Вы находитесь на странице: 1из 14

Available online at www.sciencedirect.

com

ScienceDirect
Acta Materialia 67 (2014) 8194
www.elsevier.com/locate/actamat

Calcium silicate hydrate from dry to saturated state:


Structure, dynamics and mechanical properties
Hou Dongshuai a,, Ma Hongyan a,, Yu Zhu b, Li Zongjin a,
a

Department of Civil and Environmental Engineering, Hong Kong University of Science and Technology, Clear Water Bay, Hong Kong
b
Department of Civil Engineering Henan Polytechnic University, Henan, China
Received 14 July 2013; received in revised form 12 November 2013; accepted 11 December 2013
Available online 24 January 2014

Abstract
Calcium silicate hydrate (C-S-H) is the most important hydration product of cement-based materials. In the nanostructure of C-S-H,
structural water molecules are distributed in the interlayer region and determine the mechanical performance of C-S-H gel. In this study,
C-S-H gels with dierent water contents expressed as the water/Ca ratio are characterized in the light of molecular dynamics. In order to
study the inuence of the water molecules, the structures of 12 C-S-H gel samples with water/Ca ratios from 0 to 0.95 are investigated. It
is found that the penetration of water molecules transforms the C-S-H gel from an amorphous to a layered structure by silicate depolymerization as the water content gradually increases. The structures are then tested for mechanical properties by simulated uniaxial tension and compression. The mechanical tests associated with structural analysis reveal that the structural water molecules can greatly
weaken the stiness and the cohesive force by replacing the ioniccovalent bond with unstable H-bond connections. By studying the tensile failure mechanism of C-S-H gels at dierent humidity levels, the disconnecting role of the structural water molecules is comprehensively interpreted. Because the interlayer water molecules prevent reconstruction of the bonds between the Caw and the silicate chains, the
plasticity of the C-S-H gels is reduced signicantly in the change from a dry state to a saturated state. In addition, the compressive
strength of a C-S-H gel in the saturated state is much larger than the tensile strength. This provides molecular evidence for the tensile
weakness of cement paste.
2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Calcium silicate hydrate (C-S-H); Water/Ca ratio; Molecular dynamics; Uniaxial tension/compression test

1. Introduction
When Portland cement reacts with water, cement
hydrates are produced. Calcium silicate hydrate (C-S-H)
gel is the dominant cement hydrate, and hence it determines the binding force of the cement paste. For the whole
service life of a cement paste/concrete, the interaction
between water and C-S-H gel determines the mechanical
properties and durability of cement-based materials. On
the one hand, the initial water content of the cement paste,
or water-to-cement ratio, inuences the microstructure of
Corresponding authors.

E-mail addresses: dhou@ust.hk (D. Hou), cehy@ust.hk (H. Ma),


zongjin@ust.hk (Z. Li).

the cement hydrates [1]; on the other hand, the water molecules, as essential elements of C-S-H gel, are sensitive to
the surrounding environment and humidity [2]. For example, when immersed in water, C-S-H gel is very likely to
reach a saturated state, while the structural water of the
C-S-H may be missing in the case of a re.
In order to understand the role of water in C-S-H gel, it
is necessary to investigate the intrinsic structure and
mechanical performance of the gel at dierent humidity
levels. The molecular structure of C-S-H has been studied
for more than half a century, particularly using a variety
of experimental techniques, including nuclear magnetic
resonance (NMR) imaging [3], X-ray diraction [4] and
small-angle neutron scattering (SANS) [5], and C-S-H is
now widely believed to be the analogue of the layered

1359-6454/$36.00 2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.actamat.2013.12.016

82

D. Hou et al. / Acta Materialia 67 (2014) 8194

minerals tobermorite [6,7] and jennite [8]. Based on experimental data, Pellenq et al. [9] employed the molecular
dynamics method to construct a C-S-H model that can
describe well the structural, dynamical and mechanical
properties of cement at the nanoscale. Because the model
is based on previous experimental results, it has been
widely accepted in the nanoscale simulation eld and is
dened as a realistic model. The CSHFF force eld
[10] was developed to describe the interaction between
atoms in cement systems, and this force eld demonstrated
good transferability in simulating cement-based materials.
Water molecules, essential constituents in C-S-H gel,
have been investigated since the birth of the realistic model. Pellenq et al. [9] simulated the stressstrain relation of
C-S-H gel in resisting shear force. By comparing the
mechanical performance of dry and wet C-S-H gels, it
was concluded that, during the loading process, large displacements of water molecules in wet C-S-H gels weaken
the shear strength and result in unrecovered deformations.
Youssef et al. [11] investigated the structural and dynamic
features of water molecules in C-S-H gels. Due to the
highly hydrophilic nature of calcium silicate sheets, Hbonds constructed between the non-bridging oxygen and
water are very strong. Therefore, water molecules constrained in the gel demonstrate a glassy nature: the tetrahedral spatial structure is distorted and the diusion rate is
signicantly reduced. Ji et al. [12] utilized ve classic water
models, SPC, TIP3P, TIP4P, TIP4P05 and TIP5P, to simulate calcium silicate hydrate. They concluded that SPC
and TIP5P accurately describe the intrinsic properties of
C-S-H gel, with SPC (exible) being more computationally
ecient. Recently, Bonnaud et al. [13] interpreted the cohesive force of C-S-H gel by analyzing the uid pressure of
the water molecules and the counter-ions in the interlayer
region. Under dierent humidity conditions, they found
that the cohesive force mainly results from negative pressure caused by the interaction between interlayer calcium
atoms and the calcium silicate sheets. Even though this
work took the amount of water in the C-S-H gel into consideration, the molecular structure of the calcium silicate
sheet was assumed to remain unchanged under dierent
humidity conditions. This assumption, ignoring the structural evolution due to the water eect, needs to be further
investigated.
Previous research on C-S-H gels can be categorized into
two classes: molecular structure analysis and thermodynamics investigation. Few eorts have been made to investigate the evolution of mechanical properties with water
content variation. In particular, the tensile strength of CS-H gels at the nanoscale, considered as the most essential
property of construction material, has not been studied so
far. The aim of this paper is to bridge the relationship
between the morphology and the mechanical performance
of C-S-H gels in dierent humidity conditions. On the
one hand, the structural feature evolution can be characterized by analyzing the silicate chain morphology and
chemical bonds in dierent water contents. On the other

