/b systems, and (iii) upscaling of
the elastic modulus from nanomechanical properties and the influ- The cylindrical paste samples were cut into ~12.5 mm thick
ence of material heterogeneity on predictive capability. discs using a Bruker IsoMet 1000 saw with a diamond wafer blade
after 30 ± 3 days or 90 ± 5 days of moist curing. Isopropyl alcohol
2. Experimental program (IPA) was used as a coolant for the saw to avoid further hydrating
the cement samples [27]. Samples were sanded and polished suc-
2.1. Materials and mixtures cessively using Buehler CarbiMet silicon carbide abrasive paper of
grit sizes 50, 18.3, 10.6, 9, 3, and 1 lm to ensure desired smooth-
Type I/II ordinary Portland cement (OPC) conforming to ASTM C ness for the nanoindentation testing [14,15]. Between sanding at
150, Class F fly ash conforming to ASTM C 618, limestone powder each grit size, the samples were ultrasonicated in a bath of IPA
conforming to ASTM C 568, and microsilica conforming to ASTM for 5 min to dislodge any trapped debris from the pores and
C 1240 were utilized in the preparation of ultra-high performance microstructure. Samples were polished until mirror reflectivity
cementitious pastes. Limestone powders with two different med- was achieved. Following the final ultrasonication, the samples
ian particle sizes (1.5 mm and 3.0 mm) were used to ensure efficient were stored in IPA until testing [27].
packing. The chemical composition of the source materials is
shown in Table 1, and their particle size distributions, extracted 2.3. Nanoindentation
from laser particle size analysis, in Fig. 1. A polycarboxylate ether
(PCE)-based superplasticizer with a solids content of 43% was used Nanoindentation grids were placed on the sample surface using
to ensure workability of the very low w/b pastes. an Ultra Nanoindentation Tester (UNHT3; Anton Paar) in areas that
Three different paste mixtures – two UHP pastes (w/b ~ 0.20) appeared to be smooth and free of visible surface defects when
and a companion OPC paste (w/c = 0.40) were prepared. The binder viewed under the microscope at 5, 20, and 50 magnifications.

Table 1
Chemical composition and specific gravity of the powder materials used in this study.

Components of the binder Chemical composition (% by mass) Specific gravity (g=cm3 Þ


SiO2 Al2O3 Fe2O3 CaO MgO SO3 LOI
OPC 19.60 4.09 3.39 63.21 3.37 3.17 2.54 3.15
Fly Ash (F) 58.40 23.80 4.19 7.32 1.11 3.04 2.13 2.24
Microsilica (M) >90.0 – – <1.0 – – – 2.18 [13]
Limestone (L), 1.5 mm >97% CaCO3 2.7 [26]
Limestone (L), 3 mm
E. Ford et al. / Construction and Building Materials 243 (2020) 118214 3

Each sample had at least 1250 indents, split among multiple grids ate contents. The tests were performed in an inert N2 environment
in different locations, to ensure that the heterogeneous nature of at a gas flow rate of 20 ml/s. The samples were heated from ambi-
the microstructure was adequately captured. The tests were per- ent temperature to 900 °C at a heating rate of 15 °C/min.
formed in the force control mode with a maximum displacement
cutoff of 250 nm (0.25 mm). This depth corresponded to an interac- 3. Results and discussions
tion volume idealized as a hemisphere with a radius of about 4–10
times the maximum cutoff, or 1.0–2.5 lm according to the esti- 3.1. Pore structure and CH contents of UHP pastes
mates in [18,28]. The distance between points in the grid was cho-
sen as 5 lm to ensure that the indents minimally influenced each The cumulative volume of mercury intruded and the differential
other [14]. The linear loading profile had a loading and unloading pore volumes as a function of pore diameter are plotted in Fig. 2(a)
rate of 12 mN/min with a pause for 5 s at the peak load when and (b) for the OPC and UHP pastes evaluated in this study. The
the maximum displacement cutoff was reached. pore volumes and critical pore sizes for the OPC and UHP pastes
The hardness (H) and the effective Young’s modulus (M), were after 30 days obtained from mercury intrusion porosimetry are
extracted from each of the indents according to Oliver and Pharr shown in Table 2. The 30-day porosities of the UHP pastes, calcu-
method [29,30]. The hardness is calculated as: lated from the total volume of mercury intruded and the cement
Pmax paste bulk density of about 2 g/cm3, was found to be in the 14–
H¼ ð1Þ 16% range, which is about half of that of the 0.40 w/c control
Ac
cement paste. The critical pore sizes are also lower for the UHP
Here, Pmax is the maximum load applied and Ac is the projected pastes, owing to the lower w/b and significantly improved particle
contact area, which is a function of the contact depth hc. The effec- packing through the use of ultrafine materials.
tive modulus is determined as: Thermogravimetric (TG) and differential thermogravimetric
pffiffiffiffi (DTG) traces of the OPC and UHP pastes after 30 days of hydration
p S
are shown in Fig. 3. The major peak in the 100–120 °C range corre-
M¼  pffiffiffiffiffiffiffiffiffiffiffiffi ð2Þ
2 Aðhc Þ sponds to the loss of evaporable water, the one in the 430–500 °C
S is the initial slope of the unloading curve [29,30]. The effective range corresponds to the dehydration of calcium hydroxide, and
modulus (M) is a function of the elastic moduli and Poisson’s ratios the one in the 750–850 °C range corresponds to the decarbonation
of the sample and the tip as shown below: of calcium carbonate [1,8,36]. The carbonate peaks accurately
identify the amount of added limestone in the pastes. Even with
1 1  t2 1  v 2i very high dosages of cement replacement materials (especially
¼ þ ð3Þ
M E Ei for the M20L30 paste), there was remnant CH in the pastes, once
again primarily attributed to the low w/b in these systems. The
t is the Poisson’s ratio of the sample, ti is the Poisson’s ratio of
volume fractions of CH were found to be 0.14, 0.058, and 0.047
the diamond tip, which is equal to 0.07, and Ei is the Young’s
for the OPC, F17.5M7.5L5, and M20L30 pastes respectively.
Modulus of the tip, equal to 1141 GPa [29 30]. t of different phases
was assigned after identifying them in statistical clustering [31]:
3.2. Frequency histogram of elastic properties from nanoindentation
clinker as 0.31, HD and LD C-S-H as 0.25, UHS phase as 0.29 (the
and deconvolution methods
average of C-S-H and CH), and mixed phases as 0.27 (between reac-
tants and products). As part of post-processing, abnormal load-
A representative frequency histogram for elastic moduli of dif-
depth curves representing surface pores or partial material col-
ferent phases in the 30-day cured F17.5M7.5L5 paste is shown in
lapse were removed from the data set as described in [13].
Fig. 4(a). If there exists n phases in the microstructure with each
phase occupying a volume fraction of fi (i = 1. . .n) such that
2.4. Pore structure and thermogravimetric analysis Pn
i¼1 f i ¼ 1, the properties of each phase can be approximated by
a Gaussian distribution with a probability density function (PDF)
The pore structure of the pastes was evaluated using Mercury
given as:
Intrusion Porosimetry (MIP). For MIP experiments, small pieces Xn
of the paste were weighed and placed in the low-pressure chamber PDF ¼ fw
i¼1 i i
ð5Þ
of the porosimeter (Quantachrome Instruments Pore Master). The
sample was filled with mercury starting from ambient pressure Here, wi is the property of interest of the phase. It has been
to 345 kPa (60 psi). The sample was then placed in the high- shown that the same statistical nanoindentation results can be fit
pressure chamber and the applied pressure increased to 414 MPa using different numbers of phases and volume fractions [37]. In
(60,000 psi). The pore diameter, d, as a function of the intrusion this study, a Bayesian Information Criterion (BIC) with negative
pressure was obtained from the Washburn equation [32] as: log likelihood method was implemented for statistical deconvolu-
tion. Minimizing the BIC determines the best number of phases to
4rcosðhÞ include [38]:
d¼ ð4Þ
DP
BIC ¼ 2  NlogL þ p  log ðnumÞ ð6Þ
where DP is the difference in the pressure between successive steps
(MPa), h is the contact angle between mercury and the cylindrical Here, num is the number of indentation points, p is the number
pore, taken as 130° in this study, and r is the surface tension of identifying parameters available at each indentation point (e.g.,
between mercury and the pore walls, taken as 485 mN/m M and H), and NlogL is the maximum negative log likelihood prop-
[33–35]. The critical pore size, indicative of the percolating pore size erty, which is defined as:
Y
of the material, was obtained from the peak of the differential pore NlogL ¼ maxðlogð PDFðni ÞÞÞ ð7Þ
n
volume curve. The porosity of the sample, u, was determined from
the cumulative volume of mercury intruded and bulk density of the Here, ni represents the distribution parameters, in the case of a
samples. Gaussian distribution the mean and standard deviation, that are
Thermogravimetric analysis was performed on the paste sam- iterated to maximize the likelihood function. The maximum nega-
ples to determine the calcium hydroxide (CH) and calcium carbon- tive log likelihood estimation was used to find the Gaussian normal
4 E. Ford et al. / Construction and Building Materials 243 (2020) 118214

