Understanding The Effect of Moisture On Interfacial Behaviors of

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Construction and Building Materials 385 (2023) 131404

Contents lists available at ScienceDirect

Construction and Building Materials


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

Understanding the effect of moisture on interfacial behaviors of


geopolymer-aggregate interaction at molecular level
Zhongnan Tian a, b, Zhengqi Zhang a, b, *, Xiuming Tang c, Yingnan Zhang d, Zengjian Gui c,
Junqing Tan c, Qingxi Chang c
a
School of Highway, Chang’an University, Xi’an 710064, Shaanxi, China
b
Key Laboratory of Highway Engineering Education, Chang’an University, Xi’an 710064, Shaanxi, China
c
Tangshan Meitong Road & Bridge Engineering Co Ltd, Tangshan, Hebei 063000, China
d
Liaoning Transportation Research Institute Co Ltd, Shenyang, Liaoning 110000, China

A R T I C L E I N F O A B S T R A C T

Keywords: The interaction between geopolymer and aggregate largely determines the mechanical properties and durability
Geopolymer concrete of the geopolymer concrete. The effects of moisture on interfacial behavior of geopolymer-aggregate interaction
Molecular dynamics are poorly understood, especially at molecular level. Herein, molecular dynamics (MD) simulation was employed
Moisture
to reveal the interactive behaviors of geopolymer-aggregate interfacial system with the participation of moisture.
Aggregate
Interfacial characteristics
Full atomistic models adopted for MD simulations were constructed using the sodium aluminum silicate hydrate
Mechanical behavior (N-A–S–H) gel model and the main chemical components of the aggregates, SiO2 and CaCO3. Then the wetting
characteristics of aggregate surfaces, interfacial characteristics and mechanical behaviors of the geopolymer-
aggregate interfacial systems containing interfacial moisture were elucidated and compared. It is found that
the SiO2 surface is hydrophobic while the CaCO3 surface exhibits hydrophilic characteristics. Interfacial moisture
participates in electrostatic interaction, H-bond interaction and coordination interaction in geopolymer-
aggregate interface area. Appropriate interfacial water is beneficial to the interfacial interaction of
geopolymer-aggregate system, but excessive water will increase the risk of interfacial failure. The interfacial
moisture affects the diffusion behavior of water molecules and Na+ ions in geopolymer to the interfacial region,
and the formation of H-bonds and coordination bonds at the interface. Mechanically, with the participation of
interfacial moisture, the geopolymer-SiO2 interfacial system possesses stronger tensile strength, and a greater risk
of shear failure than that of geopolymer-CaCO3. The above atomic-level findings may facilitate a better design
and fabrication of geopolymer concrete in engineering.

1. Introduction much lower global warming impact than cement concrete, foreboding
well for its potential use in carbon-friendly concrete [7–9].
Geopolymer binders, as a potentially cleaner substitute for ordinary Concrete is a multi-phase composite material, which is mainly
silicate cement (OPC), have attracted more attention from academia and composed of cementitious hydration products, aggregate and the
industry, because geopolymers can utilize wastes in their manufacturing interaction area in between, called the interfacial transition zone (ITZ).
process. Various industrial by-products can be utilized as solid pre­ The ITZ is located at the interface between cementitious binder and
cursors for geopolymer production, including fly ash, blast furnace slag, aggregate, and characterized with higher porosity than the binder ma­
red mud, steel slag and tailings [1–3]. Thus, geopolymer concrete has trix. Therefore, the ITZ with porous microstructure largely determines
the potential to reduce CO2 emissions significantly [4]. McLellan and the mechanical properties and durability of the whole concrete.
Turner presented a study of the life-cycle costs and carbon impacts of Currently, most of the ITZ studies focus on cement concrete, and just a
OPC and geopolymers, and found that geopolymer concrete shows the little on geopolymer concrete in comparison. The experimental studies
potential to reduce CO2 emissions by 14 % − 97 % at a financial cost that indicated that the geopolymer concrete has denser and more homoge­
is 7 % − 39 % lower than OPC [5,6]. Thus, geopolymer concrete has a neous ITZ than cement concrete [10,11]. It was also found that

* Corresponding author.
E-mail addresses: tzhongnan@chd.edu.cn (Z. Tian), z_zhengqi@126.com (Z. Zhang).

https://doi.org/10.1016/j.conbuildmat.2023.131404
Received 27 March 2022; Received in revised form 14 February 2023; Accepted 11 April 2023
Available online 29 April 2023
0950-0618/© 2023 Published by Elsevier Ltd.
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

aggregate types (calcareous vs siliceous) have a significant effect on the interfacial moisture were elucidated and compared. Finally, mechanical
microstructure and properties of ITZ [12–14]. Some mineral constituent behaviors of the geopolymer-aggregate interface with the participation
in aggregate have strong sensitivity to moisture, leading to great dif­ of interfacial moisture are investigated using peeling and shearing
ferences in moisture absorption of aggregate. When the aggregate is simulation. This work could facilitate a better design and fabrication of
stored or mixed, the moisture is easy to accumulate on the aggregate geopolymer concrete in engineering, and thus achieving improved
surface, thus increasing the water-cement ratio of ITZ. Nevertheless, performances of geopolymer concrete.
different from the hydration reaction of cement clinker with the
participation of moisture, the reaction of geopolymer binder involves 2. Computational methodology
the dissolution of Al and Si in the alkali medium, transportation
(orientation) of dissolved species, followed by a polycondensation, 2.1. Construction details of molecular models
finally forming a 3D network of silico-aluminate structures [15]. The
geopolymerization reaction is mainly related to the regional SiO2/Na2O As the main hydrated molecular of geopolymer binder, N-A–S–H
ratio [16]. Previous research shows that the aggregate with high water gel is the complex porous material with multi-scale feature, and at the
absorption has an obvious influence on the ITZ performance of cement nanoscale it is composed of alternating calcium-silica-aluminate sheets,
concrete, resulting in a wide and weak ITZ, but this trend is not signif­ free metal cation and water molecules. Further studies on geopolymers
icant for geopolymer concrete. This is because the hydration gel prod­ at the molecular level have shown that the molecular structure of N-
ucts of geopolymerization gradually precipitate on the edges of ITZ, and A–S–H is very similar to the Na2Si2O5 glass [27–29]. Therefore, in this
then expanded toward the interface area with the curing age increasing work, the Na2Si2O5 glass model was used as original models of N-
[17]. However, during this process, a geopolymer-moisture-aggregate A–S–H gel, and the model construction and simulation are performed
three-phase interface system potentially forms in the ITZ region, in the Materials Studio platform. Firstly, the Na2Si2O5 glass model with
which may lead to the performance decline of the whole material. At anhydrous molecules was used as the basic model, and its Si atoms were
present, there are still challenges to fully understand the interaction randomly replaced with Al atoms via Perl scripts to achieve the N-
between geopolymer and aggregate with the participation of interfacial A–S–H model with Si/Al = 1 [26,30,31]. The replaced Si atoms in
moisture, and how the moisture on the aggregate surface affects the ITZ molecular structure were four coordinated silicon, so the substitution of
properties of geopolymer concrete. Despite all of these existing efforts, Si atoms by Al atoms will cause charge loss; aiming to ensure the charge
the interfacial behaviors of geopolymer-aggregate with the participation balance of the system, it was necessary to redistribute the Na+ to reach
of moisture remain elusive, because current experimental techniques (e. charge balance. Subsequently, the Grand Canonical Monte Carlo
g., X-ray diffraction (XRD), Scanning electron microscopy (SEM), Energy (GCMC) method was performed to adsorb water molecules into the
dispersive spectroscopy (EDS)) cannot provide insights into the inter­ cavities of molecular structure. After completing all the above proced­
facial interaction at molecular level. ures, the initial models of N-A–S–H model was established, as shown in
In such a scenario, molecular dynamics (MD) simulations can be a Fig. 1(a).
excellent alternative to provide reliable insights into the molecular However, aiming to ensure the newly established molecular models
structure and interaction mechanism of interfacial zone between two at the minimum energy state, it is essential to optimize the structural
phases [18–20]. Indeed, MD simulation nicely unravels and explains the model of the geopolymer in order to obtain a stable molecular structure.
material properties and interfacial characteristics of geopolymer, Three algorithms viz. conjugate gradient, steepest descent and Newton
cement, minerals and polymer products. Lu [21] and Luo et al. [22] Raphson method were used to optimize the molecular structure box of
investigated by MD simulation and found that the adhesion and strip­ N-A–S–H in sequence. Then the box was subject to Canonical ensemble
ping behaviors between asphalt and aggregate were more strongly (NVT) runs at 300 k for 100 ps and Isobaric-Isothermal ensemble (NPT)
influenced by the aggregate properties and moisture. Cui et al. [23] runs at 500 K for 100 ps with a time step of 0.5 fs. On this basis, a
simulated the interaction of asphalt-aggregate interface modified by quenching operation was conducted; the system temperature quickly
silane coupling agents at dry and wet conditions, and analyzed the in­ dropped from 500 K to 300 K and was kept for 100 ps [32,33]. In order
fluence of silane coupling agent on asphalt-aggregate interface from the to keep the box energy at the lower and more stable level, the molecular
aspects of contact angle, transition zone thickness, interaction energy dynamics simulation using Nosé-Hoover thermostat and Berendsen
and hydrogen bond. Mutisya et al. [24] used MD simulation to deter­ method of pressure control was carried out and equilibrated 100 ps with
mine the water dynamical properties in the interlayer pores, gel pores a time step of 0.5 fs under NPT and NVT ensemble.
and connected pores of calcium silicate hydrate (C–S–H) colloid Considering that common aggregates for concrete production in­
model. Hou et al. [25] investigated the bonding mechanism of the cludes granite, basalt, limestone, etc, the silica and calcium unit are the
interfacial region between concrete and steel reinforcement, and main chemical components of aggregates, so silica (SiO2) and calcium
explained the interfacial behaviors and bonding mechanism between carbonate (CaCO3) are often selected to build the representative
passive film of steel and C–S–H at the molecular level. Kai et al. [26] aggregate models in this work, as shown in Fig. 1 (b) and (c). The
focused on the interface between geopolymer and silica aggregate using detailed information of aggregate models is listed in Table 1. The
MD simulation, and found that Si/Al ratio had an noticeable effect on aggregate models were firstly cleaved to expose their common surfaces
the density of ITZ. Up to now, the MD simulation has been widely used to aiming to investigate their surface characteristics and interfacial
investigate the interfacial behavior and interaction mechanism in behavior. The relatively stable cleavage surface was selected as the
various composite systems, to observe the interaction forms of mole­ follow-up study, i.e. (0 0 1), (1 0 1) of SiO2 and (0 1 0), (1 0 4), (2 1 4) of
cules and atoms at the nanoscale. The probing to the subject, a better CaCO3 [34,35]. Moreover, because of the layered structure of N-
understanding and explanation of the interfacial interaction of A–S–H, the (0 0 1) surface was chosen and exposed for the following
geopolymer-aggregate with the participation of interfacial moisture can simulation.
be achieved, which are usually not approachable experimentally. The determination of forcefield is a vital procedure of MD simula­
In this work, MD simulation is employed to provide a molecular-level tion, and it varies depending on the sorts of established structure model.
understanding of the effect of moisture on interfacial behaviors of COMPASS is the first high-quality forcefield to integrate parameters of
geopolymer-aggregate interaction. First, full atomistic models adopted organic and inorganic materials, marking a technological break-through
for MD simulations were constructed using the N-A–S–H gel model and in forcefield methodology. In addition, COMPASS II 1.2 is a pivotal
the main chemical components of the aggregates, SiO2 and CaCO3. Then extension to the COMPASS forcefield, extending current the scope and
the wetting characteristics of aggregate surfaces, and interfacial char­ incorporate more compounds and chemicals [36]. As the simulation in
acteristics of the geopolymer-aggregate interfacial systems containing this work is a hybrid system with metal oxides and non-metal oxides, the