hand, the mechanical properties, including the stiness


and cohesive force, can be achieved directly from uniaxial
tension/compression simulation. Combining the structural
and mechanical evolution, the water molecules role in
loading resistance can be determined and assessed.
2. Simulation method
The CSHFF force eld [10], developed for cement-based
materials, is utilized to simulate C-S-H gels at dierent
humidity states. The force eld approach has been widely
used in C-S-H simulations and has been proven to be able
to describe the structure, energy and mechanical properties
of various calcium silicate phases satisfactorily [9,1114].
The force eld parameters of Ca, Si, O and H can be
obtained from the literature [15].
2.1. Dry and saturated model
The C-S-H model in the present study is constructed
based on the procedures proposed by Pellenq et al. [9].
Firstly, the layered analogue mineral of C-S-H, tobermor without water, was taken as the initial conguraite 11 A
tion for the C-S-H model [16,17]. Silicate chains were
then broken to match the Q species distribution obtained
from NMR testing [18]. The dry disordered structure is
plotted in Fig. 1. Secondly, Grand Canonical Monte Carlo
simulation of the water adsorption is operated on the dry
disordered C-S-H structure at 300 K. The adsorption
is conducted from 0 to 100 million steps to obtain 12

Fig. 1. Initial dry C-S-H structure for water adsorption simulation.


Simulation box size: a = 21.3, b = 21.2, c = 21.9; a = 90, b = 90,
c = 90. Yellow and red bonds represents the silicate chain (SiOs); the
gray and green balls correspond to the interlayer calcium atoms (Caw) and
the layered calcium atoms (Cas), respectively. (For interpretation of the
references to color in this gure legend, the reader is referred to the web
version of this article.)

D. Hou et al. / Acta Materialia 67 (2014) 8194


23

22

Space in z direction

C-S-H gel samples with water/Ca contents from 0 to


around 0.95, respectively. The 0 step simulation gives a
dry C-S-H sample, and when the number of adsorption
steps exceeds 100 million, the C-S-H gel reaches a saturated
state [13]. The chemical formula of the saturated C-S-H
structure in the current simulation is (CaO)1.69(SiO2)1.66H2O, which is quite close to the (CaO)1.7(SiO2)1.8H2O obtained by the SANS test [5]. The molecular dynamics
simulations under constant pressure and temperature
(NPT) for 300 ps give the structures of C-S-H gel in equilibrium states. For each case, a further 1000 ps NPT run
is employed to achieve the equilibrium conguration for
structural and dynamic analysis.

21

20

19

0.0

2.2. Mechanical properties calculation

83

0.2

0.4

0.6

0.8

1.0

Water/Ca ratio

Fig. 2. Evolution of space in the z direction with water/Ca ratio.

After 12 samples from the dry state to the saturated


state had been obtained in previous simulations, described
in Section 2.1, their mechanical properties were calculated.
Youngs modulus and the tensile strength in the z direction,
describing the stiness and the interlayer cohesive force,
respectively, are obtained by uniaxial tension testing.
In order to elucidate the stressstrain relation, the C-SH samples were subjected to uniaxial tensile or compressive
loading through gradual elongation or shortening at a constant strain rate at 0.08 ps1. NPT ensembles were used for
the system throughout the whole simulation process. Take
the tension process as an example. Firstly, the simulation
box is relaxed at 300 K and coupled to zero external pressure in the x, y and z dimensions for 100 ps. Then, after the
pressures in the three directions have reached equilibrium,
the C-S-H structure is elongated in the z direction. Meanwhile, the pressures in the x and y directions are kept at
zero. The pressure evolution in the z direction is taken as
the internal stress rzz. The non-pressure setting in directions perpendicular to the tension direction, also considering Poissons ratio, can eliminate the articial constrain for
the deformation and allow the free development of tension
without any restriction. In the simulated tension process,
10,000 congurations are recorded for structural analysis.
3. Results and discussion
3.1. Molecular structures of C-S-H gel samples
3.1.1. Molecular structure transformation
After the model construction procedures, 12 C-S-H
samples were determined with water/Ca ratios of 0, 0.13,
0.3, 0.40, 0.50, 0.59, 0.66, 0.72, 0.80, 0.84, 0.86 and 0.95.
As shown in Fig. 2, with increasing water content, in the
equilibrium state, space in the z direction enlarges from
. In the saturated state, the interlayer dis18.9 to 22.6 A
, which is consistent
tance in the z direction reaches 11.3 A
with that of tobermorite [7] and the realistic model [9].
The dierences between the gel samples in the dry state
and in water-carrying states mainly result from the swelling
of C-S-H structures due to water penetration.