Fig. 2. (a) Cumulative and (b) Differential pore volume curves after 30 days of hydration for the OPC, F17.5M7.5L5, and M20L30 pastes.

Table 2 and hardness determined. An iterative Gaussian fitting algorithm


Pore volume and critical pore sizes of UHP pastes after 30 days of hydration. [39,40] was also used to statistically deconvolute the data by
Mixture MIP Porosity (%) Critical pore size (lm) pre-defining the number of phases, however the BIC-negative log
OPC (w/c = 0.40) 28.22 0.026
likelihood method produced smaller errors and thus was chosen
F17.5M7.5L5 14.36 0.019 to analyze the indentation data in this study.
M20L30 15.21 0.014

3.3. Modulus-hardness clusters from nanoindentation and insights


into nanomechanical response

Based on the results of the previous section, MH clusters


obtained from BIC and negative log likelihood statistical analysis
for the companion OPC paste and the different UHP pastes are pre-
sented here. Table 3 lists the mean and standard deviation for the
hardness and stiffness of each cluster identified as well as their
corresponding volume fraction. Due to the presence of multiple
binders and fillers used in these UHP pastes and a very low w/b,
the matrices exhibit significant heterogeneity. In general, the reac-
tion products in the UHP pastes also have (M, H)  (65.0, 3.0) as
reported elsewhere on cementitious systems [25,41,42] even
though the distribution generally tends towards higher values. This
could be attributed to several reasons, including the increased den-
sity of reaction products and the presence of multiple stiffer phases
that remain unreacted under conditions of low w/b. These attri-
butes are is discussed in detail in the forthcoming sections.
Fig. 5 shows the MH clusters for the 30-day and 90-day
hydrated companion OPC paste (w/c = 0.40). As expected, there
are four phases, corresponding to low density (LD) C-S-H, high den-
sity (HD) C-S-H, CH, and unhydrated clinker grains. The volume
fractions of individual phases are also shown in the stacked bar
Fig. 3. TGA and DTG traces of the OPC and UHP pastes at 30 days.
next to the cluster plot. With increasing curing duration, the unhy-
drated cement fraction reduces in volume along with a slight
probability distribution functions that best represent the experi- increase in the HD C-S-H phase fraction.
mental data. The smaller the resulting BIC, the better the data
was modeled by the tested number of phases. For each input data 3.3.1. Fly ash-limestone-microsilica UHP paste
set, the number of phases that resulted in the smallest BIC varied Fig. 6(a) and (b) depict the MH cluster data obtained from the
between 3 and 5. For example, for the 30-day hydrated nanoindentation experiments for the 30- and 90-day cured quater-
F17.5M7.5L5 paste, the use of five phases resulted in the smallest nary blend UHP pastes containing fly ash, limestone, and microsil-
BIC. The elastic modulus histogram overlaid with the optimal five ica. The 30-day cured paste demonstrates the presence of five
Gaussian distributions generated from the BIC method for the distinct phases based on M and H, while there are only three
30-day hydrated F17.5M7.5L5 paste are shown in Fig. 4(b). Occasion- phases in the 90-day cured paste. Most of the experimental inden-
ally the clustering algorithm yielded clusters that were embedded tation moduli of the different solid phases fall in the range of 20–
inside of one another, or unreasonably split the high hardness- 130 GPa, as is commonly reported for cementitious materials
stiffness values into multiple clusters. In such cases, the clusters [15,41,42,25]. The hardness values of the constituent phases are
were manually grouped together, and the new average stiffness somewhat higher than those reported for conventional cementi-
E. Ford et al. / Construction and Building Materials 243 (2020) 118214 5

Fig. 4. (a) Frequency histogram of stiffness in the F17.5M7.5L5 paste after 30 days, and (b) fits of BIC with maximum log likelihood scheme with 5 phases on the experimental
indentation moduli.

Table 3
Elastic properties (M, H) and volume fractions (f) of the different phases in the UHP pastes. F, M, L, and CA denotes fly ash, microsilica, limestone, and carboaluminates
respectively.

UHP paste Phase Extracted values from nanoindentation results Reference values
M (GPa) H (GPa) f M (GPa) H (GPa)
30 d 90 d 30 d 90 d 30 d 90 d
OPC LD C-S-H 19.65 ± 8.18 9.10 ± 3.37** 0.40 ± 0.27 0.39 ± 0.18 0.03 0.03 18.2 [43] 0.45 [43]
22.89 [44] 0.93 [44]
HD C-S-H 36.85 ± 5.05 33.37 ± 6.54 1.80 ± 0.33 1.55 ± 0.38 0.68 0.72 29.1 [43] 1.36 [14]
33.65 [24] 1.04 [24]
CH 49.52 ± 9.32 56.15 ± 9.94 2.54 ± 0.76 2.69 ± 0.60 0.18 0.19 39.77–44.89 1.31 [43]
[45]
Clinker 90.48 ± 18.63 87.44 ± 28.50 5.62 ± 2.03 5.72 ± 2.45 0.11 0.06 100.3 [46] 7.86 [46]
122.2 [44] 6.67 [44]
F17.5M7.5L5 LD C-S-H 26.32 ± 4.78 – 1.39 ± 0.36 – 0.22 – 18.2 [43] 0.45 [43]
22.89 [44] 0.93 [44]
HD C-S-H 37.89 ± 5.20 – 1.66 ± 0.28 – 0.34 – 29.1 [43] 1.36 [14]
33.65 [24] 1.04 [24]
UHS 44.47 ± 9.01 49.79 ± 8.69 2.15 ± 0.59 2.60 ± 0.59 0.25 0.17 42.8 [16] 1.43 [16]
Mixed (F, M, L, 67.19 ± 13.06 62.73 ± 10.25 3.97 ± 1.06 3.84 ± 0.78 0.1 0.62 75.1 (F) [21] 8.47 (F)
CA) [21]
72.8 (M) [31] 6.0 (M)
[29]
83.8 (L) [47] 1.98 (L)
[49]
51–80 (CA)
[48]
Clinker/ 100.21 ± 32.80 101.40 ± 28.14 8.12 ± 3.20 9.38 ± 2.68 0.09 0.21 100.3 [46] 7.86 [46]
Unreacted 122.2 [44] 6.67 [44]
M20L30 UHS/C-S-H 44.11 ± 5.06 40.31 ± 6.99 1.85 ± 0.26 1.85 ± 0.48 0.59 0.77 42.8 [16] 1.43 [16]
Mixed (M, L) 64.10 ± 4.97 (M) 57.82 ± 11.08 4.48 ± 0.37 (M) 2.75 ± 0.79 0.14 0.16 72.8 (M) [31] 6.0 (M)
66.29 ± 8.36 (L) 3.73 ± 0.49 (L) [29]
0.15 83.8 (L) [47] 1.98 (L)
[49]
Clinker/ 87.42 ± 25.13 96.29 ± 25.46 5.83 ± 2.81 6.21 ± 2.62 0.12 0.07 100.3 [46] 7.86 [46]
Unreacted 122.2 [44] 6.67 [44]
**
Could be an experimental anomaly where the load–displacement curves for a few indents predominantly on pores were not identified or removed by the algorithm.

