Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Case Studies in Construction Materials 16 (2022) e00981

Contents lists available at ScienceDirect

Case Studies in Construction Materials


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

Utilizing of coral/sea sand as aggregates in environment-friendly


marine mortar: Physical properties, carbonation resistance
and microstructure
Songsong He a, Chujie Jiao a, *, Yanfei Niu a, Song Li b
a
School of Civil Engineering, Guangzhou University, Guangzhou 510006, China
b
College of Urban and Rural Construction, Zhongkai University of Agriculture and Engineering, Guangzhou 510225, China

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

Keywords: The utilization of marine waste and resources to produce eco-friendly building materials is
Coral sand essential for the sustainable development of the construction industry on islands and coastal
Sea Sand areas. The sea sand and coral sand sourced from coral waste are the most promising. Therefore,
Physical properties
the objective of this work is to investigate the feasibility of coral/sea sand as fine aggregate to
Carbonation resistance
Microstructure
prepare an environment-friendly marine mortar. The effects of various volume fractions of coral
sand intended to replace sea sand on the flowability, physical performances, carbonation resis­
tance, and microstructure of the prepared marine mortars were evaluated. The experimental
results showed that as the sea sand replacement by coral sand was increased, the flowability and
dry bulk density decreased, while the water absorption increased. Encouragingly, usage of a
combination of coral and sea sands enhanced the flexural strength of marine mortars. A mixture
containing 80 vol% of coral sand showed the best effect. The carbonation depth of marine mortar
was found to noticeably increase with increasing replacement ratios of sea sand by coral sand,
closely relating to the increase in the volume of capillary and large pores. In particular, the
combined utilization of coral sand and sea sand resulted in a compacted interfacial transition zone
(ITZ) between fine aggregates and the matrix.

1. Introduction

The resource utilization and infrastructure construction of the islands in the ocean have flourished in recent years [1–3]. However,
one of the main problems for large-scale construction on such islands is the shortage of inexpensive yet reliable raw materials (e.g.,
gravel and sand) for concrete/mortar production [4,5]. Typically, such building materials are transported from the mainland, which
increases the construction cost and lengthens the construction times, not to mention high energy consumption and carbon dioxide
release [6,7]. Such an approach is unsustainable from economic, ecological and environmental points of view. In contrast, utilization
of discarded and natural marine resources (i.e., coral-based and sea sands) for manufacturing of building materials (in particular
concrete/mortar) could solve the problem of scarce island resources and protect the local ecology, both of which will realize the
sustainable development of these islands [1,8].
Coral wastes are widely accumulated on the beaches of tropical islands, mainly including coral debris and coral rubbles [9,10]. The

* Correspondence to: Guangzhou Higher Education Mega Center, 230 Wai Huan Xi Road, Guangzhou, China.
E-mail address: cejiaochujie@gzhu.edu.cn (C. Jiao).

https://doi.org/10.1016/j.cscm.2022.e00981
Received 8 December 2021; Received in revised form 20 February 2022; Accepted 23 February 2022
Available online 24 February 2022
2214-5095/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

former form due to natural weathering and erosion, whereas the latter is generated as a result of the local engineering projects such as
waterway dredging and infrastructure construction [3]. The coral debris and rubble are considered solid waste and occupy vast areas
of already limited island space. Reasonable exploitation of discarded coral materials will not deteriorate the living environment of
coral reefs nor will it destroy the marine ecological environment [11,12]. The United States was the first country to produce concrete
using coral waste as aggregate. It was used to build roads and airports on islands in the western Pacific during World War II, some of
which remain in service even today [13–15]. Since then, the utilization of coral waste as an aggregate for cement-based materials
spread worldwide [1–3]. However, previous studies [16–20] mainly focused on mechanical properties and durability of concrete with
coral wastes replacing coarse aggregates. Research on the usage of coral waste as fine aggregate to produce mortar/concrete is still
lacking. The study showed that using coral sand can reduce the workability of mortar/concrete due to the roughness and porosity of
the coral sand surface, which increases the needed unit water and paste content [3,21]. In terms of the mechanical properties and
durability, the coral sand mortar/concrete demonstrates high early compressive strength because of the internal curing effect while the
late compressive strength increases slowly [22,23]. Because coral sand is porous and possesses hardness lower than river sand, some
studies [24,25] have found that coral sand mortar/concrete possesses poorer compressive strength, drying shrinkage [21], chloride
resistance [26,27], sulphate attack [28], and abrasion performance [29] than river sand mortar/concrete. Although the addition of
silica fume and metakaolin to coral sand mortar/concrete can improve the corresponding compressive strength and durability [27,30],
this would result in additional construction costs and consume large amounts of energy. Therefore, another method is to appropriately
substitute fine aggregate with coral sand to produce mortar/concrete. Similar alternative methods to alleviate the shortage of natural
sand resources were recently reported [31–34]. Marvila et al. [32] investigated the replacement of river sand by granite residue in
cement mortars. The results indicated that the mortar with 40% granite residue achieves well applicability. Azevedo et al. [33] studied
the influence of replacing natural sand by construction waste on the properties of mortar. The results indicated that 25% was the
optimal replacement rate of construction waste. However, despite these results, the feasibility of substituting partially fine aggregate
with coral sand to produce mortar/concrete needs to be analyzed.
Besides coral sand, sea sand is an abundant resource on the oceanic islands and is also a potential substitution of fine aggregate
during mortar and concrete production [35]. Several researchers [36,37] have reported that the chemical composition and geological
origin of sea sand are similar to those of river sand; however, they have different characteristics. Existing research [35] showed that the
use of sea sand has a negligible effect on the workability of mortar/concrete. Moreover, some studies [38,39] show that the
compressive strength of concrete with sea sand is slightly higher than that of ordinary concrete. Yet, some studies disagree [40,41]. Sea
sand concrete had a higher flexural strength and dynamic modulus of elasticity relative to the ordinary concrete [42]. Additionally, sea
sand was also proved to improve the resistance of concrete on sulfate attack [43] and of carbonization [44].
The current studies have conducted a detailed analysis of the various properties of cement mortar/concrete containing coral or sea
sands. However, reports on the blended coral and sea sand properties for mortar/concrete production are lacking. The physical
properties and chemical compositions of the coral and sea sands are evidently different, and these dissimilarities might affect the
properties and microstructure of the mortar/concrete [9,35]. Both coral and sea sands contain chlorides and alkalis. Thus, the
buildings constructed with concrete containing coral and sea sands can increase the risk of reinforcement corrosion and cracking due to
volume expansion. Niu et al. [45] indicated that the chlorides in coral aggregate concrete can cause severe steel corrosion in concrete
structures. Ma et al. [46] also obtained similar results and noted that coral aggregate concrete corroded twice as fast as ordinary
concrete. Some studies [35,37] have found that the early volume expansion rate of sea sand concrete was higher than that of river sand
concrete. This was because that the Mg2+ and SO2− 4 introduced by sea sand, which could participate in the hydration reaction and
produce expanded hydration products [24]. Therefore, the application of the concrete prepared using coral and sea sands for building
structures is limited. However, the utilization of coral and sea sands to prepare mortar as a repair material in island engineering
projects for non-reinforced and fiber-reinforced polymer reinforced concrete structures is promising.
The objective of this study was to explore the feasibility of the combined utilization of coral and sea sand as fine aggregates to
prepare mortar (which called marine mortar). To analyze the effect of replacing sea sand by coral sand with various volume fractions
(0, 20, 40, 60, 80 and 100 vol%) on the performance of marine mortar. The workability of fresh marine mortar was investigated, whilst
the dry bulk density, water absorption, compressive strength, flexural strength, and accelerated carbonization properties of a hardened
marine mortar were evaluated. In addition, the hydration products, pore structure, and micromorphology of the marine mortar were
analyzed by X-ray diffraction (XRD), mercury intrusion porosimetry (MIP), and scanning electron microscopy (SEM), respectively. The
prepared marine mortar is expected to serve as a quick repair material in island engineering projects.