2
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 1. The structure model of (a) N-A–S–H, (b) SiO2 and (c) CaCO3 unit.

Fig. 2.
Table 1 In MD simulation, the contact state of water droplets and solid sur­
Detailed information of aggregate models used in MD simulation.
face can be effectively simulated to characterize the water infiltration
Mineral Composition Space Lattice Coordination characteristics of the surface. First, the cleavage surfaces of SiO2 and
type group parameters numbers
CaCO3 were cleaved with around 20 Å thickness larger than the cut-off
Silica SiO2 P3221 a = b = 4.909 Å c Si-O: 4 distance 15.5 Å (1 Å = 0.1 nm). Then a supercell aggregate surface (U ×
= 5.402 Å Si-Si: 4 V = 160 Å × 160 Å) was established with a vacuum layer of 50 Å;
α = β = 90.0◦ γ =
120.0◦
furthermore, a water droplet with a diameter of 30 Å was constructed
Calcite CaCO3 R-3C a = 4.990 Å b = Ca2+–CO2-
3: 6
with a density of 1.0 g/cm3 using the Amorphous Cell module. The water
4.990 Å c = C–O: 3 droplet was then subjected to a 100 ps NVT dynamic run at a temper­
17.061 Å ature of 300 K to make itself stable. Subsequently, the wetting models of
α = β = 90.0◦ γ =
the aggregate surface were set up by depositing the water droplet over
120.0◦
the center of aggregate surfaces, as illustrated in Fig. 3. The contact
Notes: 1 Å = 0.1 nm. angle was measured after the wetting models were equilibrated with a
500 ps NVT run to simulate the diffusion state of water molecules on the
COMPASS II force field has adequate potential parameters to ensure the aggregate surfaces.
accuracy and reliability of its application in geopolymer system, which
has been confirmed by many scholars [27,28,37–39]. Therefore, COM­ 2.2.2. Interfacial interaction simulation
PASS II forcefield was selected as the forcefield in this work. As a crucial assessment index, the interaction energy between geo­
polymer and aggregate is employed as an intermolecular combination
parameter to define the energy required to separate the interface system
2.2. Simulation model and methodology into two independent layers on the molecular scale [34,41]. Here,
geopolymer layer and aggregate layer, were established by repeating in
2.2.1. Wetting characteristics of aggregate surface the U and V directions (60 Å × 60 Å) based on the geopolymer and
Wettability of the aggregate surface is the ability of water to diffuse aggregate models, to produce the supercell surface. And the waterlayers
or keep up contact on the aggregate surface, and it is governed by wa­ with different thicknesses from 1 Å to 10 Å were constructed, then the
ter’s attraction energy to the aggregate surface. The wetting quality of sandwich models containing the geopolymer, waterlayer and aggregate
aggregate surface universally impacts the water absorption of the entire layer were combined by means of the Building layer tool to form the
aggregate; aggregates with high water absorption are more likely to geopolymer-water-aggregate interfacial system. Following that, a vac­
develop a powerless interfacial transition zone (ITZ). For geopolymer- uum layer of at least 75 Å was set up at the top of the interfacial system
aggregate system, the wettability of aggregate surface has a substan­
tial effect on the accumulation of water molecules on aggregate surfaces
[40]. In this simulation, the wettability of the aggregate surface is
characterized by the contact angle of the water droplet on its surface
after thermodynamic equilibrium. The contact angle is defined as the
angle formed between the water droplet and the surface when the water
droplet is on the aggregate surface. The schematic diagram is shown in

Fig. 2. Contact Angle Diagram. Fig. 3. Wetting model of aggregate surface.

3
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

to eliminate the impact of three-dimensional periodic boundary condi­


tions in the vertical direction. Subsequently, the established interfacial
models were first geometrically optimized for 5000 iterations, then
running NVT dynamics for 500 ps at a time step of 0.5 fs. Fig. 4 shows an
example of geopolymer-water-aggregate interfacial system based on N-
A–S–H, waterlayer and SiO2 layer.

2.2.3. Interfacial peeling simulation


It is well understood that the interface between geopolymer binder
and aggregate in concrete will bear complex stress, incurring the driving
force for the interface failure, whereas this will be worse in the presence
of interfacial moisture. The maximum normal stress in the experimental
evaluation, i.e. tensile strength, is primarily measured by pull-off test,
which can easily evaluate the tensile strength and failure mode of the
geopolymer-water-aggregate interfacial system. Here, at the atomistic
scale, the peeling simulation test was carried out by employing MD
simulations to investigate the normal separation phenomenon of
geopolymer-aggregate interface with the presence of interfacial mois­
ture. Firstly, based on the interfacial model established above, the
aggregate layer was fixed, and the interfacial waterlayer was kept free.
The top geopolymer layer was then moved upward, away from the
waterlayer and aggregate layer with a constant normal force, as shown
in Fig. 5. This process was equilibrated for 20,000 steps under NVT
ensemble with the time step of 0.1 fs, and the tensile stress and atom
coordinates were recorded every ten steps, plotting the stress-separation
curve throughout the process. This simulation was accomplished with
the help of a self-developed Perl script document in the Materials Studio
platform.