Three simulated C-S-H gel samples in the dry (water/


Ca = 0), partially saturated (water/Ca = 0.50) and saturated states (water/Ca = 0.95) are shown in Fig. 3. Correspondingly, the intensity proles of dierent atoms in the
three states are plotted in Fig. 4 vs. the distance in the z
direction. The combined information in Figs. 3 and 4 can
give a good understanding of the molecular structures of
the C-S-H gel samples. In the dry state, as shown in
Fig. 4a, the distributions of Cas (calcium sheet), Si and
Caw (interlayer calcium atom) overlap to a great extent
along the z direction and no obvious boundary exists
between the interlayer region and the calcium sheet, indicating a loss of the layered feature. As demonstrated in
Fig. 3a, the silicate chains are distributed across the z direction in a disorderly manner, which slows the amorphous
nature of the structure in the form of silicate glass [19].
In the half saturated state (water/Ca = 0.5), along the z
direction, the alternative maxima of Cas, Si, Caw and Ow
(oxygen atoms in water molecules) in the density proles,
as shown in Fig. 4b, imply that C-S-H gel has a sandwich-like structure. It can be clearly observed in Fig. 3b
that Cas and Os (oxygen atoms in the silicate structure)
form CasOs octahedrals, thereby constructing a Cas sheet.
Defective silicate chains are grafted on both sides of the Cas
sheets, and the Caw and the water molecules are distributed
between the neighboring calcium silicate sheets. In the saturated state (water/Ca = 0.95), as compared to the half saturated state, sharper density distributions of Cas and Si
(Fig. 4c) indicate a much more ordered arrangement of
the calcium silicate sheets. The intensity of Ow is distributed across the z direction, which means that water molecules are not only present in the interlayer regions but
also diuse into the defective region of calcium silicate
sheet, as shown in Fig. 3c. Furthermore, the silicate chains
also grow along the y direction with good organization, as
widely found in the mineral analogues tobermorite and jennite [8]. From the above analysis, it can be seen that water
molecules penetrate into the structure of the C-S-H gel and
transform the amorphous glassy structure to a layered
structure as a water/Ca ratio increases from 0 to 0.95.

84

D. Hou et al. / Acta Materialia 67 (2014) 8194

Fig. 3. Snapshots of the molecular structures of C-S-H gel samples from the dry state to the saturated state: (a) water/Ca = 0; (b) water/Ca = 0.5; (c)
water/Ca = 0.95. The red and white sticks corresponding to water molecules. (For interpretation of the references to color in this gure legend, the reader
is referred to the web version of this article.)

Fig. 4. Density proles of Cas, Si, Ow, Caw atoms: (a) water/Ca = 0; (b) water/Ca = 0.5; (c) water/Ca = 0.95.

3.1.2. Silicate chains morphology


Silicate chain morphologies demonstrate discrepancies
in the dry and saturated samples. In the silicate glass structure, Qn (n = 0, 1, 2, 3, 4) is an important parameter to
describe the silicate connections, where n is dened as the
number of connected neighboring silicate tetrahedrons.
Q0 is the monomer; Q1 represents a dimer structure (two
connected silicate tetrahedrons); Q2 is a long chain; Q3 is
a branch structure; and Q4 is a network structure [20]. In
the saturated samples, as shown in Fig. 5c, Q1 is the dominant phase and dimer structures develop in an orderly
fashion along the y direction. The presence of water
molecules prevents neighboring silicate chains from
approaching. However, when the water/Ca ratio is low, a
polymerization reaction can result in short silicate chain
connections and thereby change the morphology of the silicate chains. On the one hand, in the dry state (shown in
Fig. 5a), bridging silicate tetrahedrons can link with the
surrounding monomers and connect with three neighboring tetrahedrons. The solgel reaction produces Q3 species,
which are widely distributed in the amorphous silicate glass
together with cation atoms [19]. The Q3 species, i.e. the
branch structures, indicate the glassy nature of the dry CS-H gel. It should be noted that the Q3 species can bridge
both the intra- and interlayer silicate chains. Unlike the silicate chain growth along the y direction in the saturated
samples, the silicate chain in the dry sample transforms

into the branch structure and develops along two dimensions (y and z directions). Similarly, the Q3 species has also
been detected experimentally in the mineral analogue of C [21], which contains no water
S-H gel, i.e. tobermorite 9 A
molecules. On the other hand, as shown in Fig. 5b, when
the water/Ca ratio reaches 0.3, two dimers and one monomer can grow together to develop a long silicate chain
across the interlayer region. Even though the Q3 species
disappears due to the increasing water content, silicate
chains can still grow in two dimensions. The polymerization process is the evolution from short silicate chains
(Q0, Q1) to long chains (Q2).
From the analysis in this section, it can be seen that, in
the C-S-H gel, water molecules can isolate the neighboring
calcium silicate chains and prevent further polymerization
of defective silicate structures. The depolymerization role
of the water molecules has been explained in previous
silica synthesis simulations [22]. The hydration of cement,
which produces C-S-H gel, is based on the solgel reaction
in the presence of calcium ions. Water molecules associated with the non-bridging oxygen atoms in silicate
chains prevent connections between isolated silicate structures, and cause decay in the polymerization process. In
the case of low water/Ca ratio, a further polymerization
process contributes to the formation of the amorphous
silicate morphology and weakens the layered structure
of C-S-H gel.

D. Hou et al. / Acta Materialia 67 (2014) 8194

85

Fig. 5. Morphology change of silicate chain: (a) polymerization in the dry sample; (b) polymerization in the sample with water/Ca = 0.3; (c) morphology
of the silicate chain in the saturated state.