2. Experimental program

2.1. Raw materials

The fine aggregates considered in this study included coral sand, sea sand, and river sand. Table 1 presents their sources. The coral

Table 1
Sources of coral, sea and river sands used.
Type Coral sand Sea sand River sand

Source The South China Sea The East China Sea The Pearl River

2
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

sand was screened to obtain particle size fractions similar to those of sea and river sands. Fig. 1 exhibits their appearances. The particle
size distribution of all sand samples was determined using the standard GB/T 14684 [47] method, as shown Fig. 2. The physical
properties of the sand samples were tested by the GB/T 14684 [47] standard methods, as presented in Table 2.
Table 3 presents the chemical compositions of coral/sea sands identified by X-ray Fluorescence Spectrometry (XRF) method
performed by the ARL Perform’ X 4200 instrument. Fig. 3 shows the X-ray diffraction patterns of coral/sea sands measured by the
Panalytical X′ Pert Pro MPD instrument (equipped with CuKa radiation) at 45 kV and 50 mA in the 5–80◦ 2θ range at 6◦ /min scan
speed. The coral/sea sands used were ground into the powder samples (<70 µm) for XRD analyses. The SEM images of coral/sea sand
samples was observed by JEOL JSM-7001F setup at 15 kV, 85 mA, and 1000X, as exhibited in Fig. 4. The sand samples were coated
with a gold layer before testing. It can be found that the sea sand and coral sand with contrasting characteristics in physical properties
and chemical structures. To be specific, the coral sand demonstrated higher water absorption and crush index while its bulk density
and apparent density were lower than those obtained for the river and sea sand. The salt content of coral sand was higher than that of
sea sand. The coral sand was classified as calcareous sand, as confirmed by the presence of the strong and dominant peak of CaCO3, and
a small amount of halite crystal phases was also detected. The major crystalline phase of sea sand was SiO2. The coral sand particles can
be characterized by protruding horns and irregular shapes (as shown Fig. 5a). The surface of the coral sand particles is rough and
porous (as shown Fig. 4a), which endow coral sand with a good adsorption capacity. Unlike coral sand, sea sand particles exhibited
more rounded corners and somewhat regular shapes with smooth and dense surfaces (see Figs. 4b and 5b).
P II 42.5 Portland cement was selected as a binder. Its physical properties were determined using GB/T 175 [48] standard methods,
as listed in Table 4. The chemical composition of Portland cement was analyzed by XRF spectrometer using the same test method as
coral/sea sands (see Table 3). A polycarboxylate superplasticizer (SP), with a 30% water-reducing ratio and 25% solidity content was
also adopted.

2.2. Mix proportions and samples preparation

The mix proportions of marine mortars prepared in this work were listed in Table 5. The fine aggregate in the mix design of marine
mortar was changed from sea sand to coral sand at replacement ratios of 0, 20, 40, 60, 80, and 100 vol% by volume fraction, and the
marine mortar samples were denoted by SS, CS20, CS40, CS60, CS80, and CS100, respectively. In addition, the river sand mortar
(marked as RS) prepared using 100% river sand was selected as a reference for the marine mortar. The coral sand used need to be pre-
wetted to compensate for water absorption by the internal pore. In this work, the coral sand was used in the saturated surface dry state.
The cement content and mass-based water-to-cement ratio were kept constant at 675 kg/m3 and 0.45 in all the mortar mixtures,
respectively. The sand-to-cement ratio was fixed at 2 in the RS and SS mortar mixtures, which were provided in terms of mass. The SP
was added to obtain desired workability of the RS and SS mixtures, and the corresponding slump flow value was 240 ± 5 mm.
To prepare mortar samples, based on the mortar mix ratio, the fine aggregate and cement were weighed and mixed use a mortar
mixer for 60 s at 140 ± 5 r/min followed by water and SP addition as well as 60 s and 120 s mixing steps at 140 ± 5 and 280 ± 5 r/
min, respectively. The resulting uniform fresh mortar was immediately subjected to slump flow tests, performed according to the GB/T
2419 standard [49]. Another fraction of each fresh samples was loaded into steel molds and compacted according to the GB/T 17671
[50] standard method, followed by curing at 20 ± 2 ◦ C and ≥ 95% relative humidity. After 24 h, the mortar samples were demolded
and placed in a water tank at 20 ± 2 ◦ C until testing. The three parallel samples of the mortar were prepared for all the tested series.

2.3. Testing and analytical methods

2.3.1. Dry bulk density and water absorption tests


To measure the dry bulk density and water absorption of mortar samples [51,52]. The mortar samples (40 × 40 × 160 mm3) were
removed after 28 days of curing, and the weight of the samples in water was tested using an electronic static water balance. Subse­
quently, the excess water on the sample surface was wiped off with a damp cloth, and the sample weight was measured again under the
saturated surface dry condition in air. The samples were then heated at 105 ± 2 ℃ until a constant weight was obtained. The dry bulk
density (φd in kg/m3) and water absorption (Ψ w, in %) were calculated by Eqs. (1) and (2), respectively:

Fig. 1. Appearances of sand samples: (a) coral sand; (b) sea sand; (c) river sand.