2.2.4. Interfacial shearing simulation


Fig. 5. Schematic of tensile simulation based on geopolymer/steel-slag system.
Shearing failure is the relative sliding of the geopolymer-water-
aggregate under the driving of maximum tangential stress; addition­
ally, this risk of shear failure will be exacerbated with the presence of reversely at a shearing rate of 0.02 Å/ps. This process was equilibrated
interfacial moisture. Luckily, the shearing motion of the interfacial for 500 ps under NVT ensemble with the time step of 0.5 fs, and the
system can also be simulated in MD technology. Due to the inter- stress and atom coordinates were recorded every 5000 steps. The
molecular attraction of the “sandwich” system, the interfacial system schematic diagram of shearing model is depicted in Fig. 6.
will bear a downward pressure (as shown in Fig. 6), which is related to During this shearing process, the vertical pressure P and the shear
the interaction energy of the interfacial system. In this simulation, the stress were measured; then the shear coefficient μ could be calculated
geopolymer layer is the TOPWALL layer, the BOTTOMWALL layer is the using Eq.(5), similar to the friction coefficient; it is used to assess the
aggregate layer, and the waterlayers with different thicknesses at the shearing failure state of the interfacial system with varying moisture
interface is defined as the FLUID layer. And the external forces parallel content..
to the horizontal direction drew the geopolymer and aggregate to move

Fig. 4. The molecular models for (a) geopolymer layer; (b) waterlayer and (c) aggregate layer; and (d) geopolymer-water-aggregate interfacial model.

4
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 7 shows the contact angles of water droplets on the different sur­
faces of SiO2 and CaCO3. The contact angles of water droplets on the
surfaces of SiO2 all exceed 90◦ , indicating that these two crystalline
surfaces of SiO2 are hydrophobic and water is difficult to spread and
diffuse on their surfaces. The contact angles on (0 0 1), (1 0 4) and (2 1 4)
of CaCO3 are 13.8◦ , 20.8◦ and 11.9◦ respectively, indicating that these
surfaces of CaCO3 have strong hydrophilicity, allowing water to easily
spread on the CaCO3 surface, forming a thin water film. Fig. 8 shows the
typical equilibrium snapshots of water wetting process on the SiO2
(1 0 1) and CaCO3 (2 1 4) surfaces. Caused by the hydrophilicity of the
aggregate minerals, water films would be easily formed around the ag­
gregates of freshly mixed geopolymer concrete, thus causing the
powerless ITZ. Intermolecularly, the free Ca2+ and CO2– 3 on the CaCO3
surfaces are more likely to cause electrostatic adsorption, showing that
their surfaces possess a strong interaction with water. On the SiO2 sur­
faces, the bond dipole moment, which caused by the uneven charge
distribution of Si-O polar covalent bonds, provides weaker electrostatic
action, but due to the lack of free ions, the SiO2 surfaces produce
significantly less electrostatic adsorption of water than CaCO3. The
above results reveal that the two main chemical components of aggre­
gate, SiO2 and CaCO3, have different wetting qualities to water, which
Fig. 6. Schematic diagram of shearing model of geopolymer-water-
would have an important influence on the interfacial behavior of
aggregate system.
moisture. As a result, the SiO2 (1 0 1) and CaCO3 (2 1 4) surfaces with
lower contact angles, which are more sensitive to water, were selected
μ=
Sx for subsequent simulation and analysis.
P
Where Sx is the shear stress in the moving direction; P is the vertical
3.2. Effect of moisture on interactive behaviors of geopolymer-aggregate
pressure in z-direction; μ is the shear coefficient. Considering the
interface
different lattice sizes of geopolymer and aggregate, the average P and Sx
of the last ten frames of the trajectory were used to eliminate the size
3.2.1. Interfacial interaction energy of geopolymer-aggregate interface
effect and obtain reliable results.
As an intuitive index to quantitatively analyze the adhesion ability of
the two-phase interface, interaction energy is calculated from the non-
3. Result and discussion
bonding energy of each part in the interfacial system because the non-
bonding interaction is considered the most significant at the interface
3.1. Moisture wettability of aggregate surface
[42]. The interaction energy between aggregate and geopolymer,
namely adhesion energy, is defined by Eq.(2). Similarly, the interaction
The contact angle is a common parameter for assessing the moisture
energies of geopolymer-water and aggregate-water are calculated by Eq.
wettability of the aggregate surface. A contact angle larger than 90◦
(3) and Eq.(4), respectively.
indicates that this surface can hardly be wetted by water, which is
regarded as a hydrophobic surface. Conversely, when the contact angle Egeo + Eagg − Egeo + agg
Wgeo - agg = (2)
is measured less than 90◦ , the surface is considered hydrophilic. And the A
smaller the contact angle, the easier it is for water to soak the surface.
Egeo + Ewater − Egeo + water
Wgeo - water = (3)
A

Eagg + Ewater − Eagg + water


Wagg - water = (4)
A

where Wgeo-agg is the interaction energy between geopolymer and


aggregate, the Wgeo-water and Wagg-water are the interaction energy of
geopolymer-water and aggregate-water, respectively. Egeo, Eagg and
Ewater are successively the non-bonding energies of geopolymer, aggre­
gate layer and waterlayer exited lonely. Egeo+agg, Egeo+water and Eagg+water
are sequentially systematic energies of geopolymer-aggregate, geo­
polymer-water and aggregate-water systems. A is the interface area.
Fig. 9 depicts the interaction energy of geopolymer-aggregate sys­
tems with different waterlayer thickness. The interaction energy of
geopolymer-aggregate interface is mainly electrostatic energy, because
there are abundant ion clusters at the interface, which is also confirmed
by previous studies [10,26,43]. In Fig. 9, the Wgeo-agg of geopolymer-SiO2
system is smaller than that of CaCO3 system, caused by the stronger
electrostatic adsorption of free ions on the CaCO3 surface. With the in­
crease of waterlayer thickness, Wgeo-agg of the geopolymer-SiO2 and
geopolymer-CaCO3 systems first increases and subsequently falls,
reaching the highest when the water layer thickness is 0.1 nm. This
Fig. 7. Contact angles of water droplets on the different surfaces of SiO2 indicates that, compared with the interface dry condition, an appro­
and CaCO3. priate quantity of interfacial water is conducive to the interfacial

5
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 8. Water wetting process on the aggregate surfaces.

90 Wgeo-agg 90 Wgeo-agg
Wgeo-water Wgeo-water
80 80 Wagg-water
Wagg-water
Interaction energy (J/m2)

Interaction energy (J/m2)


70 70

60 60

50 50

40 40

30 30

20 20

10 10

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Waterlayer thickness (nm) Waterlayer thickness (nm)

Fig. 9. Interaction energy of geopolymer-aggregate systems with differernt waterlayer thicknesses.

interaction between geopolymer and aggregate. Affected by the increase interfacial system in the dry condition; additionally, Ww(geo-agg) is that of
of the waterlayer thickness, Wgeo-water and Wagg-water both grow dramat­ the system with the interfacial moisture. Importantly, nER is the ratio of
ically. However, in the CaCO3 system, the value of Wagg-water is much the adhesion work between geopolymer binder and aggregate to the
larger than that of Wgeo-water, but in the SiO2 system, this trend is debonding work under the presence of water in the interface area. A
inverted, which is mostly caused by the difference in the sensitivity of positive value of nER indicates that the interface of geopolymer-water-
the aggregate surface to moisture. For the interface system containing aggregate system is effectively bonded; the higher the value, the stron­
moisture, it is agreed that moisture content significantly affects inter­ ger the bonding effect. Conversely, when nER is negative, the debonding
facial characteristics, so the energy ratio (nER) was proposed to evaluate work of the interface exceeds the adhesion work, which indicates that
the effect of saturation on interface affinity, as shown in Eq.(5) the interface is at risk of failure caused by the moisture action.
[22,23,34,44]. Fig. 10 shows the energy ratio (nER) of the geopolymer-aggregate
with different waterlayer thicknesses. It is found that the nER de­
Wbonding water Wd(geo− agg) − Wagg− water − Wgeo− water
nER = = creases from positive to negative as the waterlayer thickness increases.
Wdebonding water Wagg− water + Wgeo− water + Ww(geo− agg) − Wd(geo−
For CaCO3 system, when the interface system has a waterlayer thickness
agg)

