ArticlePDF Available

Boosting the mechanical performance and fire resistivity of white ordinary portland cement pastes via biogenic mesoporous silica nanoparticles

Springer Nature
Scientific Reports
Authors:

Abstract and Figures

This study investigates how biogenic mesoporous silica nanoparticles (MS-NPs) extracted from rice straw residues, a sustainable and economical bio-source, affect White Ordinary Portland Cement (WOPC) paste performance. A comprehensive investigation using varied fractions of 0.25, 0.50, 0.75, and 1.0% MS-NPs as an additive to WOPC was conducted to analyze the physicomechanical characteristics of WOPC-MS hardened composites, including compressive strength, fire resistance, and water demand. The beneficial impact of biogenic MS-NPS was verified by X-ray diffraction (XRD), Scanning Electron Microscopy (SEM), and differential thermo-gravimetric analysis (TGA/DTG) methods, revealing several hydration products such as calcium silicate hydrates (CSHs), calcium ferrosilicate hydrates (CFSHs), and calcium aluminosilicate hydrates (CASHs). These products enhance the overall physical and mechanical properties and the thermal stability of hardened WOPC-MS. The composite comprising WOPC-0.75 MS provides numerous advantages from both an economic and environmental perspective.
This content is subject to copyright. Terms and conditions apply.
Boosting the mechanical
performance and re resistivity of
white ordinary portland cement
pastes via biogenic mesoporous
silica nanoparticles
Abdallah A. Aziz1, Hossam F. Nassar1, Mona T. Al-Shemy2 & O. A. Mohamed1
This study investigates how biogenic mesoporous silica nanoparticles (MS-NPs) extracted from rice
straw residues, a sustainable and economical bio-source, aect White Ordinary Portland Cement
(WOPC) paste performance. A comprehensive investigation using varied fractions of 0.25, 0.50,
0.75, and 1.0% MS-NPs as an additive to WOPC was conducted to analyze the physicomechanical
characteristics of WOPC-MS hardened composites, including compressive strength, re resistance,
and water demand. The benecial impact of biogenic MS-NPS was veried by X-ray diraction
(XRD), Scanning Electron Microscopy (SEM), and dierential thermo-gravimetric analysis (TGA/DTG)
methods, revealing several hydration products such as calcium silicate hydrates (CSHs), calcium
ferrosilicate hydrates (CFSHs), and calcium aluminosilicate hydrates (CASHs). These products enhance
the overall physical and mechanical properties and the thermal stability of hardened WOPC-MS. The
composite comprising WOPC-0.75 MS provides numerous advantages from both an economic and
environmental perspective.
Keywords Coats-Redfern, Eco-friendly mesoporous silica nanoparticles (MS-NPs), Sustainability, White
Ordinary Portland Cement (WOPC), ermal stability
One of the most ubiquitous building materials in the world is concrete. e composites heterogeneity further
adds to its complexity. One of the main ingredients in concrete is cement. Cement production is a major
polluter because it produces waste products that are both extremely poisonous and destructive1. Carbon
dioxide (CO2), one of these environmental pollutants, is a key exhaust gas in the cement industry and a major
global source of CO2. Approximately nine hundred kg of CO2 equivalent gases are produced for every ton of
cement produced, which undoubtedly contributes to additional environmental damage24. White cement is
a specic variety of ordinary Portland cement (OPC). While ordinary cement is frequently employed, white
cement oers unique challenges and opportunities for advancement owing to its exceptional brightness and
color uniformity. Consequently, selecting white cement enables the development of advanced materials tailored
for specic applications that necessitate functional and aesthetic performance, particularly in architectural
contexts such as decorative concrete, terrazzo ooring, precast panels, and other scenarios where appearance
is paramount. Recycling solid wastes in the white cement manufacturing process was rare due to their high
sensitivity to replacement or addition to maintain their whiteness level, chemical and physical properties, etc5.
In the fresh case, the rheological functionality of cement paste plays an important role in determining the
structure of concrete and enhances the mechanical properties and hydration characteristics of hardened white
ordinary Portland cement (WOPC)6. erefore, for a more sustainable and eco-friendly option, scientists have
investigated the possibility of using agricultural waste ash as a partial cement substitute in concrete mixes7,8.
Rice production occupies around 11% of the world’s cultivable land, amounting to 145million hectares8.