tious materials, the reasons for which can be found in the discus- either as C-S-H with a very high packing density (~0.84) [16,50]
sions below. or as an intimate nanocomposite of C-S-H and nanoscale CH [18]
For the 30-day cured paste, at the lower end of the modulus especially in low w/b binders such as the ones used in this work
spectrum, three different phases are observed, with mean values (see a later section for a more detailed analysis of the UHS phase).
of M at 26.3 GPa, 37.9 GPa, and 44.5 GPa respectively. Based on The fact that fine limestone filler (d50 of 1.5 lm or 3 lm) and
published literature [14,15,44], these can be assigned to low- microsilica (d50 < 0.5 lm in a well-dispersed state) fills the intersti-
density (LD) C-S-H, high-density (HD) C-S-H, and an ultra-high tial spaces between the cement and fly ash particles in these mix-
stiffness (UHS) phase respectively. In general, the material proper- tures likely enhances the formation of such a composite C-S-H with
ties of LD and HD C-S-H are reported to be independent of the mix- nanoscale CH phase due to the significant reduction in available
ture proportions [14,15,44], as was noticed in this study also. The capillary spaces that would otherwise enable CH precipitation as
UHS phase (M ~ 45–48 GPa, H ~ 2.0 GPa) is reported in literature microcrystals. Indeed, the lower the w/b, the lower the amount
6 E. Ford et al. / Construction and Building Materials 243 (2020) 118214

Fig. 5. MH cluster for companion OPC paste after: (a) 30 days, and (b) 90 days of curing.

Fig. 6. MH clusters of fly ash-limestone-microsilica UHP paste after: (a) 30 days, and (b) 90 days of reaction. F, M, L, and CA denotes fly ash, microsilica, limestone, and
carboaluminates. Since this is likely a mixed phase, with accurate identification difficult, the foregoing abbreviation is used.

of available water to react, thereby lowering the C/S ratio of C-S-H a composite response of multiple phases, has been put forth in
gel and increasing the elastic modulus and hardness of C-S-H [6,19,24,39] also, and is more significant for multi-component
[51,52]. Similar to the UHS phase, CH from cement hydration has blends displaying higher heterogeneity.
an indentation modulus in the 40–45 GPa range [31,45]. As The grouping at the high end of the M and H spectrum can be
observed from the DTG curves in Fig. 3, CH was present in this attributed to unhydrated clinker. The mean stiffness and hardness
UHP paste, however, the volume fraction of CH observed was of this phase are higher than those commonly reported for clinker
around 0.05. This is significantly lower than the volume fraction grains, and the spread also is much larger (compared to the spread
of UHS phase, and thus the assumption that the UHS phase is a for the OPC pastes shown in Fig. 5). This could be because there are
composite of C-S-H and CH is likely valid. Specific chemical infor- several high hardness/modulus phases in fly ash (such as hematite
mation at each indentation point can help accurate identification and mullite; XRD spectra revealed the presence of mullite
of the composition of UHS phases in multi-component binder sys- (M  220 GPa) in this fly ash). It has also been shown that accurate
tems, which is the focus of an ongoing work. identification of high modulus phases in a matrix requires a mod-
The mixed phase is a combination of multiple phases with a ulus mismatch ratio (the ratio between the stiffness of the matrix
mean M of 67.2 ± 13.1 GPa. In highly packed systems with multiple (M m ) and the modulus of the indented microstructure (M i )),
blends and a variety of fine particulates that react at varying M m =M i , between 0.2 and 5 for indentation depths smaller than or
degrees, it is not uncommon to encounter such stiff and hard equal to 10% of the characteristic length scale of the microstructure
phases [18]. The carboaluminates formed through the reaction [6,39,25]. In many conventional cementitious pastes with ‘‘softer”
between carbonates from limestone and alumina from fly ash matrices, this condition might not be satisfied, and high modulus
[53–56] have a reported moduli of 51–80 GPa [48]. Based on pre- ratio phases could be ignored [6]. A stiff UHP matrix allows the
vious work, about 20% of limestone reacts to form carboaluminates identification of these phases, resulting in values of mean indenta-
in the presence of alumina from fly ash [57]. Unreacted fly ash par- tion modulus and hardness of the unreacted phases that are higher
ticles, microsilica, and limestone used as a filler in UHP pastes have than sometimes reported.
moduli in the 75–80 GPa range [21,31,47]. It is also possible that After 90 days of hydration, the LD and HD C-S-H phases are not
some phase interfaces are also captured in the mixed phase. Thus, detected in this paste. Instead, a UHS phase is mainly observed
the mixed phase could consist of unreacted fly ash, limestone, with a mean modulus of 49.8 GPa. Enhanced hydration in very
microsilica, carboaluminates and/or the interfaces between them. low w/b systems is reported to favor formation of an UHS phase
Such a possibility, that the phases detected in the MH plots are [16], which is in line with the observation here. The pozzolanic
E. Ford et al. / Construction and Building Materials 243 (2020) 118214 7