(5) of 0.1 nm at the interface, nER has the highest positive value; subse­
quently, with the increase of waterlayer thickness, nER decreases
where Wbonding_water and Wdebonding_water are the bonding energy and
sharply, and approaches zero at the water layer of 0.4 nm. In geo­
debonding energy between geopolymer and aggregate in wet condition.
polymer-water-SiO2 system, the trend is similar; the nER is positive and
The definition of Wdebonding_water is similar to that of references [22].
the highest at 0.1 nm, and it approaches to zero at waterlayer thickness
Differently, Wbonding_water is defined as the difference between the inter­
of 0.6 nm. When nER is less than 0, the risk of interface system debonding
action energy of geopolymer-aggregate under a dry condition and the
is high. When the interfacial moisture thickness is 0.1 nm, geopolymer-
interaction energy of water-geopolymer and water-aggregate in the
CaCO3 shows a more stable interface than geopolymer-SiO2. However,
presence of water; because the later two have the weakening effect of
the decreasing rate of nER of CaCO3 system is significantly higher than
geopolymer-aggregate bonding. Wd(geo-agg) is the Wgeo-agg of the

6
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

shown in Fig. 11, whether there is interfacial water or not, the interface
area between geopolymer and aggregate is always a weak area. How­
ever, caused by the different thicknesses of interfacial waterlayer and
types of aggregates, there are obvious differences in density distribution
of the systems. When the thickness of the interfacial water layer is
0.1–0.2 nm, the interfacial density of geopolymer-aggregate is higher
than that of the system without interfacial water, which shows that a
small quantity of interfacial water can improve the density of ITZ.
Nevertheless, once the waterlayer thickness exceeds 0.3 nm, the inter­
facial density will be lower and lower, indicating that the aggregate
surface with higher moisture content would weaken the adhesion be­
tween geopolymer and aggregate. Furthermore, in the geopolymer-
CaCO3 system, the density of the geopolymer binder is at a low level
compared with the SiO2 system, especially when the waterlayer thick­
ness is 0 nm. It implies that the CaCO3 with high hydrophilicity may
attract more water molecules weakly adsorbed in the geopolymer binder
to gather on its surface, reducing the density of the geopolymer matrix
and forming pores.
Affected by the hydrophilicity of the aggregate, the water molecules
in the geopolymer will migrate to the aggregate interface, reducing the
density and strength of the geopolymer matrix. In order to analyze the
Fig. 10. Energy ratio (nER) of the geopolymer-aggregate with differernt
waterlayer thicknesses.
migration behavior of water molecules in geopolymers on aggregate
surfaces, mean squared displacement (MSD) was used to calculate the
self-diffusion coefficient of water molecules. An atom’s MSD can be
that of SiO2, causing a higher risk of interface failure. This may be
defined as an average of the deviation positions with regard to their
related to the high sensitivity of CaCO3 surface to water than SiO2 sur­
reference positions throughout time translation in equilibrium, calcu­
face, making the water molecules at the interface or in the geopolymer
lated as a function of time by the Eq.(6),
migrate and accumulate on the aggregate surface.
〈 〉 1 ∑ N

3.2.2. Behavior of moisture at geopolymer-aggregate interface MSD(t) = |r(t) − r0 |2 = |ri (t) − ri (0)|2 (6)
N i=1
At the geopolymer-aggregate interface, the thinner water layer (0.4
nm) on the CaCO3 surface may lead to a high risk of interfacial inter­ where N is the atom number; ri(0) is the reference position of the i-th
action failure, as confirmed in Fig. 10 above. This may be accounted for atom, and ri(t) is the i-th atom position of at time t. The self-diffusion
that the hydrophilicity of the aggregate surface causes more water coefficient D is calculated using Eq.(7) based on the resulting MSD
molecules to accumulate on the aggregate surface. The accumulation of curve,
water molecules at the interface will cause the density change of the
system. Here, the density distributions of different systems were coun­ 1
D = lim
MSD(t) 1 MSD(t1 ) − MSD(t0 )
≈ (7)
ted, and the density distribution of geopolymer-SiO2 and geopolymer- 6 t→∞ t 6 t1 − t0
CaCO3 systems with waterlayer thickness of 0 nm ~ 0.4 nm is shown in Where t1 and t0 represent the final and initial computing time
Fig. 11. According to the previous established models, the density of N- respectively.
A–S–H gel is 2.35 g/cm3, the density of SiO2 is 2.2 g/cm3, the density Fig. 12 summarizes the MSD curves of water molecules in geo­
of CaCO3 is 2.9 g/cm3, and the density of waterlayer is 1.0 g/cm3. As polymer binder with different thicknesses of interfacial waterlayer over

Fig. 11. Density distribution of geopolymer-SiO2 and geopolymer-CaCO3 systems with waterlayer thicknesses of 0 ~ 0.4 nm.

7
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 12. MSD curves of water molecules in geopolymer binder with different thicknesses of interfacial waterlayer.

a time scale of 500 ps. All MSD curves show that, affected by the Moreover, the self-diffusion coefficients of geopolymer-SiO2 systems are
interfacial waterlayer with different thicknesses, water molecules in much lower than that of geopolymer-CaCO3 systems. For example, the
geopolymer matrix diffuse to the interface at different rates. The less the self-diffusion coefficient of geopolymer-SiO2 system with 0.1 nm
interfacial water, the faster the diffusion rate of water molecules in the waterlayer is 2.09 × 10-13 m2/s, only half of that (4.27 × 10-13 m2/s) of
geopolymer to the interface. when the interface is dry, the diffusion rate geopolymer-CaCO3 system with the same waterlayer. This phenomenon
is fastest, which indicates that the hydrophilic aggregate surface will is caused by the hydrophilicity of the aggregate; CaCO3 surface has
indeed attract the water molecules to gather to the aggregate surface. strong hydrophilicity, while SiO2 surface is hydrophobic. Under the
The water molecules in the geopolymer are weakly bonded with other electrostatic attraction of the aggregate surface, the interfacial water­
atoms by H-bond confinement. When there are not enough water mol­ layer and the weakly bound water molecules in geopolymer binder at
ecules on the aggregate surface, the strong electrostatic action on the interface would both diffuse to the aggregate surface. The diffusion of
aggregate surface will break the constraints of weakly bound water interfacial water leads to the non-uniform porous microstructure in the
molecules in the geopolymer and cause their diffusion. interfacial transition zone between geopolymer matrix and aggregate,
Fig. 13 depicts the self-diffusion coefficient of water molecules in which has been investigated by Steinerova [45] and Huang [46].
geopolymer binder with different thicknesses of interfacial waterlayer. It Nevertheless, it is worth noting that the presence of water molecules at
is clearly evident that the self-diffusion coefficient is significantly the geopolymer-aggregate interface is not always a worrying phenom­
affected by the thickness of interfacial water layer and the type of enon, because water molecules could be bonded via H-bonds with ox­
aggregate. The thicker the water layer on the aggregate surface, the ygen atoms on the surface of geopolymer or aggregate, which is
lower the self-diffusion coefficient of water molecules to the interface. conducive to the interface interaction between geopolymer and
aggregate.

3.2.3. H-bond profile at geopolymer-aggregate interface


SiO2 A special kind of non-bonding interaction. H-bond interaction can
5
Diffusion coefficient (10-13 m2/s)

CaCO3 form highly oriented network clusters through H-bonds, connecting the
hydrogen atoms in water molecules with the oxygen atoms on the solid
surface. Compared with the dry interface, the presence of interfacial
4
water will produce H-bond interaction at the interface, which is helpful
to form a stable interface interaction. The larger quantity of H-bonds at
the interface implies the stronger H-bond interaction of two phases.
3
Water molecules at the interface between geopolymer and aggregate are
conducive to interfacial H-bond interaction. Generally speaking, the
more water molecules, the stronger the H-bond interaction. To further
2 investigate the H-bond interaction at the geopolymer/steel-slag inter­
facial system, the H-bond profile in z-direction was conducted and
counted with the criterion (acceptor distance ≤ 3.0 Å and angle cutoff ≤
1 20◦ ) by writing Perl script documents.
Fig. 14 describes the H-bond profile at the interface between geo­
polymer and aggregate with 0.1–0.6 nm waterlayer. It is found that the
0 thicker the waterlayer at the interface, the stronger the intensity of H-
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 bond and the wider the distribution range. However, the H-bond dis­
Waterlayer thickness (nm) tribution peaks on the surfaces of SiO2 and CaCO3 are significantly
different; in geopolymer-SiO2 system, the intensity of H-bond distribu­
Fig. 13. Self-diffusion coefficient of water molecules in geopolymer binder
tion on the SiO2 surface is higher, indicating that the H-bond interaction
with interfacial waterlayer.