Rice crop cultivation results in signicant amounts of agro-waste, whereas Egypt is estimated to produce
approximately 4,968,000 tons of rice straw annually as agricultural waste. Rice waste ash (RWA) can be used as
1Environmental Science and Industrial Development Department, Faculty of Postgraduate Studies for
Advanced Sciences, Beni-Suef University, Beni-Suef 62511, Egypt. 2Cellulose and Paper Department,
National Research Centre, 33El-Bohouth St. (Former El- Tahrir St.), Dokki, P.O. 12622, Giza, Egypt. email:
hossamnassarnrc@gmail.com; Ola.abdelaziz.mohamed@psas.bsu.edu.eg
OPEN
Scientic Reports | (2025) 15:2909 1
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports
Content courtesy of Springer Nature, terms of use apply. Rights reserved
a potential substitute for cementitious material in concrete, either by adding it to or replacing a signicant part
of the cement in the mixture. It improves the characteristics of both fresh and hardened concrete by promoting
the interaction between pozzolanic ingredients and Ca(OH)2, leading to the creation of more CSH gel. e
procedure described decreases the amount of calcium hydroxide, enhances the compressive strength, decreases
the permeability, and enhances the workability of the concrete9. Additionally, it mitigates plastic dehydration
and shrinkage, boosts the adhesion area between the cement matrix and aggregate, minimizes the permeability
of water and chloride, and forties the concrete’s resistance to chemical assault. As a result, it reduces the
occurrence and advancement of fractures in concrete8,10. erefore, RSA provides substantial advantages in the
eld of concrete manufacturing.
RWA-cement is less expensive than OPC, as the last one uses limestone, coal resources, and electric
power11. Furthermore, cement pastes containing RWA produce improvements in mechanical properties1214.
e temperature and duration of combustion have a great impact on the RWA’s activity. For example, we can
obtain the greatest activity of rice husks by burning it for 2h at between 600 °C and 800°C15,16. However,
the other factors that aect the RWA-cement pastes properties are unknown. e replacement ratios of RWA
admixture had a good inuence on the cement and concrete performance17,18. Considerable research in this area
conrms that the ideal replacement ratio of every admixture in various building materials is not the same. RWA
includes K2O with a high percentage, which may behave as the “activator” through the combustion process.
e treatment of rice wastes with acid prior to combustion could manage the removal of these impurities to get
an RWA with more and a larger surface area that causes higher activity19. Moreover, research showed that the
chemical contents of an RWA dier with the source or the origin due to geographical, geological, and climatic
conditions20.
Nanoindentation has been the tool of choice for numerous studies over the past 20 years to glean useful
information regarding the microstructural mechanical characteristics of cementitious composites21.
Nanoparticles (NPs) are used in a wide range of ways to develop new environmentally friendly technologies for
the industry of cement, which is the economical experience and the leading of the hour. Signicantly, these novel
procedures possess the capacity to reduce environmental pollution as compared to traditional manufacturing
methods substantially2227.
Nano-silica (NS), as ller, is a highly eective pozzolanic material. ey are very ne and glassy particles,
about a thousand times smaller than the average cement particles. It forms a premium mixture with cement
to make the quality of the cementitious complex or compounds get better to a great extent. Cement pastes
containing NS recorded rapid and accelerated hydration; this may be due to the chemical reactivity (pozzolanic
reaction) or high surface area28. e pozzolanic reaction causes a reduction in the amount of calcium hydroxide
in the concrete mixture and reduces the hydration time, water absorption, porosity, and permeability. According
to earlier studies, adding NS improves the properties of both the fresh and hardened material more than adding
other mineral additions. Using NS with cement pastes improves the cohesive force of the entire mixture in
its fresh condition, lowers the amount of water needed, and shortens setting periods compared to other silica
compounds like silica fume. Because particle agglomeration during mixing causes problems, scientists and
authors estimate that the ideal eective proportion of NS should be less than 5% by weight. According to some
scientists, the right ratio is 10% by weight, depending on the modications that must be made to the components
to prevent excessive self-drying and strength impairment from microcracks29,30. Adding or replacing a part
of the cementitious materials of WC with NS focuses on developing its unique characteristics and reduces its
manufacturing cost and environmental drawbacks31.