3
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 2. Particle size distribution of coral, sea and river sands.

Table 2
Physical properties of different types of sands.
Type Apparent density (kg/m3) Bulk density (kg/m3) Water absorption (%) Crush index (%) Fineness modulus

Coral sand 2438 1174 4.7 39.6 3.0


Sea sand 2625 1620 0.6 16.5 2.8
River sand 2614 1607 0.5 15.2 2.9

Table 3
Main chemical compositions of cement, coral sand and sea sand (in wt%).
Material SiO2 Al2O3 Fe2O3 CaO MgO SO3 K2O Na2O CI⁻ LoI

Portland cement 19.71 5.58 3.43 61.32 3.03 2.37 0.44 0.13 – 1.49
Coral sand 0.52 0.12 0.09 49.24 2.51 0.32 0.46 1.11 0.11 44.73
Sea sand 89.75 3.41 1.76 1.68 0.58 0.14 0.21 0.34 0.05 1.34

Fig. 3. XRD pattern of coral and sea sands.

GDry
φd = (1)
GSSD − GWater

4
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 4. SEM image of sand particle surface: (a) coral sand; (b) sea sand.

Fig. 5. Digital image of sand particles: (a) coral sand; (b) sea sand.

Table 4
Physical properties of the P II 42.5 Portland cement.
Specific gravity Specific surface area (m2⋅kg) Setting time (min) Flexural strength (MPa) Compressive strength (MPa)

Initial Final 3d 28d 3d 28d

3.08 371 124 208 5.7 (0.04)* 7.9 (0.04) 35.9 (0.05) 48.5 (0.05)

*(the coefficient of variation of the results obtained)

Table 5
Mix proportion of mortar mixtures.
Mix type RS SS CS20 CS40 CS60 CS80 CS100

Cement (kg/m3) 675 675 675 675 675 675 675


Water (kg/m3) 304 304 304 304 304 304 304
Superplasticizer (kg/m3) 3.1 3.1 3.1 3.1 3.1 3.1 3.1
Coral sand (vol%) 0 0 20 40 60 80 100
Sea Sand (vol%) 0 100 80 60 40 20 0
River Sand (vol%) 100 0 0 0 0 0 0

GSSD − GDry
ψw = × 100% (2)
GDry

5
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Where: GDry is the weight of the dried mortar samples (kg), GSSD is the weight of the mortar samples under the saturated surface dry
condition (kg), and GWater is the weight of the mortar samples in water (kg).

2.3.2. Mechanical property tests


The flexural and compressive strength of the mortar samples cured for 3, 7, 28 and 60 days were measured by a pressing machine
according to the GB/T 17671 standard [50] at 50 and 2400 N/s loading rates, respectively. The flexural strength was performed using
40 × 40 × 160 mm3 prismatic samples. The compressive strength was determined from broken halves of mortar prisms obtained after
the flexural tests.

2.3.3. Accelerated carbonation test


An accelerated carbonization experiment was conducted on the mortar samples [48]. The mortar samples (40 × 40 × 160 mm3)
were cured for 26 days and then heated at 60 ± 2 ℃ for 48 h. Two opposite sides of the sample block were sealed with heated paraffin
wax as shown in Fig. 6c, after which they were placed in an accelerated carbonization environment chamber under 20 ± 3% CO2
atmosphere (see Fig. 6a and 6b). The temperature and relative humidity inside the chamber were equal to 20 ± 2 ℃ and 70 ± 5%,
respectively.
The carbonized mortar samples were split at carbonization ages of 7, 14, 28, and 56 days, and 1% phenolphthalein solution was
sprayed on the split inner section. The uncarbonized inner area turned purple, while the carbonized outer area remained self-colored.
The carbonation depth was judged by the contrast between the purple and grey areas, measured with a ruler in seven locations on each
side (see Fig. 7).

2.3.4. Microstructure analysis


The microstructural characteristics of the mortar samples were analyzed by SEM, XRD, and MIP after 28 days of curing. The mortar
samples were crushed into small pieces, soaked in the anhydrous ethanol for 48 h to stop the hydration reaction and replace the free
water and then dried at 55 ◦ C for 72 h to evaporate the ethanol. The resulting samples were coated with a gold layer and analyzed by a
JEOL JSM-7001F SEM at 15 kV to observe the ITZ morphology of the mortars. MIP tests were performed using Micromeritics Autopore
IV 9500 setup at intrusion pressure in the 1.4–214 MPa range, 130◦ mercury contact angle and 0.485 N/m surface tension to analyze
the pore structure of the mortars. XRD tests were performed by the Panalytical X ’Pert Pro MPD instrument in the 5–70◦ 2θ range at 6◦ /
min speed using ground to powder samples to identify the main phase compositions of the mortars.

3. Results and discussion

3.1. Slump flow of fresh mortars

Fig. 8 shows the effect of coral sand and sea sand contents on the slump flow of fresh marine mortar. It can be found that the slump
flow of the fresh marine mortar decreased with increasing replacements of sea sand by coral sand. In detail, the slump flow decreases
significantly from 245 mm to 134 mm with the increase in the coral sand content from 0 to 100 vol%. The decrease in the slump flow
for the fresh marine mortar can be attributed to two reasons: one is that the surface of the coral sand particles is rough and porous,
which can absorb part of the cement paste and water; the other is that the coral sand contained some flake and strip particles, which
enhanced the frictional resistance between the coral sand and the cement paste [53]. In contrast, more spherical and denser sea and
river sand particles allowed fresh mortars (SS and RS, respectively) to exhibit good fluidity [54]. A consistent trend was also reported
by other researchers. Zhang et al. [21] indicated that the presence of coral sand reduced the flowability of the mortar. Fonteboa et al.
[55] observed this trend when used mussel shell waste as fine aggregate.

Fig. 6. Accelerated carbonation test: (a) setup, (b) chamber and (c) test sample.

6
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 7. Carbonation depth determination of mortar samples.

Fig. 8. Slump flow of fresh mortars.

3.2. Dry bulk density and water absorption

Fig. 9 shows the influence of replacing sea sand by coral sand on the dry bulk density of the marine mortars. It can be observed that
the dry bulk density of the marine mortar decreases with the increase in the coral sand content. Specifically, the dry bulk density of the

Fig. 9. Dry bulk density of marine mortars.