8
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 14. H-bond distribution at the interface between geopolymer and aggregate.

between geopolymer and SiO2 is higher than that of geopolymer-CaCO3 then the H-bond lengths distribution was counted. Fig. 15 shows the H-
system. Moreover, there are H-bond distribution peaks near the surface bond length distribution on the geopolymer and aggregate surfaces. As is
of SiO2 and geopolymer, indicating that H-bond exists widely on their known to all, the shorter the H-bond length, the greater the H-bond
surfaces, and yet in the geopolymer-CaCO3 system. H-bonds are mainly energy; a high proportion of small-length H-bonds implies the stronger
distributed near the surface of geopolymer and rarely near CaCO3. This H-bond interactions at the interface. It can be found in Fig. 15 that with
is because there are a large quantity of exposed O atoms on the SiO2 the increase of waterlayer thickness, the proportion of H-bond with 1.0
(1 0 1) surface, which have the potential to produce H-bonds with H Å ~ 2.0 Å length increases slowly, meaning that the H-bonding energy is
atoms in water molecules, while the CaCO3 (2 1 4) surface is mainly growing. The larger amount of interfacial water brings the water mol­
composed of Ca and C atoms. Furthermore, as shown in Fig. 14, the ecules closer to the geopolymer and aggregate surfaces, thus increasing
width of the H-bond distribution peak near the geopolymer surface is the proportion of high-energy H-bonds. However, this growth is
wider than that near the aggregate surfaces. Because a rich Na+ layer is restricted because of the adsorption sites of the geopolymer and aggre­
located on the geopolymer surfaces to balance the negative charge gate surfaces are limited to allow the adsorption of water molecules. In
generated by the aluminosilicates; this Na+ layer keeps a long distance addition, the aggregate type has a significant effect on the bond length
between the interfacial water and the O atom of aluminosilicates, distribution of H-bond at the interface; the geopolymer-SiO2 systems are
resulting in a wide H-bond distribution peak on its surface. It should be more likely to form H-bonds with shorter bond lengths at their interface
noted that, the H-bonds on the geopolymer and aggregate surfaces are than the geopolymer-CaCO3 systems. It indicates that when the geo­
helpful to enhance their interface interaction, but with the amount of polymer is in wetting contact with the aggregate, the geopolymer-SiO2
interfacial water increasing, the H-bonds at a certain distance away from interfacial system will produce stronger H-bond interaction than the
the geopolymer and aggregate surfaces are difficult to strengthen the geopolymer-CaCO3 system.
interface interaction.
To further discuss the H-bond interaction on the geopolymer and 3.2.4. Diffusion behavior of metal cations at geopolymer-aggregate
aggregate surfaces, Perl script document was written and the H-bonds in interface
the range of 1.0 Å on their surfaces were emphatically highlighted, and As discussed above, a rich Na+ layer is located on the geopolymer
surfaces to balance the negative charge generated by the aluminosili­
cates. Driven by electrostatic force, these metal cations will inevitably
diffuse to the interfacial water and aggregate surface, thus resulting in a
series of material characteristics, such as ion exchange capacity, alkali
metal immobilization and leaching [47,48]. Here, mean squared
displacement (MSD) was used to calculate the self-diffusion coefficient
of Na+. The MSD curves and self-diffusion coefficient of Na+ are sum­
marized in Fig. 16 and Fig. 17. Similar to the diffusion behavior of water
molecules, the self-diffusion coefficient of Na+ in geopolymer- CaCO3
system is lower than that in SiO2 system, which is attributed to Na+
diffusion driven by the charge dipole generated in diffused water mol­
ecules. But differently, when the interface between geopolymer and
aggregate is dry, Na+ diffuses the slowest, because Na+ is repulsed by
the positively charged aggregate surface. Then, Na+ diffuses fastest
when there is a 0.1 nm water layer at the interface, and the self-diffusion
coefficient of Na+ is three times that of the interface without water,
which indicates that a small quantity of water will promote the diffusion
of Na+ to the interface compared with no water there. However, with the
increase of waterlayer thickness, the diffusion of Na+ seems to be
limited, and the self-diffusion coefficient gradually decreases, as shown
Fig. 15. H-bonds length distribution on the geopolymer and aggre­
in Fig. 17. There may be two main reasons to explain this phenomenon.
gate surfaces.

9
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 16. MSD curves of Na+ in geopolymer binder with different thicknesses of interfacial waterlayer.

3.5
SiO2
Diffusion coefficient (10-13 m2/s)

3.0 CaCO3

2.5

2.0

1.5

1.0

0.5

0.0 Fig. 18. Interfacial system with (a) 0.2 nm and (b) 0.6 nm waterlayer at
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 equilibrium of 500 ps.
Waterlayer thickness (nm)
3.2.5. Coordination of metal cations with water at geopolymer-aggregate
Fig. 17. Self-diffusion coefficient of Na+ in geopolymer binder with interfa­
interface
cial waterlayer.
The metal cations and water molecules can form ion/dipole-bond
interaction, thus forming ionic clusters (hydrated ions) to stabilize the
The first is that the aluminosilicate with negative charge in the geo­
interfacial interaction between the geopolymer and aggregate, as shown
polymer exhibit strong spatial correlation with Na+, limiting the diffu­
in Fig. 19. For the aggregate surface with different water content, the
sion of Na+ to the water layer and aggregate surface [49]. Fig. 18 is the
hydration state of hydration ions at the interface differs among the
interfacial system with 0.2 nm and 0.6 nm waterlayer at 500 ps equi­
quantity of interfacial water molecules, resulting in the coordination
librium, showing that a rich Na+ layer is constrained on the geopolymer
between metal cations and different numbers of water molecules. The
surface, and only a small fraction of Na+ can diffuse to the interface.
interaction between metal cations and water molecules significantly
Secondly, the movement of Na+ slows down because of the collision and
affects the properties of interfacial water such as viscosity, ionic
obstruction with the neighboring water molecules; the Na+ is restricted
mobility, diffusion, etc., and the coordination number around the metal
as the center atoms in the small “cages” constructed by surrounding
cations accounts for the interaction between the ions and the water
water molecules, which is also discussed by Hou [47]. Metal ions such as
molecules [50]. In geopolymer-aggregate system, Na+ shows strong
Na+ dissolved in water are subjected to hydrolysis, and metal ions are
correlation with surrounding oxygen atoms of water molecules in the
susceptible to forming hydrated ions ([M(H2 O)z ]n+ )by coordinating short-range and ionic bonds, which has been confirmed by previous
water molecules. When water molecules are sufficient, the coordinated work using the radial distribution functions (RDF) [51].
water molecules will restrict the continued diffusion of Na+. The coordination number represents the number of the first hydra­
tion shell of Na+ ions in the RDF curve. To quantitatively investigate the
coordination of water molecules around Na+, the coordination number
of ions is calculated according to Eq.(8),

10
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 19. Ion clusters formed by Na+ and water molecules at the geopolymer-aggregate interface.