e mesoporous silicas nanoparticle (MS-NPs) surface measures over 900g/m2 32. e inuence of NS with
mesoporous MCM-41 is reviewed on plain concrete’s compressive33. Previous research examined the impact
of the method of incorporating MS-NPs into cement mortar. It was found that adding these nanospheres to
an acetone solution along with cement works better and gives the concrete higher compressive and exural
strengths than putting them into a water solution with cement34. For all these reasons, a green environment can
be obtained besides enhancing the properties of WOPC. Consequently, in order for the green environment to
accomplish its sustainability goals, it was necessary to seek out new methods and create them in order to recycle
a substantial quantity of rice waste. e new process of recycling agricultural waste allows for the evaluation of
WOPC pastes’ qualities without compromising their sensitivity.
e purpose of this research is to examine the physico-chemical properties of WOPC that contain
mesoporous silica nanoparticles (MS-NPs). So, dierent MS-NP ratios were tested to see how they aected
the mechanical and hydration features of hardened WOPC. For every cement paste, its water consistency was
recorded, and its rst and last setting times were measured. In addition, up to a 28-day curing period, the
compressive strength, chemically mixed water, and free lime values of the hardened cement specimens were
recorded. Certain hardened samples were examined using XRD, TGA/DTG, and SEM techniques to track the
hydration rate of the prepared materials.
Experimental program
Resources of materials and their characterizations
e rice straws were from Delta, Egypt, which has a moderate temperature, sunny climate, and sucient
humidity. e chemical composition of the rice straw residues was established using TAPPI standard procedures.
e results showed the following: silica 14% (T 245 om-94), ash 15.09% (T-211), lignin 15.09% (T-222 om-88),
extractives 3.95% (T 204cm-97), hemicellulose 21.35% (T-223cm-84), and α-cellulose 42.12% (T 203cm-99).
WOPC (52.5 grade, Blaine specic surface 448 m2 /kg, containing 5% limestone powder) is used in this study.
e whole chemical analysis of WOPC was performed by X-ray uorescence (XRF: Xios, style PW-1400) and the
chemical composition of WOPC was determined and tabulated in Table1.
Scientic Reports | (2025) 15:2909 2
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Fabrication of eco-friendly mesoporous silica nanoparticles (MS-NPs)
In this study, SiO2NPs with controlled shape, phase, and purity were produced using rice straw residues as a
natural and aordable bio-source. e procedure used to accomplish this was as follows: e milled RS sample
was rst leached with 0.5M HCL at a consistency of 10% (wt/v) and 80°C for 30min. e resulting slurry was
then ltered, cleaned with tap water, and allowed to dry at 50°C overnight. Leached RS was calcined twice, the
rst time at 350°C for 45min and the second time at 700°C for an additional 180min. Aer that, silica was
extracted from RS ash by dissolving it in 2M sodium hydroxide (10% wt/v) at 100°C for four hours (Eq.1) while
vigorously stirring the mixture in a closed system. To get rid of carbon leovers that hadn’t been burned, the
resulting sodium silicate solution was ltered. Following its neutralization with a 2M hydrochloric acid aqueous
solution, it underwent another precipitation (Eq.2) 35.
2NaOH (aq)+2SiO2(s)Na2SiO3(aq)+H2O
(1)
Na2SiO3(aq)+2HCl SiO2(gel)+2NaCl (aq)+H2O(l)
(2)
e reaction between sodium silicate and hydrochloric acid facilitates the generation and merging of silanol
groups (R3Si-OH), creating a three-dimensional Si-O-Si network36,37. Silica ash produced was 140g per 1000g
of rice straw residue (14% yields). Treating it with alkali using the previously indicated process formed 100g
of MS-NPs (71.5% yields). e products were then evaluated utilizing a range of analytical tools, such as XRD,
EDX, FTIR, TEM, TG, and N2 physisorption isotherm studies.
Cement paste composite preparation and testing protocols
WOPC was replaced by 0, 0.25, 0.50, 0.75, and 1.0% of MS-NPs, as shown in Table2. e water-to-cement
ratio (W/C ratio) was adjusted according to the weight% of MS-NPs combined with cement pastes. To avoid
MS-NP coagulation, pre-dispersion of MS-NPs is necessary prior to cement mixing. erefore, the MS-NPs are
dissolved in the entire mixing water, then dispersed by using the ultra-sonication technique for 20min, aer
which the emulsion distributed in a multi-speed blender was mixed with cement to form cement paste, followed
by pouring the pastes into cubic molds having the following dimensions of 2.5 × 2.5 × 2.5cm3, then kept at high
relative humidity (nearly 100% RH) for 24h. Aer that, the samples were demolding and immersed under tap
water at normal temperature for 28 days of curing time38.