3.1.3. Calcium sheet and interlayer calcium


Calcium atoms in C-S-H gel have complicated local
structures. According to the geometric location, calcium
atoms are categorized as sheet calcium (Cas) or interlayer
calcium (Caw). The nearest neighboring oxygen atoms of
the two types of Ca are both from Ow (O in water) and
from Os (O in silicate chains). A radial distribution function (RDF) describes the local structure of CaO of the saturated sample. To better understand the structure features
of CaO, the RDF (CaO) is further dierentiated into
four pairs, as shown in Fig. 6. For bonded atoms, the relevant bond distances can be readily determined by the positions of the peaks in the corresponding RDF. The lengths
of the four CaO bonds are ranked in the following order:
CasOs < CawOw < CawOs < CasOw.
The bond length dierence reects the local environment
of Ca atoms with respect to chemical bonds. On the one

hand, silicate chains graft onto the calcium sheet and, in


comparison with Caw, Cas atoms are more likely to associate with Os in the silicate chain. Hence, CasOs has a
shorter bond length than CawOs. On the other hand,
water molecules energetically prefer the Caw and have
fewer chances to form bonds with Cas, contributing to
the shorter bond length of CawOw.
The coordination number of Ca atoms is dened as the
from the
total neighboring O atoms that are within 3 A
central Ca atom. The CaO bond length is dened as
, as used in many references using CSHFF [9,10,15].
3A
As shown in Fig. 7, the chemical bonds of both CasOs
and CawOs reduce with increasing water content. This is
more pronounced in the CawOs bond because, in a dry
sample, 6.5 Os atoms can form chemical bonds with the
nearest Caw, whereas more than 35% of the neighboring
Os atoms are substituted by Ow atoms in the saturated case.

Cas-Os
Cas-Ow
Caw-Os
Caw-Ow

CN of Ca

0
0.0

0.2

0.4

0.6

0.8

1.0

Water/Ca ratio

Fig. 6. RDF of CaO pairs.

Fig. 7. Average coordination number of Ca and Caw with increasing


water content.

D. Hou et al. / Acta Materialia 67 (2014) 8194

Therefore, it can be observed from Fig. 7 that the numbers


of CasOw and CawOw bonds per calcium atom increase
to around 1 and 2.5, respectively, in the saturated state.
The variations in the coordinated oxygen atoms result in
dierent bridging bonds between neighboring calcium silicate sheets. The chemical bond evolution is illustrated schematically in Fig. 8. In the case of dry samples, the ionic
bonds of OsCawOs can connect directly to the nearby silicate chains. However, since the maximum CawO bond
, it is dicult to form an OsCawOs
length is about 3 A
connection, as the distance to the neighboring Os atoms
. In this case, the structural water molecules
exceeds 6 A
can play a role in bridging near the Caw by CawOw bonds.
When the water/Ca ratio further increases, multiple layers
of water molecules are packed between the calcium silicate
sheets. An H-bond network is developed between neighboring water molecules and the surrounding Os atoms.
3.1.4. H-bonds in the interlayer region
In addition to the SiO and CaO bonds, the H-bond is
another important chemical connection in the C-S-H gel.
Water molecules can form H-bonds both with silicate
chains and with neighboring water molecules. RDFs of
OwHw and OsHw describe local structures of the two
types of H-bond connections. As shown in Fig. 9, the rst
, which is shorter
peak value of OsHw occurs at 1.64 A
of OwHw. This implies that Os atoms
than the 1.75 A
are energetically preferable to the surrounding Hw atoms.
, the dierence
In the medium range, from 2 to 4.5 A
between the two RDF curves is more obvious. Double
and a single peak
peaks of OsHw occur at 3.1 and 4.0 A
. Douof OwHw with larger intensity occurs only at 3.1 A
ble peaks mainly result from the fact that water molecules
are geometrically constrained by the surrounding silicate
chains, and the spatial correlation between Os and Hw
can be maintained in the medium range. Compared to
the constraint eect from the silicate chain, the restrained
inuence from neighboring water molecules is much
weaker and the arrangement of water molecules in the
medium range is not well ordered.

2. 0

Os-Hw
Ow-Hw
1. 5

RDF (O H)

86

1. 0

0. 5

0. 0
0

10

Distance (Ang)

Fig. 9. RDF of OH in saturated sample.

The average number of H-bonds per water is calculated


as a function of water/Ca content. According to the literature [23], the formation of an H-bond requires two conditions: the distance between two water neighbors DHO
and the angle h between the
should be less than 2.45 A
OH vector and OO vector should be less than 30. As
shown in Fig. 10, the bond number of OsHw gradually
decreases from around 1.6 to 1.2, while the number of
OwHw increases from 0.3 to 1.18, as the water/Ca ratio
increases from 0.13 to 0.95. When the water/Ca ratio is
low, OsHw accounts for the majority of the H-bonds.
Water penetration into the C-S-H gel can signicantly
reduce the OsHw ratio, and H-bonds are progressively
substituted by OwHw bonds. Both OsHw and OwHw
occupy 50% of the H-bonds in the saturated sample. As
a comparison, 1.2 and 1.18 H-bonds connected by silicate
chains and neighboring water molecules, as seen in
Fig. 10, are consistent with the H-bond number calculated
in Ref. [11] using Kumar et al.s method [24].
The evolution of H-bond types also inuences the bridging between nearby calcium silicate sheets, as schematically
2. 0

O s -H w
Ow-Hw

H-bonds/water

1. 6

1. 2

0. 8

0. 4

0. 0
0.2

0.4

0.6

0.8

1.0

Water/Ca

Fig. 8. Schematic diagrams of the interlayer connections.

Fig. 10. Evolution of average H-bond number with water/Ca.

D. Hou et al. / Acta Materialia 67 (2014) 8194

87

(TCF) is utilized to describe dynamical properties of various chemical bonds.