7
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

marine mortar reduced from 2116 to 1896 kg/m3 with the substitution level of coral sand increasing from 0 to 100 vol%, which is
attributed mainly to the lower density of coral sand than sea sand. In addition, owing to the coral sand particles have protruding
corners and irregular shapes, resulting in a higher air content and lower density in the marine mortar [9,56]. The dry bulk densities of
the RS and SS mortars were higher than that of the marine mortar containing coral sand. Those data agree with the expected results
that a mortar containing denser fine aggregates with smoother and rounder shapes (both of which are favorable for providing a better
filling effect) will possess higher dry bulk density [57]. The same results were reported by Poon et al. [58]. and Senhadji et al. [59].
The effect of varying the coral sand content on the water absorption of the marine mortars is summarized in Fig. 10. It can be found
that the water absorption of marine mortar increases from 7.7% to 10.5% with the increase in the replacement ratio of sea sand by
coral sand from 0 to 100 vol%. The water absorption of the marine mortar increases with the addition of coral sand because the coral
sand has a high-water absorption due to the porous surface [53]. Moreover, the water absorption of marine mortars generally increases
as their densities decrease because it becomes more difficult for water to penetrate the dense matrix [54]. As expected, the RS mortar
has a low water absorption under the above affecting factors. Marvila et al. [31]. demonstrated similar water absorption results ob­
tained for the mortar prepared with PET recycled sand. When 20% of river sand was replaced by PET recycled sand, the water ab­
sorption exceeded 14%.

3.3. Mechanical strength

3.3.1. Compressive strength


Fig. 11 shows the effect of replacing sea sand with coral sand on the compressive strength of the marine mortar at varying curing
ages. In the first 7 days, the compressive strength of the marine mortar increases rapidly with the increase in the coral sand content. For
instance, compared with the SS mortar (30.4 MPa), the 3-day compressive strengths of marine mortar for coral sand contents of 20%,
40%, 60%, 80%, and 100% increased by 4.9%, 7.5%, 11.3%, 21.7%, and 23.6%, respectively. The obtained test results indicate that
the addition of coral sand enhanced the early compressive strength of the marine mortar. Consistent consequences were found in the
study by Gao at al. [8]. This phenomenon could be explained by the fact that coral sand absorbed some water from the cement paste,
which reduced the local water/binder ratio and improved the matrix’s strength [54]. Secondly, the coral sand stored water in the
casting process and released water in the hardening process, and played the role of internal curing, which is also conducive to enhance
the compressive strength of marine mortar [8,21]. Additionally, the marine mortars were produced from coral sand with high chloride
content, the chloride ions promote the cement hydration rate, resulting in improved early compressive strength of the resulting marine
mortar [3,8]. By contrast, the compressive strength of the marine mortar decreased steadily with the increase in the coral sand content
after 28 days of curing. To be specific, the 28-day compressive strength of the SS mortar is 51.6 MPa, while those of marine mortars
with 20%, 40% 60%, 80%, and 100% of coral sand decrease by 6.7%, 7.8%, 9.9%, 11.9%, and 14.0%, respectively. It should be noted
that the reduction in compressive strength was directly associated with the hardness of the fine aggregate [60]. Coral sands have higher
brittleness and breakage index than sea sand due to its porous structure [61]. The negative influence on strength be resulted from the
porous structure of coral sand overwhelmed the aforementioned multiple advantageous effects. Therefore, the later compressive
strength decreased with the coral sand content increase. The RS mortar exhibited the highest compressive strength (53.99 MPa at
28 day) compared with the marine mortars because the physical and geometric properties of river sand are conducive to generate a
high strength. The typical compressive strength of mortar for repairs must be above 20 MPa according to the standard JC/T 2381 [62].
Thus, the studied marine mortar containing coral/sea sands met these requirements. Furthermore, the high early compressive strength
confirms the feasibility of utilizing coral and sea sand as fine aggregates for the marine mortars as rapid repair materials.

Fig. 10. Water absorption of marine mortars.

8
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 11. Compressive strength of mortars at varying curing ages.

3.3.2. Flexural strength


Fig. 12 shows the influence of replacing sea sand with coral sand on the flexural strength of the marine mortars at varying curing
ages. The flexural strength of the marine mortar increased as the content of coral sand in the fine aggregate was increased. This in­
crease was the most rapid during the first 7 days of cure and then increases gradually. The flexural strength of the SS mortar sample
cured for 28 days was 7.83 MPa. The flexural strengths of the marine mortar increased by 0.8%, 2.1% and 5.1% as the coral sand
content was increased to 20, 40 and 60 vol%, respectively. The marine mortar CS80 exhibits the highest flexural strength, which is
8.26% higher than that of the mortar SS at 28 days. However, marine mortars containing 100% of coral sand started showing a slight
decrease in their flexural strength, which was still 3.9% higher than for the SS mortar sample. These findings demonstrate that the
combined utilization of coral sand and sea sand as fine aggregate could improve the flexural strength of marine mortar, and the
optimum replacement proportion of the coral sand was found to be 80 vol%. This phenomenon could be explained by the rough and
porous surface of coral sand, which was expected to improve the compactness and the formation of the dense ITZ between the sand and
the matrix [63]. Consequently, the flexural strength of the marine mortar is improved with the increase in the coral sand content. In
addition, the coral sand particles with flake- and strip-like morphologies very likely behaved as mineral fibers to enhance the flexural
strength of the marine mortar [8].
The schematic of the fracture characteristics of the marine mortar containing coral sand is proposed on the basis of the observed
results, as shown in Fig. 13. From the fracture surface of the marine mortar samples after the flexural strength test, the broken coral
sands can be directly observed; however, almost no fracture failure occurred inside sea sand. This indicates that the primary failure
modes of marine mortar containing coral sand are manifested as the fracture of the coral sand and the debonding fracture of the sea
sand interface. The interfacial bond of the sea sand and the matrix is normally weak due to the smooth surface of the sea sand particles
[63,64]. Therefore, the fracture failure occurs more easily in the ITZ between the matrix and sea sand, rather than inside the matrix
itself or inside the sea sand grains. At the same time, the surfaces of the coral sand particles are rough and porous, which is very
beneficial for creating strong bonds of the ITZ between coral sand and the matrix. Wang et al. [8] demonstrated that the microhardness
of the coral sand/matrix ITZ is much higher than of the ITZ between the matrix and the river sand. Therefore, the strength of the coral
sand particles is below that of the matrix and the ITZ bonds, and fracture extension and advancement could proceed directly through
the coral sand.