In accordance with industry standards (ES 2421-3; UNI EN 196-3), setting time and standard water
consistency for these pastes were examined. Using a VICAT device to measure the setting time39. According a
ASTM C109M (2016), the compressive strength (CS) was determined by using Ton industrie instrument (West
Germany) at dierent curing time intervals of 1, 3, 7, and 28 days40. e hydration reaction was stopped using
stopping (1:1 acetone - methanol) mixture with stirring for 1h, then ltered and dried at 75°C for 3h. Aer that,
the stopped samples were maintained in desiccators41.
By heating the ground dried specimens at a temperature of 105°C to remove any free water to prepare the
sample for measuring the chemically combined water content, then for one hour at 1000°C, the chemically
combined water content (Wn, %) was calculated from the ignition loss. For every specimen, duplicate
measurements were made, and the average value was noted. Wn, %= [(W0 – Wf) / Wf] × 100; where W0 = dried
specimens mass and Wf =ignited specimens mass.
Chemically combined water (Wn, %) and free lime (CaO, %) contents were calculated42.
Aer 28 days of curing time, a re resistance test was performed in which the hardened composite pastes
were maintained at approximately 75–80°C for 1day and then exposed to ring at 250, 500 and 750°C for 3h.
e cooling process was applied to the red samples in Two dierent ways (gradually and rapidly cooling)43.
Mixes
Mix proportion
(Wt, %)
W/C (%) Initial setting time (min) Final setting time (min)WOPC MS-NP
Mix WMS0 100 - 0.29 125 180
Mix WMS1 100 0.25 0.292 95 140
Mix WMS2 100 0.5 0.293 105 150
Mix WMS3 100 0.75 0.295 115 160
Mix WMS4 100 1 0.296 125 175
Tab le 2. Optimum water of consistency and setting time for the WOPC-MS hardened composites.
SiO Components Cl SO3 K2O MgO Na2O CaO Fe2O3 Al2O3
23.3 WOPC 0.04 3.43 0.06 0.29 0.12 67.2 0.15 2.38
Tab le 1. Chemical oxide compositions of WOPC.
Scientic Reports | (2025) 15:2909 3
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Characterization
A JEOL JEM-2100 transmission electron microscopy (TEM) system was used to conduct high-resolution TEM
investigations on the produced MS-NPs. e elemental analysis of the extracted MS-NPs was conducted using
an Energy Dispersive X-ray (EDX) spectrometer on a Quanta FEG-250 microscope, operating at a voltage of
10kV.
e surface area and pore size of the MS-NPs were assessed using a Quantachrome (USA; Nova 2000 series)
device employing the Brunauer-Emmett-Teller (BET) model. e samples were subjected to overnight heating at
a temperature of 150°C in order to remove any trapped gases. e physisorption isotherm at 77K was conducted
using nitrogen gas. e FT-IR spectrum of the MS-NPs was recorded using a JASCO FTIR 6100 spectrometer.
e spectrometer had a resolution of 4cm− 1 and a range of 4000 –400cm− 1, and 60 scans were performed.
e phase purity, composition, and crystallinity of the dierent produced samples were determined by
X-ray powder diraction (XRD) in the 2θ range of 5°–80°. e Panalytical Empyrean X-ray diractometer was
employed for this purpose.
e cement and the MS-NPS samples that were made were both tested for thermal degradation using a
dierential scanning calorimeter (SDTQ600 V20.9 Build 20). e measurements were taken in an inert nitrogen
environment at a testing rate of 20°C/minute, with temperatures ranging from 25 to 800°C, in order to avoid
degradation occurring too soon. Equation3 was utilized in conjunction with the Coats-Redfern approach to
ascertain the energy of activation (Ea)44:
log log
Wf/
(
WfW
)
T2
= log
[
AR
θEa
(
1
2
RT
Ea
)]
Ea
2.303RT (3)
Here, R is the gas constant, θ is the heating rate, and W and Wf are the sample’s actual and nal weights (in grams)
up to temperature T (in Kelvin). Equation(4) were used to compute the additional kinetic parameters: e free
energy change of activation (ΔGa), the entropy of activation (ΔSa), and the enthalpy of activation (ΔHa)45.