reaction of fly ash/microsilica that results in C-S-H gel with a lower Limestone and microsilica have similar moduli but different hard-
C/S ratio could also contribute to this observation. Lowering the C/S ness as shown in Table 3. These clusters could have been also com-
ratio of C-S-H gel, which is a result of increased silica polymeriza- bined without any appreciable loss of accuracy. The size of
tion, is noted to increase the elastic modulus and hardness limestone powder (d50 of 1.5–3 mm) and microsilica (d50 < 1 mm)
[51,52,58]. lends credence to the fact that the indentation measurements
Between the 30- and 90-day cured specimens, the number of attributed to them are likely to be those of mixed phases contain-
clusters dropped from five to three, with a large volume fraction ing these materials and the hydrated products. After 90 days of
being occupied by the mixed phases. As reaction progresses in such hydration, the UHS phase has a volume fraction roughly equal to
multi-component blends through pozzolanic reaction of fly ash the sum of the volume fractions of the UHS phase and the Mixed
and silica fume and the preferential deposition of products closer (M) phase in the 30-day sample. This is an indication of the reac-
to reacting surfaces in densely packed systems, this can be tion of microsilica to further form low C/S C-S-H gel. Lowering of
expected. A closer look at the MH cluster of the 30- and 90-day C/S is reported to result in a pronounced increase in modulus
cured mixtures also suggests that for M < 65 GPa, about 10–25% and hardness [51,52], which can also be noticed in Fig. 7(b). The
of the reaction products show H > 3 GPa, indicating that a multi- unhydrated cement phase reduces in volume fraction from 30 to
phase response is indeed acquired. The increased volume fraction 90 days of hydration as expected. The Mixed (L, M) phase with
of this phase at later ages can also be attributed to this reason. indentation modulus of ~60 GPa, attributed to limestone powder
The cluster with (M, H) of (101.4, 9.38) can be attributed to unre- and some hydrates, is also present in this system at a similar vol-
acted clinker and other stiffer/harder phases. The hardness of dif- ume fraction as Mixed (L) in the 30-day paste. This points to fur-
ferent phases in the fly ash-based UHP matrix discussed above ther consumption of microsilica with time.
are higher than those reported for conventional cementitious
pastes, which is contributed by the lower w/b in the UHP paste
3.4. Multi-step homogenization for macroscale elastic property
mixtures and corroborated by a stronger matrix [6,25,39], since
prediction from nanoscale response
hardness correlates directly with compressive strength.
The use of nanomechanical properties of individual phases in
3.3.2. Microsilica-limestone UHP paste homogenization models to predict the macroscale response of
The reaction products in microsilica-limestone UHP paste (with composites has been carried out for normal and high-
a 50% cement replacement) are fundamentally different from that performance concretes [14,60]. The analytical homogenization
of the fly ash-based system discussed earlier. The MH clusters of models are based on Eshelby’s solution of a strain localization
this paste after 30 and 90 days of hydration are shown in Fig. 7(a) tensor for ellipsoidal inclusions embedded in a matrix [61,62].
and (b) respectively. A UHS phase is noticed at both ages, with This method has been applied regularly in homogenization
M ~ 42 ± 6 GPa, and H ~ 2.0 GPa, and the LD/HD C-S-H phases schemes of cementitious materials [61–65]. The shear and bulk
are absent as independent phases. It is known that the incorpora- moduli of the individual phases are obtained from their elastic
tion of microsilica or nanosilica increases the volume fraction of moduli and Poisson’s ratios [61,64]. In this study the bulk and
high-modulus C-S-H phases due to enhanced pozzolanic reaction shear moduli were determined based on the nanoindentation
and consequent reduction of the C/S ratio [59]. The high microsilica cluster data, converting the indentation modulus to Young’s mod-
content and the high overall cement replacement level could have ulus as shown in Eq. (3) with the following Poisson’s ratios
contributed to denser reaction products even at earlier ages, result- assigned to each cluster based on the major component [31]: clin-
ing in high M and H values, and the high volume fraction of this ker as 0.31, HD and LD C-S-H as 0.25, and UHS phase as 0.29 (the
phase. average of C-S-H and CH). For the mixed phase where every
In the 30-day paste an interesting observation is that there are indent is a composite response across multiple possible phases,
two distinct clusters with similar M values (~65 GPa), but different the Poisson’s ratio was taken as 0.27, which lies between the
H values. The lower hardness phase called Mixed (L) phase likely Poisson’s ratios of the reactants and hydration products. The fine
corresponds to the mixed phase of limestone powder (which and coarse aggregate were considered to have a Poisson’s ratio of
constitutes 30% of the starting materials) and some hydrates, while 0.25 [66]. The homogenized bulk and shear moduli for two-phase
the higher hardness phase called Mixed (M) can be attributed to materials can be quantified from the individual phase properties
the mixed phase of unreacted microsilica and some hydrates. as shown in Eqs. (8) and (9).

Fig. 7. MH clusters of microsilica-limestone UHP paste after: (a) 30 days, and (b) 90 days.
8 E. Ford et al. / Construction and Building Materials 243 (2020) 118214

P   1
1 þ Ar KKri  1
if iK i
¼ ð8Þ
P   1
K hom
Ki
i f i 1 þ Ar K r  1

P   1
1 þ Br GGri  1
i f i Gi
¼ ð9Þ
P   1
Ghom
Gi
i f i 1 þ Br Gr  1

where fi is the volume fraction of each inclusion phase,K r and Gr rep-


resent the bulk moduli and shear moduli of the reference medium
(the matrix), K i and Gi represent the bulk moduli and shear moduli
of the inclusion phases. The coefficients Ar and Br are given as:

3K r
Ar ¼ ð10Þ
3K r þ 4Gr
Fig. 8. Components of the 30 day old pastes used for homogenization.

6K r þ 12Gr
Br ¼ ð11Þ
15K r þ 20Gr other identified cementitious phases were used as double inclu-
sions in the previously homogenized matrix. In the third step, both
The double-inclusion method [67,68] based on Eshelby’s solu-
coarse and fine aggregates were used as inclusions. The homoge-
tion consists of an ellipsoidal inclusion of stiffness SI1 embedded nization approaches are schematically represented in Fig. 9.
in another ellipsoidal matrix of stiffness SI2 , which is further Table 4 reports the stiffness results from the analytical
embedded in an infinitely extended homogeneous medium of stiff- homogenization process for both the paste and UHPC. The elastic
ness S. The double-inclusion method is capable of accounting for modulus determined from quasi-static, radial strain controlled
the inclusion-inclusion interactions in addition to the inclusion- compression tests from a recent study of 28-day cured UHPC
matrix interactions considered in the Mori-Tanaka approach, and by the authors [9] are also provided. The homogenized elastic
has also be implemented for cementitious materials [63,69]. The modulus of the fly ash-based UHPC is similar to that of the
average elastic moduli can be calculated as: experimental value from compression tests, but the microsilica-
h i1 based mixture shows a higher predicted value. The higher pre-
S ¼ S1 þ f 1 a þ f 2 b ð12Þ dicted stiffness of the microsilica-based paste can be attributed
to the presence of large volume fractions of UHS phase as well
In the above equation, f1 and f2 are volume fractions of two as the mixed phases observed from the cluster plots. It is
inclusion phases, a and b are functions of f1, f2, and the Eshelby’s hypothesized that the high volume of microsilica used in this
tensors S, SI1 , and SI2 . Detailed derivation and analysis procedure paste was less well-dispersed in a low w/b system, despite
are described in [67,69]. For both Mori-Tanaka and double- extended mixing and the use of a high volume of superplasti-
inclusion methods, a multi-level nested homogenization approach cizer, as reported elsewhere [71]. The reduced dispersion of
is implemented as described below. microsilica reduces its degree of reaction, and consequently, high
Previous studies have used a combination of techniques includ- stiffness mixed phases are present instead of lower density C-S-
ing MIP, X-ray tomography, and electron microscopy to identify H phases, which results in an over-prediction of the paste mod-
the pore volumes that are of interest in predicting macroscale ulus. When aggregates are incorporated into UHPC, the shearing
response from microscale observations [60,63,70]. In this work, action of aggregates on the paste de-agglomerates many of these
considering the maximum indentation depth of 250 nm and a larger associations. UHPC mixtures are mixed in a high-shear
probed volume with a radius of ~1.5–2.5 lm, it is assumed that mixer for a longer duration [9], and hence more reaction prod-
the effect of pores smaller than 1 lm would be accounted for in ucts of lower stiffness are formed at the expense of higher stiff-
the deconvoluted stiffness of the phases. Hence, the pore volume ness agglomerated phases. Thus the experimental stiffness of the
corresponding to d > 1 lm from MIP is used in the homogenization concrete mixture is lower than the upscaled value obtained
process for the standalone void phase. The volume fractions of through analytical homogenization. Ongoing chemical species
phases obtained from statistical nanoindentation, which is used mapping of indentation locations is expected to shed more light
in the homogenization process, are shown in Fig. 8. into this effect.
Utilizing the Mori-Tanaka scheme, two different homogeniza-
tion approaches were used. In the first case, a two-step homoge- 3.5. The nature of the UHS phase in UHP pastes
nization method was used. Here, the cementitious phase with
the highest volume fraction was used as the matrix (HD CSH for Previous studies have identified the UHS phase as a very high
F17.5M7.5L5 and UHS Phase for M20L30), and the other cementitious packing density (~0.84) C-S-H phase [16,50] or as a nanocomposite
phases and voids as inclusions. In the multi-step Mori-Tanaka of HD C-S-H and nanoscale CH [18,72]. In very low w/b pastes
method, the order of homogenization was varied based on the where the particles are efficiently packed through the use of ultra-
modulus of the different phases. The phase with the highest vol- fine materials, there is deficiency of water and space for the forma-
ume fraction was chosen as the matrix, and the voids were added tion of CH microcrystals outside the gel pores. In such highly
as an inclusion. In the subsequent steps, the phase with the lowest confined reaction zones, nanosized crystals are preferred since
modulus was homogenized with the resultant of the previous step. there is limited space between the reactants, and the supersatura-
The remaining process was carried out similarly. For the double tion in the liquid is high [18].
inclusion model, the homogenization is not capable of directly Assuming that C-S-H is nanogranular in nature as described in
accounting for the voids. Hence, the first step in this process was [43,73–75], and that the UHS clusters identified through nanoin-
the homogenization of voids with the reaction product with the dentation contain no other products or interfaces with similar
highest volume fraction using the Mori-Tanaka method. Next, the mechanical characteristics, the phase packing density (g) can
E. Ford et al. / Construction and Building Materials 243 (2020) 118214 9