3.4. Carbonation resistance

Fig. 14 shows the carbonization area of the tested mortar samples subjected to the accelerated carbonization tests for 7–56 days.
The corresponding carbonation depths were measured, as shown in Fig. 15. Notably, the carbonation depth of the marine mortar
markedly increases with the increase in the replacement ratio of sea sand by coral sand for the same carbonation age and further
increases with the extension of the accelerated carbonization age. Taking the accelerated carbonization at 56 days as an example, the
RS sample has the lowest carbonization depth of 4.78 mm. Compared with the carbonization depth of 5.32 mm for the SS sample, the
carbonization depth of the marine mortar for coral sand contents of 20%, 40%, 60%, 80%, and 100% (by volume) increased by 46.2%,
94.4%, 144.4%, 195.5%, and 230.6%, respectively. The main reason for the increase in carbonization depth was that the pore structure
of the marine mortar may change due to the addition of coral sand. An increase in the volume of harmful pores can lead to CO2 readily
transferring from the outside to the inside of the marine mortar, which would accelerate the carbonization process [55]. Additionally,
the porous structure of the coral sand particles with its channels promoted CO2 diffusion deeper into the matrix [65]. The addition of
supplementary cementitious materials (SCMs) can improve the carbonization resistance of the cement mortar [66]. Therefore, to
alleviate the adverse effects of the coral sand presence on carbonation resistance performance of the corresponding marine mortar,

9
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 12. Flexural strength of mortars at varying curing ages.

Fig. 13. Schematic of fracture pattern of marine mortar containing coral sand.

SCMs could be considered for adding.

3.5. Microstructural analysis

3.5.1. XRD analysis


Fig. 16 shows the XRD patterns of all the mortar samples aged for 28 days. The main crystalline phases in the marine mortar
contained quartz, Ca(OH)2, CaCO3, and Gismondine. The sources of CaCO3 and quartz in the marine mortar samples were considered
to be from coral sand and sea sand during the sample preparation, respectively [67]. The diffraction peak positions in the XRD patterns
of marine mortars with different volumes of the coral sand are similar, indicating that coral sand has little effect on the composition of
the hydration products. In addition, it can be found that the Ca(OH)2 and Gismondine contents increased with the coral sand fraction.
This could be explained by the fact that the addition of coral sand promotes mortar hydration as the porous coral sand particles adsorb
more chloride than sea sand particles. This explanation was supported by Wang et al. [53], which demonstrated that the chloride ions
from coral aggregate reacts with tricalcium aluminate in cement to form Friedel’s salt, resulting in accelerated hydration rate of
cement. Additionally, the internal curing effect of coral sand also promotes cement hydration [3].

3.5.2. Pore structure


The MIP test was used to further discuss the pore structure of the marine mortars. Fig. 17 shows the pore size distributions of the
mortar samples, and the corresponding pore size parameters are presented in Table 6. The incremental pore volume curves of the
marine mortars in this study show a bimodal distribution, and the peak value presents a significant rise trend with the increase in the
replacement ratio of sea sand by coral sand, as shown in Fig. 17a. The cumulative pore volume curves (Fig. 17b) show that with the
coral sand content increased, the curve shifts upward. The measurement results indicated that the porosity of marine mortar increases
distinctly due to the increase in the coral sand content. More specifically, the porosity increases from 14.44% for the SS sample by

10
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 14. Fracture surface of carbonized mortar samples treated with phenolphthalein, columns from left to right: RS, SS, CS20, CS40, CS60, CS80,
CS100; rows from up to bottom: carbonization days at 7, 14, 28, and 56 days.

Fig. 15. Carbonization depth of mortar samples at different carbonization ages.

10.2%, 20.1%, 24.9%, 28.8%, and 36.4% for marine mortar containing coral sand with contents of 20%, 40%, 60%, 80%, and 100% by
volume, respectively. Similarly, the average pore size of the marine mortar increases with the increasing addition of coral sand. To be
specific, the average pore size increases from 17.4 nm for the SS sample to 18.4, 20.3, 23.6, 36.1, and 52.6 nm for CS20, CS40, CS60,
CS80, and CS100 samples, respectively. The worsening of the pore structure of marine mortar could be contributed to the incorpo­
ration of the coral sand particles with porous surface, and its porosity is typically above 30% [3]. Additionally, flaky and rod-like
particles (with numerous angles and protruding facets) of coral sand do not stack and pack dense enough, which also increases the
porosity of the corresponding marine mortar [9,63].
The pore sizes of the cement mortar can typically be categorized as gel pores (with diameters (D) below 10 nm), transition pores (D
between 10 nm and 100 nm), capillary pores (D between 100 nm and 1000 nm) and large pores (D above 1000 nm) [68,69]. As the
coral sand content was increased from 0% to 100 vol% in the marine mortar, the increase in the porosity of marine mortar manifested

11
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

Fig. 16. XRD patterns of mortar samples at 28 days.

Fig. 17. (a) Incremental and (b) cumulative pore volume curves of the mortar samples.

Table 6
Pore size parameters of mortar samples.
Mix types RS SS CS20 CS40 CS60 CS80 CS100

Porosity /% 14.43 14.44 15.91 17.34 18.03 18.6 19.7


Average pore size /nm 17.4 17.4 18.4 20.3 23.6 36.1 52.6
Gel pores (D < 10 nm) /% 21.37 19.93 15.26 15.16 12.73 2.96 2.98
Transition pores (10 nm ≤ D < 100 nm) /% 66.26 64.30 63.96 56.14 51.67 52.67 50.39
Capillary pores (100 nm ≤ D < 1000 nm) /% 6.91 8.11 12.69 18.68 24.92 31.92 33.82
Large pores (D ≥ 1000 nm) /% 5.46 7.66 8.09 10.02 10.68 12.45 12.81

12
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

in the volume increase from 7.7% to 12.8% for large pores and from 8.1% to 33.8% for capillary pores and in the volume decrease from
19.9% to 3.0% for gel pores and from 64.3% to 50.4% for transition pores. It can be observed that the 85% reduction of gel pore volume
in marine mortar samples as the coral sand content was increased from 0% to 100%, while the capillary pore volume increased by
317.0%. Previous studies [70,71] reported that water and gas transports typically occur in the capillary pores of mortar, while the gel
pores hardly participate in mass transfer inside the mortar matrix. The increase in the volume of the capillary pores provides a channel
for water and CO2 to penetrate, which increases the water absorption and carbonation depth of marine mortar. Thus, this result
confirms that the marine mortar samples with higher coral sand content showed increased water absorption capacity and carbonation
depth. The porosities of RS and SS mortar samples were similar. Besides, the capillary pore volume of the RS sample (which was equal
to 8.1%) was higher than that of the SS sample (6.9%). It also explains the slightly higher carbonization depth and water absorption of
the SS sample than of the RS sample. Additionally, the resulting mortar with the higher the porosity values possessed the lower the
28-day compressive strength. However, the same trend was not well accorded with the flexural strength test results. The reason was
because that the mechanical strength of mortars was not only concerned to their porosity, but also decided on the matrix and ITZ
properties [63].