The TCF, C(t), of a bond is described by Eq. (1):
Ct

Fig. 11. Schematic diagrams of the interlayer H-bond connections.

illustrated in Fig. 11. At a low water/Ca ratio, OsHw


bonds can be distributed across the interlayer region and
directly bridge the neighboring calcium silicate sheets. With
increasing water content, the water molecules in the interlayer region transform from single-layer to multi-layer
packing. In the case of the single layer, water molecules
can bridge the neighboring calcium silicate sheets by Hbonds of both HwOs and HwOw, while, in the case of
multi-layers, the water connects indirectly to the surface
water molecules by HwOw bonds. In this way, the connection between neighboring calcium silicate sheets is screened
by the increasing number of water layers.
3.1.5. Stability of chemical bonds
The previous discussion describes the molecular structure of C-S-H gel in both dry and saturated states in terms
of chemical bonds. As shown in the snapshots of the interlayer region in Fig. 12, the chemical bonds of SiOs, Cas
Os, CawOs, CawOw, HwOw and OsHw play roles in
bridging the neighboring calcium silicate sheets. The
strengths of the bonds depend to a large extent on the stability of the chemical bonds. As in previous research studying H-bond strength [11], the time-correlated function

< dbtdb0 >


< db0db0 >

where db(t) = b(t)  <b>, b(t) is a binary operator that


takes a value of 1 if the pair (e.g. CaO) is bonded and 0
if not, and <b> is the average value of b over all simulation
times and pairs. The chemical bonds break and form as
time passes, which result in the connectivity variation in
C-S-H gels. If the connectivity persists unchanged, the
TCF of the bonds will maintain a constant value of 1.
Otherwise, the breaking of the bonds leads to lower TCF
values, and the more frequent the bond breakage, the lower
the TCF value. By comparing the deviations from 1 in the
TCF curves, the stability of various bonds can be
estimated.
Fig. 13 demonstrates the evolution of the TCFs of
SiOs, CasOs, CawOs, CawOw, HwOw and OsHw in
1000 ps. The chemical bonds can be categorized into two
types: bonds connected with Os and bonds connected with
water. The C(t) values of Os-connected bonds, including
SiOs, CasOs and CawOs bonds, almost maintain a constant value with limited uctuations, which indicates that
they are stable chemical connections. In contrast, C(t)
values of Ow-connect bonds, like CawOw, HwOw and
OsHw bonds, reduce signicantly with time. For H-bonds
of HwOw especially, the C(t) value decreases to less than
0.5, implying that the connectivity of more than half of
the bonds changes. According to the reduction extent of
C(t), the strength of chemical bonds is ranked in the following
order:
SiOs > CasOs > CawOs > CawOw >
HwOs > HwOw.
The stability of the bonds can be explained by the mobility of dierent atoms in C-S-H gel. The mean square displacement MSD(t) [25], a parameter used to estimate the
dynamic properties of water molecules, can be dened by
Eq. (2):
2

MSDt < jri t  ri 0j >

Fig. 12. Chemical bonds in interlayer region.

88

D. Hou et al. / Acta Materialia 67 (2014) 8194

3.2. Mechanical properties


1. 0

C (t)

0. 8

0. 6

0. 4

Si-O
Ca-O
Caw-O
Caw-Ow
Ow-Hw
Hw-O

0. 2

0. 0
0

200

400

600

800

1000

Time (Ps)

Fig. 13. TCF of various chemical bonds in C-S-H gel.

Ow
Caw
Cas
Si

MSD (Ang )

0
0

200

400

600

800

1000

Time (Ps)

Fig. 14. MSD of Ow, Caw, Cas and Si in 1000 ps.

where ri(t) represents the position of atom i at time t, ri(0) is


the original position of atom i and MSD takes into account
three-dimensional coordinates. A large MSD value at time
t indicates that the atoms diuse rapidly and are displaced
far away from the original position. Fig. 14 exhibits that
the diusion rate of atoms ranks in the following order:
Ow > Caw > Cas > Si. It can be seen from Fig. 14 that the
MSD of Ow (water) at 1000 ps is more than four times that
of the other three atoms. The high diusion rate of water
molecules results in the frequent change of the chemical
bonds with water. It is worth noting that the MSD of
Caw has a similar variation trend to that of water. This is
attributed to the fact that Caw forms clusters with water
in the interlayer region, and the diusion rate is thus accelerated to some extent. In addition, in the calcium silicate
sheet, the diusion rate of Si is lower than that of Cas.
The dynamic properties of the SiO bonds and CaO
bonds revealed match previous ndings using the ab initio
method well [26]. The ndings prove that SiO bonds in silicate chains have more binding energy than CaO bonds
and are hard to break.