a.
Voids
Matrix Other CSH or Coarse agg. Homogenized
(product with mixed phases Fine agg.
highest f) Clinker from Step 1

n steps unl paste


is homogenized
b.
Fine agg.
Matrix Homogenized Inclusion 2 Homogenized
Inclusion 1
(product with from Step 1 Coarse from Step n
highest f) agg.

c.
Less sff Fine agg.
Matrix Voids Homogenized phase Homogenized
(product with from Step 1 Sffer Coarse from Step n
highest f) phase agg.

Fig. 9. Schematic representation of the homogenization schemes: (a) two-step Mori-Tanaka scheme, (b) multi-step Mori-Tanaka scheme, and (c) double inclusion method.

Table 4
Analytical homogenization results for the 30-day cured UHPC mixtures.

UHP paste Method of homogenization Elastic modulus from homogenization Elastic modulus of UHPC from experiments (GPa) [9]
E of UHP Paste (GPa) E of UHPC (GPa)
F17.5M7.5L5 Mori-Tanaka (Two step) 36.96 46.56 47.53
Mori-Tanaka (Multi-step) 37.17 44.38
Double inclusion 36.78 44.09
M20L30 Mori-Tanaka (Two step) 46.05 51.18 43.1
Mori-Tanaka (Multi-step) 45.68 50.90
Double inclusion 45.67 50.89

determined based on its relationship with the indentation modulus The HD C-S-H in the OPC- and fly ash-based mixtures demon-
and hardness described as follows [16,17]: strate packing densities of 0.76 ± 0.03 and 0.72 ± 0.03 respectively
(note that the microsilica-based paste does not show a HD C-S-H
M Y
¼ ðts ; g; g0 Þ ð13Þ phase) based on the above-mentioned granular model, which cor-
ms M
responds to FCC or HCP packing (g  0.74) [17,78]. The packing
densities reported are average values from a large number of
H Y
¼ ðas ; g; g0 ; hÞ ð14Þ indents. The volume fraction of C-S-H globules in HD C-S-H is
hs H
approximately equal to the packing density of HD C-S-H, ~0.75,
where M is the indentation modulus, g is the packing density, ms is therefore a gel porosity of /HDCSH 0.25 can be used for HD C-S-
the asymptotic elastic modulus =Mðg ! 1Þ, ts is the Poisson’s ratio H for all the pastes. If the CH nanocrystals are formed within the
of the solid, g0 is the solid percolation threshold, H is the indenta- HD C-S-H structure, they result in an effective reduction in the
tion hardness, hs is the asymptotic hardness = Hðg ! 1Þ of the HD C-S-H porosity (/HD-CSH) by an amount equal to the CH volume
cohesive-friction solid that obeys the Druker-Prager criterion, as is fraction (fCH). Thus, the residual gel porosity of the UHS phase
the solid friction coefficient, and h is the indenter cone angle. (/residual) can be defined as:

Es /residual ¼ /HDCSH  f CH ð17Þ


ms ¼ ð15Þ
1  t2s
Note that it is being implicitly assumed here that the local con-
hs ¼ cs  Að1 þ Bas þ ðC as Þ þ ðDas Þ Þ
3 10
ð16Þ straints force the formation of nanoscale CH and not microcrystals
as in conventional pastes. The theoretical maximum CH fraction
where Es is the plane-stress elastic modulus and cs is the cohesion that can precipitate as nanocrystals in HD C-S-H to form the UHS
of the solid. The variables A, B, C, and D are the fitted parameters phase is equal to /HD-CSH, even though it could generally be lower.
corresponding to a simulated indentation experiment on a granular, The volume fraction of CH, fCH, in each of the indentation volumes
isotropic porous solid with a conical indenter with equivalent cone cannot be accurately determined without chemical mapping of the
angle to a Berkovich tip [76 77]. For a Berkovich indenter with a indentation locations using energy dispersive x-ray analysis. A
cone angle of 70.32°, A = 4.76438, B = 2.5934, C = 2.1860, and Mori-Tanaka homogenization scheme is used to determine the
D = 1.6777 [76]. The solid properties ms, cs, ms, as, and the packing effective modulus of the UHS phase consisting of HD C-S-H and
density g can be back-calculated from the experimental M and H nanoscale CH. The homogenized, theoretical UHS phase moduli
values [16,17]. are plotted as a function of residual gel porosity in Fig. 10, shown
10 E. Ford et al. / Construction and Building Materials 243 (2020) 118214