3.5.3. Micromorphology of ITZ


Due to the defects in the aggregate/matrix ITZ, it is generally considered the weakest zone of the concrete and mortar, negatively
affecting the mixture strength [69,70]. Fig. 18 presents the SEM images of the ITZ of mortar samples. As shown in Fig. 18a and 18b,
there are evident gaps in the ITZ between sand and cement paste in the RS and SS sample, which reflects a potentially poor bonding
between the sand and the matrix. Thus, resulting in a lower flexural strength compared with those of other mortar series. Since river
and sea sands have low water absorption capacity and smooth particle surfaces, a water-rich zone forms easily on the particle surfaces
due to the sidewall effects, which results in debonding around the sand particles [72].
After replacing sea sand with 60% of coral sand, CS60 sample showed a tight connection appeared in the ITZ between the sea sand
and cement paste (see Fig. 18c). This means that the evident interval of the ITZ between sea sand and cement paste in the marine
mortar is improved with the partial substitution of sea sand with coral sand. This beneficial phenomenon can be explained that the
high-water absorption of coral sand added reduces the water-rich zone on the surface of sea sand. SEM images of the ITZ of the CS60
and CS100 samples revealed that the coral sand was closely connected to the cement paste, and it is difficult to distinguish the
boundary of the ITZ (see Fig. 18d and 18e). This tight ITZ can be explained by the following reasons: (1) The surface of the coral sand
particles is very porous. The hydration products of needle-like ettringite and C-S-H gel can fill the pores to promote cement paste
interaction with the coral sand (see Fig. 18f). (2) The surface of the coral sand particles is also rough, which improves the interlocking
capacity of the coral sand and cement paste. Thus, the above results indicate that the combined use of sea sand and coral sand is
beneficial to alleviate the problem of poor bonding between sea sand and cement paste as well as provide a dense ITZ between the
cement paste and coral sand. The densification of ITZ improved, which enhanced the flexural strength of the marine mortar, which
corroborates the flexural strength test results.

Fig. 18. SEM images of ITZ between sand and cement paste for the (a) RS, (b) SS, (c, d) CS60 and (e, f) CS100 mortar samples.

13
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

4. Conclusions

This study demonstrated the feasibility of using sea sand and coral sand as fine aggregate to produce environment-friendly marine
mortar. The effects of coral sand content in this mixture on physical properties, carbonation resistance and microstructure of the
marine mortar were investigated. The main conclusions can be summarized as follows:

(1) The slump flow of the fresh marine mortar and dry bulk density of the marine mortar decreased with the increase in the
replacement rate of sea sand by coral sand, while the water absorption increased under the same state.
(2) The combined usage of coral sand and sea sand could increase the early compressive strength and flexural strength of marine
mortar. The flexural strength of marine mortar containing 80 vol% of coral sand in the fine aggregate was the best. The fracture
path under the bending load went through the coral sand particles and extended along the sea sand boundaries.
(3) The addition of coral sand impaired the capability of carbonization resistance for the marine mortar, and the carbonation depth
increased as the coral sand content was increased.
(4) The combined utilization of coral sand and sea sand in the marine mortar could form an interlocking interface between coral
sand and cement paste, and the compactness of the ITZ between sea sand and cement paste improved efficiently. The pore
volume of the capillary and large pores increased as more coral sand was added, whereas the pore volume of gel and transition
pores decreased.

In summary, it can be concluded that the utilization of coral sand and sea sand to substitute conventional sand during the pro­
duction of marine mortar is feasible. The combination of 80 vol% coral sand and 20 vol% sea sand is recommended. This approach will
contribute to the sustainable development of marine construction. The potential effects of chlorides and alkalis carried by coral/sea
sand particles on the durability of the resulting mortar should be further evaluated. Thus, it can be recommended that future studies
focus on the influence of coral/sea sands on chloride corrosion and alkali-aggregate reaction of the marine mortars, to complement the
understanding of using these sands on the properties of the prepared mortars.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements

The authors are grateful to the grant from the National Natural Science Foundation of China (Grant Nos. 52078148, 51778158 and
52108125); Natural Science Foundation of Guangdong Province (Grant No. 107059961065); Special Research Projects in Key Areas
for Colleges and Universities in Guangdong Province (Grant No. 2021ZDZX4009); Key Innovation Program of Water Conservancy
Science and Technology of Guangdong Province (No. 2017-32).

References

[1] A. Wang, B. Lyu, Z. Zhang, et al., The development of coral concretes and their upgrading technologies: a critical review, Constr. Build. Mater. 187 (30) (2018)
1004–1019.
[2] J. Liu, Z. Ou, P. Wei, et al., Literature review of coral concrete, Arab. J. Sci. Eng. 43 (4) (2017) 1–13.
[3] L. Zhou, S. Guo, Z. Zhang, et al., Mechanical behavior and durability of coral aggregate concrete and bonding performance with fiber-reinforced polymer (FRP)
bars: a critical review, J. Clean. Prod. 289 (2021), 125652.
[4] Y. Huang, X. He, H. Sun, et al., Effects of coral, recycled and natural coarse aggregates on the mechanical properties of concrete, Constr. Build. Mater. 192
(2018) 330–347.
[5] W.Y. Xu, S.T. Yang, C.J. Xu, et al., Study on fracture properties of alkali-activated slag seawater coral aggregate concrete, Constr. Build. Mater. 223 (2019)
91–105.
[6] L. Ma, Z. Li, J. Liu, et al., Mechanical properties of coral concrete subjected to uniaxial dynamic compression, Constr. Build. Mater. 199 (28) (2019) 244–255.
[7] X. Wang, R. Yu, Z. Shui, et al., Mix design and characteristics evaluation of an eco-friendly Ultra-High Performance Concrete incorporating recycled coral based
materials, J. Clean. Prod. 165 (2017) 70–80.
[8] Y. Wang, Z. Shui, X. Gao, et al., Utilizing coral waste and metakaolin to produce eco-friendly marine mortar: hydration, mechanical properties and durability,
J. Clean. Prod. 219 (2019) 763–774.
[9] A. Wang, Z. Zhang, K. Liu, et al., Coral aggregate concrete: numerical description of physical, chemical and morphological properties of coral aggregate, Cem.
Concr. Compos. 100 (2019) 25–34.
[10] Y.U. Hongbing, S.U.N. Zongxun, T. Cheng, Physical and mechanical properties of coral sand in the Nansha Islands, Mar. Sci. Bull. 8 (2) (2006) 31–39.
[11] R. Vasseur, B. Lathuilière, I. Lazăr, et al., Major coral extinctions during the early Toarcian global warming event, Glob. Planet. Change 207 (2021), 103647.
[12] L. Zhang, D. Niu, B. Wen, et al., Initial-corrosion condition behavior of the Cr and Al alloy steel bars in coral concrete for marine construction, Cem. Concr.
Compos. 120 (2021), 104051.
[13] D.L. Narver, Good concrete made with coral and sea water, Civ. Eng. 24 (11) (1954) 49–52.
[14] C.H. Scholer, Examination and study of certain structures in the Pacific Ocean area, Progress Report, USNCEL, Contract NBy-3171, USN CEL, (1959).
[15] P.A. Howdyshell, The use of Coral as an Aggregate for Portland Cement Concrete Structures, Construction Engineering Research Lab (Army), Champaign IL,
1974.
[16] D.S. Chen, Effect of Different Mixing Methods on Performance of Porous Aggregate Concrete (M.S. Thesis), Guangxi University, Nanning, China, 2017.
[17] Yingjie Chu, Aiguo Wang, Yingcan Zhu, et al., Enhancing the performance of basic magnesium sulfate cement-based coral aggregate concrete through gradient
composite design technology, Compos. Part B Eng. 227 (2021), 109382.