∫ r
between geopolymer and aggregate in the form of H-bonds, coordina­
nij (r) = 4πρj r2 gij (r)dr (8)
0 tion bonds and ion bonds. However, with the increase of interfacial
waterlayer thickness, this coordination interaction will inevitably
where nij(r) is the coordination number, gij(r) is the radial distribution weaken, and more water will weaken the electrostatic interaction be­
function, and ρj is the density. A self-written Perl script was used to tween geopolymer and aggregate, resulting in a tendency for mechanical
calculate the coordination numbers of Na+ at the interface with different failure of geopolymer-moisture-aggregate system under the action of
waterlayer thicknesses, and the results are shown in Fig. 20. Since the external force.
free Na+ ions on the geopolymer surface are with a limited quantity, the
number of water molecules at the interface will have a substantial in­
3.3. Effect of moisture on mechanical behaviors of geopolymer-aggregate
fluence on the coordination number of Na+. As shown in Fig. 7, as the
interface
quantity of water molecules at the interface increases, limited free Na+
ions from the geopolymer surface coordinate with increasing water
3.3.1. Peeling simulation of geopolymer-aggregate interface
molecules until they are saturated. Along with the increasing number of
The peeling separations of geopolymer-aggregate interface were
interfacial water molecules, the excess water molecules which cannot be
simulated to investigate the effects of interfacial moisture on the me­
coordinated or adsorbed, may play an adverse effect on the interaction
chanical responses. The typical result of stress-separation relationship of
between geopolymer and aggregate. Additionally, at the same interfa­
geopolymer-aggregate system in tension-induced interface failure pro­
cial water thickness, the Na+ coordination number of geopolymer-SiO2
cess is shown in Fig. 21. It is found in Fig. 21 that the stress experiences
interface is higher than that of geopolymer-CaCO3 interface, which is
an initial linear increase up to a peak value at approximately 0.02 nm
that some water molecules are adsorbed on the hydrophilic CaCO3
displacement; after that first peak, the stress drops sharply. Then in the
surface and coordinated with the spilled Ca2+. The hydrated ion clusters
range of 0.05 nm ~ 0.08 nm displacement, the stress increases slightly
can be used as intermediates to stabilize the interfacial interaction
and then drops subsequently, forming the second stress peak; and finally
the interfacial stress decreases to a small value, indicating a final peeling
8 separation between geopolymer and aggregate. The first peak is located
SiO2 in the range of 0 nm ~ 0.05 nm with a high and sharp shape, manifesting
7 a strong and transient force; at this stage, the electrostatic interaction
CaCO3
between geopolymer and aggregate is mainly overcome steadily,
6 accompanied by the violent vibration of water molecules and Na+. In the
peeling separation range of 0.05 nm ~ 0.08 nm, the stress produces a
Coordination Number

5 lower peak, indicating that the system is overcoming a weak interaction


force. And the H-bonds and coordination bonds at the interface exhibits
4 weak force and long action distance, thus they may be overcome at this
stage. In Fig. 21, it is also found that some Na+ ions, escaping from the
3 geopolymer surface, are attracted by the coordinated water molecules
and aggregate surface, which reveals the presence of cation -water co­
2 ordination interactions at the geopolymer-aggregate interface.
Fig. 22 depicts the tensile strength of the geopolymer-aggregate
1 interface with different waterlayer thicknesses throughout peeling
simulation. Compared with the dry interface, the tensile strength be­
0 tween geopolymer and aggregate first increases, reaches the peak when
the waterlayer thickness is 0.1 nm ~ 0.2 nm, and then decreases rapidly
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 with the increase of the waterlayer thickness. An appropriate amount of
water molecules at the interface can strengthen the tensile strength of
Waterlayer thickness (nm)
geopolymer-aggregate interface, which is consistent with the results of
Fig. 20. Coordination number of the Na+ at the geopolymer- interaction energy, indicating that the water molecules at the interface
aggregate interface. will participate in the interface behavior of geopolymer and aggregate,

11
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 21. Stress-separation response in peeling simulation with snapshots in the separation process.

CaCO3 system under the same moisture content. SiO2 surface can
SiO2 contain more water molecules, and ensure its tensile strength with
1600
geopolymer binder not lower than dry aggregate, that is because of the
CaCO3
wetting characteristics of aggregate surface. As previously discussed, the
1400 strong hydrophilicity of CaCO3 surface drives the water molecules in the
Adhesion strength (MPa)

geopolymer to gather significantly in the interface interaction area,


1200 resulting in the decrease of the density of the interaction area. And as
confirmed by the experimental work of Tasong [12,52] and He et al.
[53], the limestone aggregate (CaCO3-based aggregate) produced a
1000
porous ITZ because of its high water absorption, and the ITZ of limestone
aggregates deform with more tendency than that with basalt and granite
800 aggregate (SiO2-based aggregate) using digital image correlation (DIC)
and nanoindentation tests [54]. Therefore, comprehended from the MD
600 simulation, the tensile strength between geopolymer and aggregate
containing more SiO2 mineral is less affected by interfacial water than
that of aggregate containing more CaCO3 mineral.
400
3.3.2. Shearing simulation of geopolymer-aggregate interface
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Shear failure is the failure of geopolymer and aggregate in geo­
Waterlayer thickness (nm) polymer concrete under horizontal force. Fig. 23 shows the shearing
simulation process of geopolymer-aggregate interface. As the geo­
Fig. 22. Tensile strength of the geopolymer-geopolymer interface with
different waterlayer thicknesses. polymer and aggregate layer move horizontally in the opposite direc­
tion, the interfacial waterlayer is stretched, accompanied by the
vibration, collision and friction of water molecules and Na+ ions. In this
strengthening electrostatic interaction and forming H-bonds and hy­
simulation, the shear coefficient μ is calculated as an evaluation index,
drated ion clusters. Furthermore, the continuous increase of interfacial
which is used to reflect the horizontal shear stress under the unit
waterlayer thickness expands the spacing between geopolymer and
constraint press. The larger the shear coefficient μ, the greater the shear
aggregate; a long action distance between geopolymer and aggregate
stress required for the shear movement of the interfacial system,
results in a weak electrostatic interaction force, which reduces the
implying that the geopolymer-aggregate interface bears a lower risk of
electrostatic interaction rapidly. Although H-bond and water-ion coor­
shear failure. Fig. 24 shows the shear coefficient results of different
dination interaction increase slightly with the increase of waterlayer
geopolymer-aggregate interfaces with different waterlayer thicknesses.
thickness, H-bond and coordination bond, as weak non-bonding inter­
The shear coefficient of the geopolymer-aggregate interface increases
action, are difficult to have a crucial impact on the interaction of
and then decreases with the thickness of the interfacial water layer. In
geopolymer-aggregate; electrostatic interaction is still the most
this simulation, there is a layer of water molecules, free ions and hy­
essential.
drated ions on the interface between geopolymer and aggregate, which
It is also found in Fig. 22 that the type of aggregate plays a significant
will hinder the horizontal movement of geopolymer-aggregate interface,
role in interfacial tensile strength of geopolymer-aggregate. When the
so as to stabilize the interface. However, with the increasing number of
interface is dry, the geopolymer binder is stronger bonded to the CaCO3
interfacial water molecules, more water molecules can not be adsorbed
surface than SiO2 surface, due to the strong coulomb potential of the
or coordinated, and herein play a lubricating role in shearing movement
CaCO3 surface. Moreover, when the interface is wet, the geopolymer-
instead, resulting in a decrease in the shear resistance of the
SiO2 system possesses stronger tensile strength than that of geopolymer-
geopolymer-aggregate interface.

12
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Fig. 23. Shearing simulation process: (a) initial state; (b) intermadiate state; (c) final state.

4. Conclusion
12
SiO2 (1) As the main chemical components of the aggregates, SiO2 and
10
CaCO3 CaCO3 exhibit very different surface wetting characteristics; the SiO2
surface has a contact angle of more than 90◦ and is hydrophobic, while
the CaCO3 surface has a contact angle of about 10◦ ~ 20◦ and exhibits
Shear coefficient