was supported using thermogravimetric analysis and nanogranular


C-S-H packing density calculations. Both the UHP pastes contained
several mixed/composite phases as determined from the MH
response, which is not uncommon in packed systems with multiple
blends of fine, incompletely reacted particles. This renders the
indentation-based analysis of such heterogeneous systems more
cumbersome; however, the presence of phases stiffer than the
hydrates function as micro-inclusions that enhance the properties
of the UHPC. Nanomechanical analysis reveals the enhanced influ-
ence of physical packing of particles in low w/b systems such as
UHPC.
Analytical homogenization models based on Eshelby’s solution
for inclusions embedded in a matrix were used to upscale the elas-
tic response of the individual phases in UHP paste to the elastic
response of UHPC. For the fly ash-based UHPC, the homogenized
Fig. 10. Homogenized UHS phase modulus as a function of residual porosity of the E was in good agreement with that determined experimentally
UHS phase.
using strain controlled compression tests on UHPC specimens.
However, for the microsilica-based UHPC, the upscaled E was
found to be higher than that from experiments, likely due to poten-
as solid lines. For a residual porosity of zero, fCH is 0.25 (i.e., the
tial agglomeration of microsilica in the UHP paste. In the presence
composite is 75% HD C-S-H and 25% nano CH), whereas for a resid-
of coarse aggregates, these agglomerates would be broken down by
ual porosity of 0.25, fCH is zero. The lower and upper bounds pro-
the shearing action of aggregates, reducing the volume of what
vided in this plot correspond to indentation plane stress elastic
were identified as mixed phases. The local heterogeneity in the
modulus of the C-S-H globules of 65 GPa and 75 GPa respectively.
microstructure in such cases are more likely to be significant in
The larger value is used to account for the potential enhancement
upscale property prediction. While increasing the number of grids
in globule stiffness with decreasing C/S ratio of the gel [51,52]. The
and indents might appear to be a straightforward solution, it is
indentation modulus of CH is taken as 46 GPa [18].
unlikely to account for effects that are fundamentally different in
Considering the microsilica-based UHP paste after 30 days of
pastes and concrete – e.g., ultra-fine particle agglomeration.
hydration as an illustration (the nanoindentation stiffness is
~45 GPa, shown using the dotted line; See Table 3), if the UHS
phase is a composite of HD C-S-H and CH, the stiffness bounds indi- CRediT authorship contribution statement
cate that residual porosity of this phase should be between 0.17
and 0.24 (or fCH between 0.08 and 0.01). The fCH determined from Emily Ford: Methodology, Validation, Investigation, Formal
TGA for this mixture is 0.047, which lies in this range. Similar analysis, Writing - original draft, Data curation, Visualization.
results for the other mixtures are noted and they all fall within Aashay Arora: Investigation, Data curation. Barzin Mobasher:
or reasonably close to the bounds. Note that /HD-CSH is obtained Resources, Funding acquisition, Writing - review & editing. Chris-
from the packing density of C-S-H, which is an averaged value over tian G. Hoover: Resources, Supervision, Writing - review & editing.
many indents within a representative volume. Similarly, fCH Narayanan Neithalath: Conceptualization, Methodology, Writing -
obtained from TGA is also averaged over a tested volume. As a con- review & editing, Supervision, Project administration, Resources,
firmation, UHS stiffness data and fCH values from [18] are also plot- Funding acquisition.
ted, which is also observed to fall within the bounds. This analysis
confirms that the UHS phase can be mechanically modelled as a
Declaration of Competing Interest
composite of HD C-S-H and nanoscale CH.
The authors declare that they have no known competing finan-
4. Conclusions cial interests or personal relationships that could have appeared
to influence the work reported in this paper.
This paper has discussed the nanoscale mechanical characteri-
zation of UHP cementitious matrices containing high volumes of Acknowledgements
commonly available cement replacement materials. Nanoindenta-
tion experiments coupled with a Bayesian information criterion This study was partly supported by U.S. National Science Foun-
were carried out to determine the average modulus and hardness dation (CMMI: 1463646) and Arizona Department of Transporta-
of the statistical phases. The microstructure was found to be highly tion (ADOT: SPR 745). The first author acknowledges a Dean’s
heterogeneous, attributable mainly to the low w/b used and mul- Fellowship from Arizona State University. Omar Castillo is
titude of cement replacement materials of differing size ranges. acknowledged for his help with specimen preparation.
LD and HD C-S-H along with an UHS phase was identified for the
fly ash-based UHP paste (w/b ~ 0.20) after 30 days of hydration,
while only the UHS phase was observed for the fly ash-based paste Appendix A. Supplementary data
after 90 days of hydration. For the microsilica-based UHP paste, the
volume fractions of LD and HD C-S-H phases were so low that the Supplementary data to this article can be found online at
points were clustered together with the UHS phase at all ages. The https://doi.org/10.1016/j.conbuildmat.2020.118214.
UHS phase was shown to be a nanocomposite of HD C-S-H phase
and nanoscale CH. The reduced w/b in UHP pastes as well as the References
presence of several fine materials such as limestone powder and
microsilica was postulated to enhance the formation of such a pro- [1] A. Arora, M. Aguayo, H. Hansen, C. Castro, E. Federspiel, B. Mobasher, N.
Neithalath, Microstructural packing- and rheology-based binder selection and
duct through the absence of capillary spaces that would likely characterization for Ultra-high Performance Concrete (UHPC), Cem. Concr. Res.
favor CH nanocrystal formation. The formation of such a product 103 (2018) 179–190.
E. Ford et al. / Construction and Building Materials 243 (2020) 118214 11