14
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

[18] Y. Tan, H. Yu, R. Mi, et al., Compressive strength evaluation of coral aggregate seawater concrete (CAC) by non-destructive techniques, Eng. Struct. 176 (2018)
293–302.
[19] H. Ma, C. Yue, H. Yu, et al., Experimental study and numerical simulation of impact compression mechanical properties of high strength coral aggregate
seawater concrete, Int. J. Impact Eng. 137 (2020), 103466.
[20] B. Liu, J. Guo, X. Wen, et al., Study on flexural behavior of carbon fibers reinforced coral concrete using digital image correlation, Constr. Build. Mater. 242
(2020), 117968.
[21] B. Zhang, H. Zhu, K.W. Shah, et al., Performance evaluation and microstructure characterization of seawater and coral/sea sand alkali-activated mortars, Constr.
Build. Mater. 259 (2020), 120403.
[22] Li Wenjie, Chen Yihu, Fan Liyun, et al., Experimental study and microstructure analysis of calcareous sand cement mortar, Mater. Rep. 34 (S1) (2020) 224–228.
[23] X. Zhang, J. Zuo, Z. Wang, et al., The evolution of the microstructure and mechanical properties of coral aggregate mortar under uniaxial compression using
ultrasonic analysis, Constr. Build. Mater. 300 (2021), 124000.
[24] J.H. Mei, J. Wu, L.X. Wang, J.P. Mei, et al., Mechanical properties and microstructural characteristics of coral sand mortar, J. Build. Mater. 23 (02) (2020)
263–270.
[25] W. Zhou, P. Feng, H. Lin, Constitutive relations of coral aggregate concrete under uniaxial and triaxial compression, Constr. Build. Mater. 251 (2020), 118957.
[26] Y. Wang, Z. Shui, R. Yu, et al., Chloride ingress and binding of coral waste filler-coral waste sand marine mortar incorporating metakaolin, Constr. Build. Mater.
190 (2018) 1069–1080.
[27] Y. Wang, Z. Shui, Y. Huang, et al., Properties of coral waste-based mortar incorporating metakaolin: part II. Chloride migration and binding behaviors, Constr.
Build. Mater. 174 (2018) 433–442.
[28] J. Tang, H. Cheng, Q. Zhang, et al., Development of properties and microstructure of concrete with coral reef sand under sulphate attack and drying-wetting
cycles, Constr. Build. Mater. 165 (2018) 647–654.
[29] Q. Qin, Q. Meng, H. Yang, et al., Study of the anti-abrasion performance and mechanism of coral reef sand concrete, Constr. Build. Mater. 291 (2021), 123263.
[30] W. Wu, R. Wang, C. Zhu, et al., The effect of fly ash and silica fume on mechanical properties and durability of coral aggregate concrete, Constr. Build. Mater.
185 (2018) 69–78.
[31] A.F. Campanhão, M.T. Marvila, A.R.G. de Azevedo, et al., Recycled PET sand for cementitious mortar, Materials 15 (1) (2022) 273.
[32] A.R.G. Azevedo, M.T. Marvila, L. da Silva Barroso, et al., Effect of granite residue incorporation on the behavior of mortars, Materials 12 (9) (2019) 1449.
[33] A.R.G. Azevedo, D. Cecchin, D.F. Carmo, et al., Analysis of the compactness and properties of the hardened state of mortars with recycling of construction and
demolition waste (CDW), J. Mater. Res. Technol. 9 (3) (2020) 5942–5952.
[34] L.F. Amaral, G.C.G. Delaqua, M. Nicolite, et al., Eco-friendly mortars with addition of ornamental stone waste-a mathematical model approach for granulometric
optimization, J. Clean. Prod. 248 (2020), 119283.
[35] Y. Zhao, X. Hu, C. Shi, et al., A review on seawater sea-sand concrete: mixture proportion, hydration, microstructure and properties, Constr. Build. Mater. 295
(2021), 123602.
[36] T. Dhondy, A. Remennikov, M. Neaz Sheikh, Properties and application of sea sand in sea sand–seawater concrete, J. Mater. Civ. Eng. 32 (12) (2020), 04020392.
[37] J. Xiao, C. Qiang, A. Nanni, et al., Use of sea-sand and seawater in concrete construction: current status and future opportunities, Constr. Build. Mater. 155
(2017) 1101–1111.
[38] P. Li, W. Li, T. Yu, et al., Investigation on early-age hydration, mechanical properties and microstructure of seawater sea sand cement mortar, Constr. Build.
Mater. 249 (2020), 118776.
[39] Y. Huang, X. He, Q. Wang, et al., Mechanical properties of sea sand recycled aggregate concrete under axial compression, Constr. Build. Mater. 175 (2018)
55–63.
[40] J. Limeira, M. Etxeberria, L. Agulló, et al., Mechanical and durability properties of concrete made with dredged marine sand, Constr. Build. Mater. 25 (11)
(2011) 4165–4174.
[41] C.G. Girish, D. Tensing, K.L. Priya, Dredged offshore sand as a replacement for fine aggregate in concrete, Int. J. Eng. Sci. Emerg. Technol. 8 (3) (2015) 88–95.
[42] D. Pan, S.A. Yaseen, K. Chen, et al., Study of the influence of seawater and sea sand on the mechanical and microstructural properties of concrete, J. Build. Eng.
42 (2021), 103006.
[43] S. Han, J. Zhong, Q. Yu, et al., Sulfate resistance of eco-friendly and sulfate-resistant concrete using seawater sea-sand and high-ferrite Portland cement, Constr.
Build. Mater. 305 (2021), 124753.
[44] W. Liu, H. Cui, Z. Dong, et al., Carbonation of concrete made with dredged marine sand and its effect on chloride binding, Constr. Build. Mater. 120 (2016) 1–9.
[45] D. Niu, L. Zhang, Q. Fu, et al., Critical conditions and life prediction of reinforcement corrosion in coral aggregate concrete, Constr. Build. Mater. 238 (2020),
117685.
[46] Z. Wu, H. Yu, H. Ma, et al., Rebar corrosion in coral aggregate concrete: determination of chloride threshold by LPR, Corros. Sci. 163 (2020), 108238.
[47] GB/T 14684-2001. Sand for Building, China Building Industry Press, Beijing, 2001 (In Chinese).
[48] GB/T 175-2007. Common Portland Cement, China Building Industry Press, Beijing, 2007 (In Chinese).
[49] GB/T 2419-2016. Test Method for Fluidity of Cement Mortar, Standards Press of China, Beijing, 2016.
[50] GB/T 17671-1999, Method of testing cements - Determination of strength (ISO 679:2009 Cement test methods - Determination of strength, MOD), China
Building Industry Press, Beijing, 1999 (in Chinese).
[51] GB/T 50081-2019. Standard for Test Methods of Concrete Physical and Mechanical Properties, China Building Industry Press, Beijing, 2019 (in Chinese).
[52] BS EN 12390-7. Testing Hardened Concrete-Part 7 Density of Hardened Concrete, British Standard Institution, London, 2019.
[53] Y. Wang, S. Zhang, D. Niu, et al., Strength and chloride ion distribution brought by aggregate of basalt fiber reinforced coral aggregate concrete, Constr. Build.
Mater. 234 (2020), 117390.
[54] W. Xu, X. Wen, J. Wei, et al., Feasibility of kaolin tailing sand to be as an environmentally friendly alternative to river sand in construction applications, J. Clean.
Prod. 205 (2018) 1114–1126.
[55] C. Martínez-García, B. González-Fonteboa, D. Carro-López, et al., Impact of mussel shell aggregates on air lime mortars. Pore structure and carbonation, J. Clean.
Prod. 215 (2019) 650–668.
[56] I. Raini, R. Jabrane, L. Mesrar, et al., Evaluation of mortar properties by combining concrete and brick wastes as fine aggregate, Case Stud Constr. Mater. 13
(2020), e00434.
[57] A.K.H. Kwan, C.F. Mora, Effects of various shape parameters on packing of aggregate particles, Mag. Concr. Res. 53 (2) (2001) 91–100.
[58] J.X. Lu, P. Shen, H. Zheng, et al., Synergetic recycling of waste glass and recycled aggregates in cement mortars: physical, durability and microstructure
performance, Cem. Concr. Compos. 113 (2020), 103632.
[59] T. Cheboub, Y. Senhadji, H. Khelafi, et al., Investigation of the engineering properties of environmentally-friendly self-compacting lightweight mortar
containing olive kernel shells as aggregate, J. Clean. Prod. 249 (2020), 119406.
[60] V. Akyuncu, F. Sanliturk, Investigation of physical and mechanical properties of mortars produced by polymer coated perlite aggregate, J. Build. Eng. 38 (2021),
102182.
[61] J. Shen, X.U. Dongsheng, Z. Liu, et al., Effect of particle characteristics stress on the mechanical properties of cement mortar with coral sand, Constr. Build.
Mater. 260 (2020), 119836.
[62] GB/T 175-2016. Repairing mortar, China Building Industry Press, Beijing, 2016 (In Chinese).
[63] Q. Huang, X. Zhu, G. Xiong, et al., Recycling of crushed waste clay brick as aggregates in cement mortars: An approach from macro-and micro-scale
investigation, Constr. Build. Mater. 274 (2021), 122068.
[64] C. Perry, J.E. Gillott, The influence of mortar-aggregate bond strength on the behaviour of concrete in uniaxial compression, Cem. Concr. Res. 7 (5) (1977)
553–564.