Previous structural and dynamics analyses provide a


clear picture of chemical bonds in C-S-H gels. Due to the
strong chemical strength and low diusion rate of Si, the
silicate chains provide the most stable backbone of C-SH gels. Cas atoms are associated with the neighboring Os
in the silicate chains to produce high-strength calcium
silicate sheet. In dierent humidity states, the Caw atoms
associated with various amounts of interlayer water play
an important role in connecting the neighboring calcium
silicate sheets. The mechanical properties of C-S-H gels
in dierent humidity states are determined by the combination of these chemical bonds. In light of the method introduced in Section 2.2, the mechanical properties of C-S-H
gel are discussed in this section, and explained in terms
of the chemical bonds.
3.2.1. Stressstrain relation
Stressstrain curves are used to assess the mechanical
performance of C-S-H gels in the whole tensile loading process. Tensile and compressive stressstrain relations of the
12 samples at dierent humidity states are plotted in
Figs. 15 and 16, respectively. Fig. 15 shows two types of
failure mode during the tensile process. When the water/
Ca ratio is less than 0.6 (samples 15), the samples demonstrate good plasticity. During the tensile process, the stress
increases up to the failure strength at a strain of around
A
1, then reduces slowly and uctuates in a saw-like
0.1 A
style. The structure of the dry C-S-H gel (sample 1) cannot
be directly stretched to fracture even when the strain
A
1. In contrast, when a samples water/Ca
reaches 0.8 A
ratio exceeds 0.6, it is more likely to be stretched to fracture. As shown in Fig. 15, in samples 612 the stress
increases to the maximum quickly at a strain of less than
A
1 and subsequently drops to zero. With increasing
0.1 A
water/Ca ratio from 0.6 to 1, the strain at fracture reduces
A
1. That is to say, the interlayer water
from 0.8 to 0.4 A
molecules make the C-S-H gel more brittle.
In case of the compressive stressstrain relation in
Fig. 16, two distinguished failure modes can also be
observed. When the water/Ca ratio is less than 0.5 (samples
14), the samples demonstrate good plasticity. During the
compressive process, the stress increases to the maximum,
then uctuates at the failure strength until the strain
A
1. In the post-failure region, the
reaches around 0.35 A
stress can maintain at a long-strain plateau value without
obvious reduction. On the other hand, when the samples
water/Ca ratio is larger than 0.5, the stressstrain plateau
disappears and the stress reduces quickly in the post-failure
stage.
3.2.2. Youngs modulus and tensile strength
The Youngs modulus, tensile strength and compressive
strength, obtained from the stressstrain curves, are important quantitative parameters for estimating the mechanical
properties of C-S-H gels. As shown in Fig. 17a, with a

D. Hou et al. / Acta Materialia 67 (2014) 8194

89

Fig. 15. Stressstrain relations of 12 samples under uniaxial tension loading along the z direction.

progressively increasing water/Ca ratio from 0 to 1,


Youngs modulus reduces from 67 to less than 47 GPa.
The Youngs moduli obtained in the current simulations
are consistent with the results from a nanoindentation test
(60 GPa) [27] and previous ab initio calculations (55
68 GPa) [9]. It should be noted that the moduli of samples
with water/Ca ratios of <0.3 are maintained at around
67 GPa, much larger than other samples. It is known from
previous structural analyses that having less water can contribute to further polymerization of the silicate chains, thus
the three samples may have Q3 species or developed silicate
chains in the z direction which achieve the high moduli.
When the water/Ca ratio is high, the silicate chains which
act as the backbone of C-S-H are shorter, and the water
molecules themselves can also signicantly weaken the stiness of the C-S-H gel. In terms of chemical bonds, in dry CS-H gel, the Caw atoms ll in the defects of the silicate
chains, which contribute to the CawOs connection. As
mentioned above, following an increase in the water/Ca

ratio, water molecules penetrate and replace the partial


CawOs connections with H-bonds, which can soften the
C-S-H gel. Furthermore, water molecules adsorbed on
the defective silicate chains result in larger interlayer distances and thus a smaller coulombic attraction between
neighboring calcium silicate layers.
Similar evolution trends in the tensile strength can be
observed in Fig. 17b, decreasing from 7.5 GPa in the dry
state to 3.8 GPa in the saturated state. The failure stresses
obtained are consistent with the rupture strength (3 GPa)
by shear simulation [9] and the uid pressure (2.75 GPa)
calculation [13]. The huge dierence in C-S-H gel tensile
strength between the dry and saturated states indicates different failure mechanisms. The chemical bonds in the C-SH gel determine the mechanical performance in dierent
humidity states to a large extent. When C-S-H gel is transformed from a dry state to a saturated state, the stable and
high-strength CaOs bonds are partially substituted by
unstable CaOw bonds and H-bonds, while the structure

90

D. Hou et al. / Acta Materialia 67 (2014) 8194

Fig. 16. Stressstrain relations of 12 samples under uniaxial compression loading along the z direction.

transforms from an amorphous glass state to a layered


structure. The variations in the chemical bonds and structure that play important roles in bridging the calcium silicate sheets result in a weakening of the cohesive force. In
view of the uid pressure, Bonnaud et al. proposed that
water molecules have a disjoining eect in C-S-H grains
[13]. The uid pressure in C-S-H gel is contributed by the
cohesive force from the Ca2+ counter-ions and the repulsive force by the water molecules. From the dry state to
the saturated state, the repulsive eect of the water reduces
the cohesive force of the C-S-H gel.
As shown in Fig. 17c, the compressive strength of samples with dierent water contents varies from 8 to 6 GPa.
No pronounced reduction trend can be observed, implying
that the water molecules have little inuence on the compressive performance of the C-S-H gel. Furthermore, in

the saturated state, the compressive strength of the C-SH gel is about twice as large as the tensile strength. It
should be noted that, at the macrolevel, the cement paste
is strong in compression and weak in tension. The
discrepancy between the compressive and tensile behavior
of C-S-H gel at the nanolevel gives molecular insight into
the tensile weakness of cement paste.
3.2.3. Structural analysis
In order to support the discussions on the relation between
mechanical properties and chemical bonds, structural analysis is required. The strain-correlated function (SCF) is used
to describe the variation in chemical bonds in comparison
with the original structure during the tensile process. The
SCF can be interpreted as a TCF in the presence of strain
loading, and is dened in a similar way to TCF as

D. Hou et al. / Acta Materialia 67 (2014) 8194


70

Young' s modulus in z (GPa)

65

60

55

50

(a)

45

0.0

0.2

0.4

0.6

0.8

1.0

0.8

1.0

Water/Ca ratio

Tensile strength (GPa)

(b)
0.0

0.2

0.4

0.6

Water/Ca

Compressive strength (GPa)

(c)
0
0.0

0.2

0.4

0.6

0.8

1.0

water/Ca

Fig. 17. Evolution of mechanical properties of C-S-H gel with water/Ca


ratio: (a) Youngs modulus; (b) tensile strength; (c). compressive strength.