[2] R. Yu, P. Spiesz, H. Brouwers, Effect of nano-silica on the hydration and [30] W. Oliver, G. Pharr, Measurement of hardness and elastic modulus by
microstructure development of Ultra-High Performance Concrete (UHPC) with instrumented indentation: advances in understanding andrefinements to
a low binder amount, Constr. Build. Mater. 65 (2014) 140–150. methodology, J. Mater. Res. 19 (1) (2004) 1–20.
[3] M.M. Norhasri, M. Hamidah, A.M. Fadzil, O. Megawati, Inclusion of nano [31] C.-J. Haecker, E.J. Garboczi, J.W. Bullard, R.B. Bohn, Z. Sun, S.P. Shah, T. Voigt,
metakaolin as additive in ultra high performance concrete (UHPC), Constr. Modeling the linear elastic properties of Portland cement paste, Cem. Concr.
Build. Mater. 127 (2016) 167–175. Res. 35 (2005) 1948–1960.
[4] W. Meng, K.H. Khayat, Mechanical properties of ultra-high-performance [32] E.W. Washburn, ‘‘Note on a method of determining the distribution of pore
concrete enhanced with graphite nanoplatelets and carbon nanofibers, sizes in a porous material,” in Proceedings of the National Academy of Sciences
Compos. B 107 (2016) 113–122. of the United States of America, vol. 7, no. 4, pp. 115–116, 1921.
[5] M. Alkaysi, S. El-Tawil, Z. Liu, W. Hansen, Effects of silica powder and cement [33] H. Alghamdi, A. Dakhane, A. Alum, M. Abbaszadegan, B. Mobasher, N.
type on durability of ultra high performance concrete (UHPC), Cem. Concr. Neithalath, Synthesis and characterization of economical, multi-functional
Compos. 66 (2016) 47–56. porous ceramics based on abundant aluminosilicates, Mater. Des. 152 (2018)
[6] S. Zhao, W. Sun, Nano-mechanical behavior of a green ultra-high performance 10–21.
concrete, Constr. Build. Mater. 63 (2014) 150–160. [34] A.A. Liabastre, C. Orr, An evaluation of pore structure by mercury penetration,
[7] C. Wang, C. Yang, F. Liu, C. Wan, X. Pu, Preparation of ultra-high performance J. Colloid Interface Sci. 64 (1) (1978) 1–18.
concrete with common technology and materials, Cem. Concr. Compos. 34 [35] D. Shi, D.N. Winslow, Contact angle and damage during mercury intrusion into
(2012) 538–544. cement paste, Cem. Concr. Res. 15 (1985) 645–654.
[8] W. Huang, H. Kazemi-Kamyab, W. Sun, K. Scrivener, Effect of cement [36] M.S. Kim, Y. Jun, C. Lee, J.E. Oh, Use of CaO as an activator for producing a price-
substitution by limestone on the hydration and microstructural development competitive non-cement structural binder using ground granulated blast
of ultra-high performance concrete (UHPC), Cem. Concr. Compos. 77 (2015) furnace slag, Cem. Concr. Res. 54 (2013) 208–214.
86–101. [37] P. Lura, P. Trtik, B. Münch, Validity of recent approaches for statistical
[9] A. Arora, Y. Yao, B. Mobasher, N. Neithalath, Fundamental insights into the nanoindentation of cement pastes, Cem. Concr. Compos. 33 (4) (2011) 457–
compressive and flexural response of binder and aggregate-optimized ultra- 465.
high performance concrete (UHPC), Cem. Concr. Compos. 98 (2019) 1–13. [38] K.J. Krakowiak, W. Wilson, S. James, S. Musso, F.-J. Ulm, Inference of the phase-
[10] D.-Y. Yoo, M.J. Kim, S.-W. Kim, J.-J. Park, Development of cost effective ultra- to-mechanical property link via coupled X-ray spectrometry and indentation
high-performance fiber-reinforced concrete using single and hybrid steel analysis: application to cement-based materials, Cem. Concr. Res. 67 (2015)
fibers, Constr. Build. Mater. 150 (2017) 383–394. 271–285.
[11] P. Li, Q. Yu, H. Brouwers, Effect of coarse basalt aggregates on the properties of [39] G. Constantinides, K.R. Chandran, F.-J. Ulm, K.V. Vliet, Grid indentation analysis
ultra-high performance concrete (UHPC), Constr. Build. Mater. 170 (2018) of composite microstructure and mechanics: principles and validation, Mater.
649–659. Sci. Eng. A 430 (2006) 189–202.
[12] L. Xu, F. Wu, Y. Chi, P. Cheng, Y. Zeng, Q. Chen, Effects of coarse aggregate and [40] J. Němeček, V. Šmilauer, L. Kopecký, Nanoindentation characteristics of alkali-
steel fibre contents on mechanical properties of high performance concrete, activated aluminosilicate materials, Cem. Concr. Compos. 33 (2011) 163–170.
Constr. Build. Mater. 206 (2019) 97–110. [41] M.J. DeJong, F.-J. Ulm, The nanogranular behavior of C-S-H at elevated
[13] C. Hu, Z. Li, Property investigation of individual phases in cementitious temperatures (up to 700 °C), Cem. Concr. Res. 37 (2007) 1–12.
composites containing silica fume and fly ash, Cem. Concr. Compos. 57 (2015) [42] G. Constantinides, F.-J. Ulm, The effect of two types of C-S-H on the elasticity of
17–26. cement-based materials: Results from nanoindentation and micromechanical
[14] L. Sorelli, G. Constantinides, F.-J. Ulm, F. Toutlemonde, The nano-mechanical modeling, Cem. Concr. Res. 34 (2004) 67–80.
signature of ultra high performance concrete by statistical nanoindentation [43] G. Constantinides, F.-J. Ulm, The nanogranular nature of C-S–H, J. Mech. Phys.
techniques, Cem. Concr. Res. 38 (2008) 1447–1456. Solids 55 (2007) 64–90.
[15] C. Hu, Z. Li, A review on the mechanical properties of cement-based materials [44] P. Mondal, S.P. Shah, L. Marks, A reliable technique to determine the local
measured by nanoindentation, Constr. Build. Mater. 90 (2015) 80–90. mechanical properties at the nanoscale for cementitious materials, Cem.
[16] M. Vandamme, F.-J. Ulm, P. Fonollosa, Nanogranular packing of C-S–H at Concr. Res. 37 (2007) 1440–1444.
substochiometric conditions, Cem. Concr. Res. 40 (2010) 14–26. [45] P.J. Monteiro, C.T. Chang, The elastic moduli of calcium hydroxide, Cem. Concr.
[17] F.-J. Ulm, M. Vandamme, C. Bobko, J.A. Ortega, K. Tai, C. Ortiz, Statistical Res. 25 (8) (1995) 1605–1609.
indentation techniques for hydrated nanocomposites: concrete, bone, and [46] C. Hu, Nanoindentation as a tool to measure and map mechanical properties of
shale, Am. Ceram. Soc. 90 (9) (2007) 2677–2692. hardened cement pastes, MRS Commun. 5 (2015) 83–87.
[18] J.J. Chen, L. Sorelli, M. Vandamme, F.-J. Ulm, G. Chanvillard, A coupled [47] F. Lavergne, A.B. Fraj, I. Bayane, J.F. Barthélémy, Estimating the mechanical
nanoindentation/SEM-EDS study on low water/cement ratio portland cement properties of hydrating blended cementitious materials: an investigation
paste: Evidence for C–S–H/Ca(OH)2 Nanocomposites, J. Am. Ceram. Soc. 93 (5) based on micromechanics, Cem. Concr. Res. 104 (2018) 37–60.
(2010). [48] J. Moon, S. Yoon, R.M. Wentzcovitch, P.J. Monteiro, First-principles elasticity of
[19] W. Wilson, J. Rivera-Torres, L. Sorelli, A. Durán-Herrera, A. Tagnit-Hamou, The monocarboaluminate hydrates, Am. Mineral. 99 (2014) 1360–1368.
micromechanical signature of high-volume natural pozzolan concrete by [49] N.K. Dhami, A. Mukherjee, M.S. Reddy, Micrographical, minerological and
combined statistical nanoindentation and SEM-EDS analyses, Cem. Concr. Res. nano-mechanical characterisation of microbial carbonates from urease and
91 (2017) 1–12. carbonic anhydrase producing bacteria, Ecol. Eng. 94 (2016) 443–454.
[20] K. J. Krakowiak, W. Wilson, S. James and F.-J. Ulm, ‘‘In situ Chemo Mechanical [50] Zechuan Yu, Ao Zhou, Denvid Lau, Mesoscopic packing of disk-like building
Characterization of Cementitious Microstructures with Coupled X Ray blocks in calcium silicate hydrate, Sci. Rep. 6 (1) (2016), https://doi.org/
Microanalysis and Indentation Technique,” in Mechanics and Physics of 10.1038/srep36967.
Creep, Shrinkage, and Durability of Concrete: A Tribute to Zdenek P. Bažant: [51] F. Pelisser, P.J.P. Gleize, A. Mikowski, Effect of the Ca/Si molar ratio on the
Proceedings of the Ninth International Conference on Creep, Shrinkage, and micro/nanomechanical properties of synthetic CSH measured by
Durability Mechanics, Cambridge, 2013. nanoindentation, J. Phys. Chem. 116 (2012) 17219–17227.
[21] C. Hu, Microstructure and mechanical properties of fly ash blended cement [52] M.A. Qomi, K. Krakowiak, M. Bauchy, K. Stewart, R. Shahsavari, D. Jagannathan,
pastes, Constr. Build. Mater. 73 (2014) 618–625. D. Brommer, A. Baronnet, M. Buehler, S. Yip, F.-J. Ulm, K. Van Vliet, R-. Pellenq,
[22] V.Z. Zadeh, C.P. Bobko, Nanoscale mechanical properties of concrete containing Combinatorial molecular optimization of cement hydrates, Nat. Commun.
blast furnace slag and fly ash before and after thermal damage, Cem. Concr. (2014) 1–10.
Compos. 37 (2012) 215–221. [53] V.L. Bonavetti, V.F. Rahhal, E.F. Isassar, Studies on the carboaluminate
[23] C. Hu, Z. Li, Y. Gao, Y. Han, Y. Zhang, Investigation on microstructures of formation in limestone filler-blended cements, Cem. Concr. Res. 31 (2001)
cementitious composites incorporating slag, Adv. Cem. Res. 26 (4) (2014) 222– 853–859.
232. [54] G. S. Barger, J. Bayles, B. Blair, D. Brown, H. Chen, T. Conway, P. Hawkins, R. A.
[24] M. Vandamme, F.-J. Ulm, Nanoindentation investigation of creep properties of Helinski, F. A. Innis, M. D. Luther, W. C. McCall, D. Moore, W. O’Brien, E. R.
calcium silicate hydrates, Cem. Concr. Res. 52 (2013) 38–52. Orsini, M. F. Pistilli, D. Suchorski and O. Tavares, ‘‘Ettringite Formation and the
[25] M. Vandamme, The Nanogranular Origin of Concrete Creep: A Performance of Concrete,” Portland Cement Association, pp. 1–16, 201.
Nanoindentation Investigation of Microstructure and Fundamental [55] A. Ipavec, R. Gabrovsek, T. Vuk, V. Kaucic, J. Macek, A. Meden, Carboaluminate
Properties of Calcium-Silicate-Hydrates, Massachusetts Institute of phases formation during the hydration of calcite-containing portland cement,
Technology, 2008. J. Am. Ceram. Soc. 94 (4) (2011) 1238–1242.
[26] J. Fu, Nano-indentation experiment for determining mechanical properties of [56] G. Puerta-Falla, M. Balonis, G. Le Saout, G. Falzone, C. Zhang, N. Neithalath, G.
typical cement phases at nano/micro-scale, IOP Conf. Ser. Mater. Sci. Eng. 439 Sant, Elucidating the role of the aluminous source on limestone reactivity in
(2018). cementitious materials, Am. Ceram. Soc. 98 (21) (2015) 4076–4089.
[27] C.G. Hoover, F.-J. Ulm, Experimental chemo-mechanics of early-age fracture [57] Kirk Vance, Matthew Aguayo, Tandre Oey, Gaurav Sant, Narayanan Neithalath,
properties of cement paste, Cem. Concr. Res. 75 (2015) 42–52. Hydration and strength development in ternary portland cement blends
[28] F.-J. Ulm, M. Vandamme, H.M. Jennings, J. Vanzo, M. Bentivegna, K.J. containing limestone and fly ash or metakaolin, Cem. Concr. Compos. 39
Krakowiak, G. Constantinides, C.P. Bobko, K.J. Van Vliet, Does microstructure (2013) 93–103, https://doi.org/10.1016/j.cemconcomp.2013.03.028.
matter for statistical nanoindentation techniques?, Cem Concr. Compos. 32 [58] D. Hou, T. Zhao, H. Ma, Z. Li, Reactive molecular simulation on water confined
(2010) 92–99. in the nanopores of the calcium silicate hydrate gel: structure, reactivity, and
[29] W. Oliver, G. Pharr, An improved technique for determining hardness and mechanical properties, J. Phys. Chem. C 119 (2015) 1346–1358.
elastic modulus using load and displacement sensing indentation [59] P. Mondal, S.P. Shah, L.D. Marks, J.J. Gaitero, Comparative study of the effects of
experiments, J. Mater. Res. 7 (6) (1992). microsilica and nanosilica in concrete, Transp. Res. Rec. 2141 (2010) 6–9.
12 E. Ford et al. / Construction and Building Materials 243 (2020) 118214