15
S. He et al. Case Studies in Construction Materials 16 (2022) e00981

[65] J.J. De Oliveira Andrade, E. Possan, J.Z. Squiavon, et al., Evaluation of mechanical properties and carbonation of mortars produced with construction and
demolition waste, Constr. Build. Mater. 161 (2018) 70–83.
[66] C. Wang, K. Sun, H. Niu, et al., Effect of binary admixture of sepiolite and fly ash on carbonation and chloride resistance of modified cement mortar, Constr.
Build. Mater. 279 (2021), 122509.
[67] H. Shi, Z. Yu, J. Ma, et al., Properties of Portland cement paste blended with coral sand powder, Constr. Build. Mater. 203 (2019) 662–669.
[68] M. Sun, C. Zou, D. Xin, Pore structure evolution mechanism of cement mortar containing diatomite subjected to freeze-thaw cycles by multifractal analysis,
Cem. Concr. Compos. 114 (2020), 103731.
[69] R.A. Olson, C.M. Neubauer, H.M. Jennings, Damage to the pore structure of hardened Portland cement paste by mercury intrusion, J. Am. Ceram. Soc. 80 (9)
(1997) 2454–2458.
[70] H.W. Song, S.J. Kwon, Permeability characteristics of carbonated concrete considering capillary pore structure, Cem. Concr. Res. 37 (6) (2007) 909–915.
[71] P. Hou, X. Cheng, J. Qian, et al., Characteristics of surface-treatment of nano-SiO2 on the transport properties of hardened cement pastes with different water-to-
cement ratios, Cem. Concr. Compos. 55 (2015) 26–33.
[72] K.L. Scrivener, A.K. Crumbie, P. Laugesen, The interfacial transition zone (ITZ) between cement paste and aggregate in concrete, Interface Sci. 12 (4) (2004)
411–421.

16

You might also like