8 hydrophilic characteristics.
(2) Compared with the dry interface condition, an appropriate
quantity of interfacial water is conducive to the interfacial interaction of
6
geopolymer-aggregate system. With the continuous increase of interfa­
cial waterlayer thickness, the adhesion work of geopolymer-aggregate
4 decreases and the debonding work increases, forming a weak interface
area with lower density, and the aggregate with strong hydrophilicity
will aggravate this trend.
2
(3) Attracted by the aggregate surface, weakly bonded water mole­
cules and Na + ions in the geopolymer at interface will diffuse to the
0 interface area. Compared with the dry aggregate surface, the interfacial
water will hinder the diffusion of water molecules in geopolymer to the
interface, while Na+ is just the opposite. The diffusion of water mole­
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
cules and Na + ions is limited with the increase of the amount of water at
Waterlayer thickness (nm) the interface. The interfacial water molecules can form H-bond in­
Fig. 24. Shear coefficient of geopolymer-aggregate interface with different teractions, also coordinate with metal cations to compose hydrated ion
waterlayer thicknesses. clusters, reinforcing the interfacial interaction between the geopolymer
and the aggregate. The more interfacial moisture, the stronger the H-
In addition, the type of aggregate has a significant effect on the shear bond interaction and coordination interaction, but this growth is
coefficient of the geopolymer-aggregate interface. Shear coefficient μ of restricted.
the geopolymer-SiO2 interface peaks at the interfacial waterlayer of 0.2 (4) The hydrophilicity of aggregate surface will significantly affect
nm thickness, while the shear coefficient μ of the CaCO3 system con­ the interfacial behavior of water at the geopolymer-aggregate interface.
taining a 0.3 nm waterlayer is the largest, which is also greater than the Compared to the hydrophobic SiO2 surface, the strongly hydrophilic
peak shear coefficient of the geopolymer-SiO2 interface. This indicates CaCO3 surface exhibits weaker adhesion work and stronger debonding
that the geopolymer-CaCO3 system is more resistant to shear failure work with the geopolymer with the involvement of interfacial water. In
under interfacial wetting conditions, which may be related to the free the geopolymer-CaCO3 carbonate system, the self-diffusion coefficient
ions and surface configuration of the aggregate minerals. Firstly, the of water molecules and sodium ions is twice that of SiO2 system, which
interfacial region of the geopolymer-CaCO3 system possesses a large accelerates the accumulation of water molecules and Na+ in the inter­
number of hydrated ionic clusters with Na+ and Ca2+ ions as nuclei, face region. However, the accumulated water forms a weak H-bond
which can capture more water molecules to reduce the proportion of interaction with aggregates and geopolymers compared with
free water molecules. Secondly, the surface configuration of CaCO3 geopolymer-silica interface system, it is because that fewer oxygen atom
matrix is more disordered compared to that of SiO2, which increases the contacts on the CaCO3 surface.
roughness on the cell surface and increases the interfacial contact area (5) Mechanically, a proper amount of interfacial water at the inter­
and friction coefficient. It is also confirmed in the experimental tests that face between geopolymer and aggregate will improve the interfacial
the surface of limestone aggregate is rougher and the surface of basalt adhesive strength and shear strength, and reduce the risk of interface
and granite is smoother, and the ITZ region of smooth basalt or granite failure. Inspired by the MD simulation, with the participation of inter­
aggregate is more prone to shear damage under shear stress [55,56]. facial water, the geopolymer-SiO2 interfacial system possesses stronger
Therefore, comprehended from MD simulation, compared with aggre­ tensile strength than that of geopolymer-CaCO3, while the interfacial
gate containing more CaCO3 minerals, the interface between geo­ system between geopolymer and SiO2-based aggregate will bear greater
polymer and SiO2-based aggregate will bear greater risk of shear failure risk of shear failure when the interface is wet.
when the interface is wet.