[60] W. da Silva, J. Němeček, P. Štemberk, Application of multiscale elastic [69] C.C. Yang, R. Huang, Double inclusion model for approximate elastic moduli of
homogenization based on nanoindentation for high performance concrete, concrete material, Cem. Concr. Res. 26 (1) (1996) 83–91.
Adv. Eng. Softw. 62–63 (2013) 109–118. [70] P. Trtik, B. Münch, P. Lura, A critical examination of statistical nanoindentation
[61] Y. Li, P. Wang, Z. Wang, Evaluation of elastic modulus of cement paste on model materials and hardened cement pastes based on virtual experiments,
corroded in brine solution with advanced homogenization method, Constr. Cem. Concr. Compos. 31 (10) (2009) 705–714.
Build. Mater. 157 (2017) 660–1609. [71] Z. Wu, K.H. Khayat, C. Shi, Changes in rheology and mechanical properties of
[62] X. Gao, Y. Wei, W. Huang, Effect of individual phases on multiscale modeling ultra-high performance concrete with silica fume content, Cem. Concr. Res.
mechanical properties of hardened cement paste, Constr. Build. Mater. 153 123 (2019) 105786.
(2017) 25–35. [72] N. Hernandez, J. Lizarazo-Marriaga, M.A. Rivas, Petrographic characterization
[63] S. Das, P. Yang, S.S. Singh, J.C. Mertens, X. Xiao, N. Chawla, N. Neithalath, of Portlandite crystal sizes in cement pastes affected by different hydration
Effective properties of a fly ash geopolymer: aynergistic application of X-ray environments, Constr. Build. Mater. 182 (2018) 541–549.
synchrotron tomography, nanoindentation, and homogenization models, Cem. [73] H. Jennings, The colloid/nanogranular nature of cement paste and properties,
Concr. Res. 78 (2015) 252–262. Nanotechnol. Construct. 3 (2009) 27–36.
[64] L. Brown, P.G. Allison, F. Sanchez, Use of nanoindentation phase [74] W. Shen, W. Zhang, J. Wang, C. Han, B. Zhang, J. Li, G. Xu, The microstructure
characterization and homogenization to estimate the elastic modulus of formation of C-S-H in the HPC paste from nano-scale feature, J. Sustain. Cem.
heterogeneously decalcified cement pastes, Mater. Des. 142 (2018) 308–318. Based Mater. (2019) 1–15.
[65] W. Ashraf, J. Olek, N. Tian, Multiscale characterization of carbonated [75] E. Masoero, E. Del Gado, R.J.-M. Pellenq, S. Yip, F.-J. Ulm, Nano-scale mechanics
wollastonite paste and application of homogenization schemes to predict its of colloidal C-S–H gels, R. Soc. Chem. 10 (2014) 491–499.
effective elastic modulus, Cem. Concr. Compos. 72 (2016) 284–298. [76] B. Gathier, Multiscale Strength Homogenization: Application to Shale
[66] D. P. Bentz, J. Arnold, M. J. Boisclair, S. Z. Jones, P. Rothfeld P. E. Stutzman, ‘‘NIST Nanoindentation, Massachuesetts Institute of Technology, Cambridge, MA,
Technical Note 1963: Influence of Aggregate Characteristics on Concrete 2008.
Performance,” National Institute of Standards and Technology, 2017. [77] J.A. Ortega, Microporomechanical Modeling of Shale, Cambridge
[67] M. Hori, S. Nemat-Nasser, Double-inclusion model and overall moduli of Massachusetts Institute of Technology, MA, 2010.
multi-phase composites, Mech. Mater. 14 (1993) 189–206. [78] H.M. Jennings, J.J. Thomas, J.S. Gevrenov, G. Constantinides, F.-J. Ulm, A multi-
[68] S. Nemat-Nasser, M. Hori, ‘‘Section 10 Self-Consistent, Differential, and Related technique investigation of the nanoporosity of cement paste, Cem. Concr. Res.
Average Methods,” in Micromechanics Overall Properties of Heterogeneous 37 (2007) 329–336.
Materials, North-Holland Series in Applied Mathematics and Mechanics, 1993,
pp. 325–366.

You might also like