Ce

< dbedb0 >


< db0db0 >

where e denotes the strain. During uniaxial tensile testing,


the chemical bonds can break and form frequently to resist
the loading, which results in structural transformation. As
shown in Fig. 18a, in the case of a dry C-S-H sample, the

91

C(e) values of CasOs bonds and CawOs bonds decrease


to 0.4, implying a large amount of breakage of these chemical
bonds. In contrast, the slight reduction in C(e) of SiOs
bonds indicates that most of the silicate chains maintain their
integrity. On the other hand, in the saturated state, as shown
in Fig. 18b, the C(e) of CawOs bonds reduces more than that
of the CasOs bonds and the C(e) of the SiOs bonds remains
constant. This means that the CawOs bonds are more likely
to be broken in the presence of water molecules. As discussed
previously, CasOs, SiOs and CawOs are the main contributors to the mechanical performance of C-S-H gel from the
dry state to the saturated state. Silicate chains act as the backbone of the C-S-H gel, and only at large deformation in the
dry state can the SiOs bonds be slightly damaged. Regarding the CasOs and CawOs bonds, their variations
determine the tensile strength of the C-S-H gel. With increasing amounts of water, the role of CawOs bonds becomes
more and more critical.
More information can be obtained from the molecular
structural evolution of C-S-H gel in resisting tensile loading,
as shown in Fig. 19. It can be seen from Fig. 19a that, when
the stain is small, even though the silicate chain distribution
is amorphous, the locations of the Cas and Caw atoms can
be clearly distinguished in the z direction. Only a small
amount of the Q3 species distributed across the interlayer
region can be stretched. When the strain reaches a moderate
A
1, local structures become more disordered
level of 0.3 A
and small nanocracks are present where bonds breakages
occur. When the strain level increases further, Caw atoms
diuse into the defective regions of the calcium silicate sheets
and some debonded silicate chains grow into the interlayer
region. Meanwhile, small cracks coalescence both in the interlayer region and in the calcium silicate sheets, which conrms
that the CasOs bonds play as important a role as CawOs
bonds in the loading resistance, as illustrated by the C(e)
shown in Fig. 18a. Caw and Cas atoms are mixed together
and silicate chains are stretched along the z direction, with
both contributing to the plasticity of the dry C-S-H gel. The
saw-like stressstrain relationship shown in Fig. 15 can be
attributed to the merger of Caw and silicate chains. In the
saturated state, as shown in Fig. 19b, the tensile damage of
the C-S-H gel exhibits a dierent failure mechanism. During
the tensile process, the breakage of the CawO bonds
and the H-bonds results in the separation of neighboring calcium silicate sheets. As the strength of the H-bond is weak, the
H-bonds between the water and the neighboring calcium
silicate sheets are formed and broken frequently. Therefore,
as shown in Fig. 19b, nanocracks occur along the interlayer
region where the H-bonds are rich, and coalesce to cause
complete fracture of the layered structure. In addition, the
calcium silicate sheets maintain their original layered
morphology, which indicates only a slight contribution of
the CasOs bond in loading resistance. The failure mode of
interlayer rupture is consistent with previous results by
SCF, as shown in Fig. 18b, i.e. CawOs plays a major role in
resisting the tensile loading and in determining the tensile
strength.

92

D. Hou et al. / Acta Materialia 67 (2014) 8194


1.0

0.8

0.8

0.6

0.6

C ( )

C ()

1.0

0.4

0.4

Cas-Os
Si-Os
Caw-Os

0.2

0.0
0.0

0.2

0.2

(a)
0.4

0.6

0.8

Strain (Ang/Ang)

0.0
0.0

Cas-Os
Si-Os
Caw-Os
0.2

(b)
0.4

0.6

0.8

Strain (Ang/Ang)

Fig. 18. SCFs of CaO, SiO and CawO bonds in dierent states: (a) dry state; (b) saturated state.

Fig. 19. Molecular structural evolution of C-S-H gel in resisting tensile loading in dierent states (the strains from top to the middle and the bottom are
small, moderate and large, respectively): (a) dry state; (b) saturated state.

Compressive loading additionally results in molecular


structure deformation, as shown in Fig. 20. For the dry
C-S-H sample, illustrated in Fig. 20a, the compressive
loading reduces the interlayer distance so that the calcium
sheet and silicate chains merge together. Meanwhile, the
morphology of the silicate chain changes during the compressive process: some silicate chains grow across the interlayer region and some react with neighboring silicate
chains to enhance the polymerization degree. The morphology variation indicates better plasticity, which can explain
the long strain plateau in the stressstrain relation, as
shown in Fig. 16. On the other hand, for the saturated

C-S-H sample, exhibited in Fig. 20b, the interlayer water


molecules prevent the silicate species from polymerizing
so that the layered arrangement can be maintained even
at a high strain state. In addition, the loading shortens
the interlayer distance, reducing the diusion rate for water
molecules. Thus the H-bond network is hard to break due
to compressive loading, which contributes to the high compressive strength for the C-S-H gel at saturated state.
Therefore, the dierent tensile and compressive behaviors
of interlayer water molecules determine the large strength
discrepancy of C-S-H gel when under tension and when
compressed.

D. Hou et al. / Acta Materialia 67 (2014) 8194

93

Fig. 20. Molecular structural evolution of C-S-H gel in resisting compressive loading in dierent states (he strains from top to the middle and the bottom
are small, moderate and large, respectively): (a) dry state; (b) saturated state.