13
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Declaration of Competing Interest [20] B. Pang, Y.T. Jia, S.D. Pang, Y.S. Zhang, H.J. Du, G.Q. Geng, H.M. Ni, J.J. Qian, H.
X. Qiao, G.J. Liu, The interpenetration polymer network in a cement paste-
waterborne epoxy system, Cem. Concr. Res. 139 (2021), 106236.
The authors declare that they have no known competing financial [21] Y. Lu, L.B. Wang, Atomistic modelling of moisture sensitivity: a damage
interests or personal relationships that could have appeared to influence mechanisms study of asphalt concrete interfaces, Road Materials and Pavement
the work reported in this paper. Design 18 (2017) 200–214.
[22] L. Luo, L.J. Chu, T.F. Fwa, Molecular dynamics analysis of moisture effect on
asphalt-aggregate adhesion considering anisotropic mineral surfaces, Appl. Surf.
Data availability Sci. 527 (2020), 146830.
[23] B. Cui, H. Wang, Molecular interaction of Asphalt-Aggregate interface modified by
silane coupling agents at dry and wet conditions, Appl. Surf. Sci. 572 (2022),
No data was used for the research described in the article. 151365.
[24] S.M. Mutisya, J.M. de Almeida, C.R. Miranda, Probing the dynamics of water over
Acknowledgement multiple pore scales in cement by atomistic simulations, Appl. Surf. Sci. 565
(2021), 150426.
[25] D.S. Hou, X.Q. Xu, M.H. Wang, Z. Chen, J.R. Zhang, B.Q. Dong, J.J. Miao, C.W. Liu,
The study is supported by Research Project by the National Natural Nanoscale insights on the interface between passive film of steel and cement
Science Foundation of China (52278428), Science and Technology hydrate: Diffusion, kinetics and mechanics, Appl. Surf. Sci. 514 (2020), 145898.
[26] M.-F. Kai, J.-G. Dai, Understanding geopolymer binder-aggregate interfacial
Project of Transportation Department of Hebei Province (2018-04); The characteristics at molecular level, Cem. Concr. Res. 149 (2021), 106582.
Fundamental Research Funds for the Central Universities, CHD [27] R. Wang, J. Wang, T. Dong, G. Ouyang, Structural and mechanical properties of
(300203211211). The authors gratefully acknowledge their financial geopolymers made of aluminosilicate powder with different SiO2/Al2O3 ratio:
Molecular dynamics simulation and microstructural experimental study, Constr.
support.
Build. Mater. 240 (2020), 117935.
[28] L.Y. Xu, Y. Alrefaei, Y.S. Wang, J.G. Dai, Recent advances in molecular dynamics
References simulation of the N-A-S-H geopolymer system: modeling, structural analysis, and
dynamics, Constr. Build. Mater. 276 (2021), 122196.
[1] Z. Tian, X. Liu, Z. Zhang, K. Zhang, X. Tang, Potential using of water-soluble [29] Z. Tian, Z. Zhang, H. Liu, J. Shi, X. Yang, Static and Dynamic Interfacial
polymer latex modified greener road geopolymeric grouts: Its preparation, Interactions of Geopolymer-Aggregate Considering Anisotropic Mineral Surfaces:’A
characterization and mechanism, Constr. Build. Mater. 273 (2021), 121757. Molecular Dynamics Study, China Journal of Highway and Transport (2022) 1–13.
[2] Z. Tian, Z. Zhang, N. Li, Y. Xu, H. Tang, Z. Gui, Composition and Performance [30] D.S. Hou, Z.J. Li, T.J. Zhao, Reactive force field simulation on polymerization and
Enhancement of Geopolymer Grouting Materials with Industrial Residue: a Review, hydrolytic reactions in calcium aluminate silicate hydrate (C-A-S-H) gel: structure,
Materials Review 34 (10A) (2020) 19034–19042. dynamics and mechanical properties, RSC Adv. 5 (1) (2015) 448–461.
[3] Z. Tian, Z. Zhang, K. Zhang, X. Tang, S. Huang, Statistical modeling and multi- [31] S. Chitsaz, A. Tarighat, Molecular dynamics simulation of N-A-S-H geopolymer
objective optimization of road geopolymer grouting material via RSM and MOPSO, macro molecule model for prediction of its modulus of elasticity, Constr. Build.
Constr. Build. Mater. 271 (2021), 121534. Mater. 243 (2020), 118176.
[4] Z. Zhang, Z. Tian, K. Zhang, X. Tang, Y. Luo, Preparation and characterization of [32] Y. Zhang, J. Zhang, J. Jiang, D. Hou, J. Zhang, The effect of water molecules on the
the greener alkali-activated grouting materials based on multi-index optimization, structure, dynamics, and mechanical properties of sodium aluminosilicate hydrate
Constr. Build. Mater. 269 (2021), 121328. (NASH) gel: A molecular dynamics study, Constr. Build. Mater. 193 (2018)
[5] B.C. McLellan, R.P. Williams, J. Lay, A. van Riessen, G.D. Corder, Costs and carbon 491–500.
emissions for geopolymer pastes in comparison to ordinary portland cement, [33] F. Lolli, H. Manzano, J.L. Provis, M.C. Bignozzi, E. Masoero, Atomistic Simulations
J. Cleaner Prod. 19 (9–10) (2011) 1080–1090. of Geopolymer Models: The Impact of Disorder on Structure and Mechanics, ACS
[6] L.K. Turner, F.G. Collins, Carbon dioxide equivalent (CO2-e) emissions: A Appl. Mater. Inter. 10 (26) (2018) 22809–22820.
comparison between geopolymer and OPC cement concrete, Constr. Build. Mater. [34] Y. Zhou, Z.C. Peng, J.L. Huang, T. Ma, X.M. Huang, C.W. Miao, A molecular
43 (2013) 125–130. dynamics study of calcium silicate hydrates-aggregate interfacial interactions and
[7] H. Gullu, A.A. Agha, The rheological, fresh and strength effects of cold-bonded influence of moisture, J. Cent. South Univ. 28 (1) (2021) 16–28.
geopolymer made with metakaolin and slag for grouting, Constr. Build. Mater. 274 [35] J.Z. Liu, B. Yu, Q.Z. Hong, Molecular dynamics simulation of distribution and
(2021). adhesion of asphalt components on steel slag, Constr. Build. Mater. 255 (2020),
[8] H. Gullu, M.M.D. Al Nuaimi, A. Aytek, Rheological and strength performances of 119332.
cold-bonded geopolymer made from limestone dust and bottom ash for grouting [36] H. Sun, Z. Jin, C.W. Yang, R.L.C. Akkermans, S.H. Robertson, N.A. Spenley,
and deep mixing, Bulletin of Engineering Geology and the Environment 80 (2) S. Miller, S.M. Todd, COMPASS II: extended coverage for polymer and drug-like
(2021) 1103–1123. molecule databases, J. Mol. Model. 22 (2) (2016) 10.
[9] H. Gullu, H.I. Fedakar, On the prediction of unconfined compressive strength of [37] D. Tavakoli, A. Tarighat, Molecular dynamics study on the mechanical properties
silty soil stabilized with bottom ash, jute and steel fibers via artificial intelligence, of Portland cement clinker phases, Comput. Mater. Sci. 119 (2016) 65–73.
Geomechanics and Engineering 12 (3) (2017) 441–464. [38] P.K. Sarkar, N. Mitra, Molecular level study of uni/multi-axial deformation
[10] Z.Y. Luo, W.G. Li, K.J. Wang, A. Castel, S.P. Shah, Comparison on the properties of response of tobermorite 11 angstrom: A force field comparison study, Cem. Concr.
ITZs in fly ash-based geopolymer and Portland cement concretes with equivalent Res. 145 (2021), 106451.
flowability, Cem. Concr. Res. 143 (2021), 106392. [39] J.S.R. Murillo, W. Hodo, A. Mohamed, R.V. Mohan, A. Rajendran, R. Valisetty,
[11] W. Zhang, M. Zheng, L. Zhu, Y. Ren, Y. Lv, Investigation of steel fiber reinforced A molecular dynamics investigation of hydrostatic compression characteristics of
high-performance concrete with full aeolian sand: Mix design, characteristics and mineral Jennite, Cem. Concr. Res. 99 (2017) 62–69.
microstructure, Constr. Build. Mater. 342 (2022), 128065. [40] C. Zhang, Z. Liu, P. Deng, Contact angle of soil minerals: A molecular dynamics
[12] W.S. Tasong, C.J. Lynsdale, J.C. Cripps, Aggregate-cement paste interface. II: study, Comput. Geotech. 75 (2016) 48–56.
Influence of aggregate physical properties, Cem. Concr. Res. 28 (10) (1998) [41] L. Kong, X. Quan, W. Luo, Y. Chen, B. Yang, H. Wang, Y. Zeng, Exploration of
1453–1465. molecular dynamics for the adsorption of anionic emulsifier on the main chemical
[13] E. Farahi, P. Purnell, N.R. Short, Supercritical carbonation of calcareous composition surface of aggregate, Constr. Build. Mater. 292 (2021), 123210.
composites: Influence of mix design, Cem. Concr. Compos. 43 (2013) 12–19. [42] W. Sun, H. Wang, Moisture effect on nanostructure and adhesion energy of asphalt
[14] P. Xie, J.J. Beaudoin, R. Brousseau, Effect of aggregate size on transition zone on aggregate surface: A molecular dynamics study, Appl. Surf. Sci. 510 (2020),
properties at the portland cement paste interface, Cem. Concr. Res. 21 (6) (1991) 145435.
999–1005. [43] Z. Tian, Z. Zhang, H. Liu, W. Zheng, X. Tang, Z. Gui, Interfacial characteristics and
[15] P.D. Silva, K. Sagoe-Crenstil, V. Sirivivatnanon, Kinetics of geopolymerization: mechanical behaviors of geopolymer binder with steel slag aggregate: Insights
Role of Al2O3 and SiO2, Cem. Concr. Res. 37 (4) (2007) 512–518. from molecular dynamics, J. Cleaner Prod. 362 (2022), 132385.
[16] S.J. Lyu, Y.H. Hsiao, T.T. Wang, T.W. Cheng, T.H. Ueng, Microstructure of [44] G.J. Xu, H. Wang, Study of cohesion and adhesion properties of asphalt concrete
geopolymer accounting for associated mechanical characteristics under various with molecular dynamics simulation, Comput. Mater. Sci. 112 (2016) 161–169.
stress states, Cem. Concr. Res. 54 (2013) 199–207. [45] M. Steinerova, J. Schweigstillova, POROUS MICROSTRUCTURE OF THE
[17] Y. Zhang, W. Sun, W. Lin, K. Zheng, J. Sha, S. Liu, In situ quantitatively tracking INTERFACIAL TRANSITION ZONE IN GEOPOLYMER COMPOSITES, Ceramics-
the hydration process of interfacial transition zone between coarse aggregate and Silikaty 57 (4) (2013) 328–335.
k-psds geopolymer matrix with esem, Journal of the Chinese Silicate Society 31 (8) [46] Q. Huang, C. Zhou, X. Gu, W. Zhang, Experimental study on moisture transport
(2003) 806–810. property of interfacial transition zone in concrete, Journal of Building Structures
[18] M.F. Kai, L.W. Zhang, K.M. Liew, New insights into creep characteristics of calcium 40 (1) (2019) 174–180.
silicate hydrates at molecular level, Cem. Concr. Res. 142 (2021), 106366. [47] D. Hou, X. Xu, Y. Ge, P. Wang, J. Chen, J. Zhang, Molecular structure, dynamics
[19] D.S. Hou, W. Zhang, M. Sun, P. Wang, M.H. Wang, J.R. Zhang, Z.J. Li, Modified and adsorption behavior of water molecules and ions on [0 1 0] surface of
Lucas-Washburn function of capillary transport in the calcium silicate hydrate gel γ-FeOOH: A molecular dynamics approach, Constr. Build. Mater. 224 (2019)
pore: A coarse-grained molecular dynamics study, Cem. Concr. Res. 136 (2020), 785–795.
106166. [48] Y. Zhou, D.S. Hou, H. Manzano, C.A. Orozco, G.Q. Geng, P.J.M. Monteiro, J.P. Liu,
Interfacial Connection Mechanisms in Calcium-Silicate-Hydrates/Polymer

14
Z. Tian et al. Construction and Building Materials 385 (2023) 131404

Nanocomposites: A Molecular Dynamics Study, ACS Appl. Mater. Inter. 9 (46) [52] W.A. Tasong, C.J. Lynsdale, J.C. Cripps, Aggregate-cement paste interface: Part I.
(2017) 41014–41025. Influence of aggregate geochemistry, Cem. Concr. Res. 29 (7) (1999) 1019–1025.
[49] Y. Tu, Q. Yu, R. Wen, P. Shi, L. Yuan, Y. Ji, G. Sas, L. Elfgren, Molecular dynamics [53] J. He, D. Lei, W. Xu, In-situ measurement of nominal compressive elastic modulus
simulation of coupled water and ion adsorption in the nano-pores of a realistic of interfacial transition zone in concrete by SEM-DIC coupled method, Cem. Concr.
calcium-silicate-hydrate gel, Constr. Build. Mater. 299 (2021), 123961. Compos. 114 (2020), 103779.
[50] C. Liu, F. Min, L. Liu, J. Chen, Hydration properties of alkali and alkaline earth [54] L. Kong, Y. Du, Interfacial interaction of aggregate-cement paste in concrete,
metal ions in aqueous solution: A molecular dynamics study, Chem. Phys. Lett. 727 Journal of Wuhan University of Technology-Mater, Sci. Ed. 30 (1) (2015) 117–121.
(2019) 31–37. [55] J. Li, Q. Zhang, Z. Chen, Aggregate Shape Analysis and Evaluation of Modified
[51] Y. Li, G.S. Zhang, D.S. Hou, Z.G. Wang, Nanoscale insight on the initial hydration Epoxy Wear Layer for Cement Pavement, Journal of Shenyang Jianzhu University,
mechanism of magnesium phosphate cement, Constr. Build. Mater. 276 (2021), Natural Science 33 (6) (2017) 1055–1064.
122213. [56] L. Hong, X.L. Gu, F. Lin, Influence of aggregate surface roughness on mechanical
properties of interface and concrete, Constr. Build. Mater. 65 (2014) 338–349.

15

You might also like