This structural analysis proves that the hydrolytic weakening found in the silicate composite also occurs in C-S-H
gels. In the dry sample, in the interlayer region, the coordinated atoms of Caw are Os and CawOs connections have a
high potential energy. In addition, even though the breakage of local CawOs bonds occurs, Ca atoms can immediately reconstruct the chemical bonds with neighboring Os.
The reconstruction of the chemical bonds contributes to
the recovery of small defects in the elastic region. Nevertheless, in the saturated sample, interlayer Ow atoms substitute
partial Os atoms and form unstable CawOw bonds. Due to
the disjoining inuence of the water molecules, the broken
CaO connection cannot be reconstructed as easily as in
dry C-S-H gel. Therefore, saturated C-S-H gel has a more
brittle structure.
4. Conclusion
By considering the molecular dynamics, the structure,
dynamics and mechanical performance of C-S-H gels at
dierent states have been investigated. The following
conclusions can be drawn from this study.

(1) Structurally, as the water content increases, the penetration of the water transforms the C-S-H gel from an
amorphous to a layered structure. Meanwhile, the
bridging bonds between the calcium silicate sheets
vary from only CawOs bonds to a combination of
CawO bonds and H-bonds. That is to say, the presence of water molecules weakens the stability of the
chemical bonds in the system.
(2) The morphology and chemical bond variation
induced by the water determine the mechanical
properties evolution. Both the stiness and the
cohesive force are reduced with increasing water
content.
(3) In the tensile loading process, the amorphous dry
sample, due to the stretching of the silicate branch
and the breakage of the CasOs bonds, demonstrates
high stiness, high strength and plasticity in the
post-failure stage. In contrast, in the saturated state,
the much weaker CawOs bonds and H-bonds play
major roles in the loading resistance, which leads
to lower stiness and strength, and makes the gel
more brittle.

94

D. Hou et al. / Acta Materialia 67 (2014) 8194

(4) In the loading process, some broken bonds can be


reconstructed. The presence of water molecules prevents this reconstruction to some extent, which is
another mechanism that inuences the mechanical
properties of C-S-H gel.
(5) In the compressive process, the H-bond network is
hard to break so that the compressive strength of
C-S-H gel in the saturated state is much larger than
the tensile strength at the molecular level.

Acknowledgements
Support from the China Ministry of Science and Technology under grant 2009CB623200 is gratefully acknowledged.
References
[1] Ma H, Li Z. Comput Concr 2013;11(4):31736.
[2] Li Z. Advanced concrete technology. Hoboken, NJ: John Wiley &
Sons; 2011.
[3] Cong X, Kirkpatrick R. Adv Cem Mater 1996;3(34):14456.
[4] Janika JA, Kurdowsk W, Podsiadey R, Samset J. Acta Phys Polonic
2001;100:52937.
[5] Allen AJ, Thomas JJ, Jennings HM. Nat Mater 2007;6:3116.
[6] Merlino S, Bonnacorsi E, Armbruster T. Eur J Mineral 2001;13(3):
57790.
[7] Hamid S. Zeitschrit fur Kristallographie 1981;154(3-4):18998.
[8] Bonnacorsi E, Merlino S, Taylor H. Cem Concr Res 2004;34(9):
14818.

[9] Pellenq RJ, Kushima A, Shahsavari R, Van Vliet KJ, Buehler MJ,
Yip S. PNAS 2009;106(38):161027.
[10] Shahsavari R, Pellenq RJM, Ulm FJ. Phys Chem Chem Phys
2011;13(3):100211.
[11] Youssef M, Pellenq RJM, Yildiz B. J Am Chem Soc 2011;133(8):
2499510.
[12] Ji Q, Pellenq RJM, Van Vliet KJ. Comput Mater Sci 2012;53(1):
23440.
[13] Bonnaud PA, Ji Q, Coasne B, Pellenq RJ-M, Van Vliet KJ. Langmuir
2012;28(31):1142232.
[14] Manzano H, Moeini S, Marinelli F, van Duin A, Ulm F, Pellenq R. J
Am Chem Soc 2011;134(4):220815.
[15] Shahsavari R. PhD thesis. Massachusetts Institute of Technology;
2011.
[16] Murray SJ, Subramani VJ, Selvam RP, Hall KD. J Transport Res
Board 2010;2142(11):7582.
[17] Selvam RP, Subramani VG, Murray S, Hall K. Project Report
MBTC DOT 2095/3004; 2009.
[18] Chen JJ, Thomas JJ, Taylor HFW, Jennings HM. Cem Concr Res
2004;34:1499519.
[19] Mead RN, Mountjoy G. J Phys Chem 2006;110(29):2738.
[20] Feuston BP, Garofalini SH. J Phys Chem 1990;94(13):53516.
[21] Merlino S, Bonaccorsi E, Armbruster T. Eur J Min 2000;12(2):
41129.
[22] Rao NZ, Gelb LD. J Phys Chem B 2004;108(33):1241828.
[23] Wang JW, Kalinichev AG, Kirkpatrick RJ. Geochim Cosmochim
Acta 2004;68(16):335165.
[24] Kumar R, Schmidt J, Skinner J. J Chem Phys 2007;126(20):204107.
[25] Kerisit S, Liu CX. Environ Sci Technol 2009;43(3):77782.
[26] Shahsavari R, Buechler MJ, Pellenq RJM, Ulm FJ. J Am Ceram Soc
2009;92(10):232330.
[27] Costantinide G, Ulm F. J Mech Phys Solids 2006;55(1):6490.

Вам также может понравиться