Ha=Ea
RT;∆Sa=2.303R
(
log
AC
KbT)
,and Ga=∆Ha
TSa (4)
Here, Kb and C are Boltzmann and Planck constants, respectively.
Results and discussion
Extraction and characterization of MS-NPs
e N2 physisorption isotherm of the manufactured MS-NPs sample is depicted in Fig.1a. Physisorption pores
can be conveniently classied based on their size, following the guidelines of the International Union of Pure
and Applied Chemists (IUPAC). Micropores are dened as pores with a width of 2 nm or less, mesopores are
pores with a width between 2 nm and 50nm, and macropores are pores with a diameter of 50nm or greater46.
Given that the average pore width of the manufactured SiO2NP sample is more than 6nm, it can be classied
as a mesoporous material. e mesoporous materials’ type IV isotherm was visible in the produced MS-NPs47.
MS-NPs’ adsorption and desorption assays revealed signicant BET surface areas (187 m2/g), total pore volumes
(0.29 cm3/g), and average pore diameters (6.14nm), all of which were caused by the presence of inter-particulate
holes between the NPs. e N2 adsorption-desorption isotherm showed an H3 hysteresis loop because there
was no saturation adsorption at a high relative pressure (P/P0). is was because the particles were arranged in
a exible way, like lamina46,48. is hysteresis loop structure in fabricated MS-NPs is a consequence of capillary
condensation with hysteresis, which occurs with mesoporous sorbents with an average pore size of more than
4nm and is controlled by the sorbent texture, according to the IUPAC44,47.
By using EDX analysis, the produced MS-NPs’ purity was identied. According to Fig. 1b, the only
components present in the MS-NPs produced were silica and oxygen.
e synthesized MS-NPs used in this study have a molecular structure that is discernible from FTIR analysis,
as shown in Fig.1c. e isolated MS-NPs’ vibrational bands at about 3436, 2921, 1627, 1552, 1095, 806, and
470cm− 1 included adsorbed moisture and/or amorphous silica and/or silicic acid37,49. Si-OH showed symmetric
and asymmetric vibrational absorption bands at 3436 cm− 1 and 806cm− 1, respectively50,51. e estimated
adsorbed moisture of 4.38 wt% for MS-NPs, as determined by TG analysis, may be the reason for the broadening
of the silanol peak between 3070 and 3750cm− 1 owing to the H-bonded silanol groups. At around 1627 and
3436cm− 1, respectively, the bending and stretching vibrational peaks related to adsorbed moisture occurred52.
e Si-O-Si groups from the silicate matrix also showed symmetric and asymmetric vibrational absorption
bands at around 1095 and 806cm− 1, respectively. Along with bending vibration at around 470cm[− 1 53,54.
Additionally, TG analysis was carried out to assess the purity of isolated MS-NPs (Fig.1d). In the temperature
range of 40–150°C, TGA thermograms for as-produced MS-NPs only show one dominant pyrolysis phase with
weight loss corresponding to 4.38%. Volatilization of the adsorbed water molecules and complete siloxane linkage
conversion of the remaining silanol groups cause this process55. Another indication that pure mesoporous MS-
NPs were completely manufactured is their thermal stability following this stage of weight reduction56.
MS-NPs were subjected to an XRD examination to show whether the materials were crystalline or amorphous.
According to Fig.1e, the major diraction peak in the retrieved MS-NPs’ diraction proles is indicative of the
amorphous phase of mesoporous MS-NPs that occurs at a Bragg angle of 22°54. As a result, the XRD prole
shows that the amorphous silica was successfully extracted.
Due to their consistent mesoporosity, large surface area and pore size, particle size (10–1000nm), tunable
pore diameter (2–30 nm), surface functionalization, exible morphology, superior biodegradability, and
Scientic Reports | (2025) 15:2909 4
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
biocompatibility, MS-NPs have attracted a lot of attention5759. Using TEM, the shape of the produced mesoporous
MS-NPs derived from burned RS was investigated. Figure2: Single uneven chunks with a width varying from
160 to 900nm are visible in TEM micrographs. ese masses of MS-NPs are homogeneous, spherical, and
somewhat regular in their holes and voids, which range in diameter from 9.11 to 5.53nm, much like mesoporous
structures. In line with the N2 physisorption isotherm data, the TEM data show that the mesoporous MS-NPs
that were successfully made were in fact periodic matrices consisting of medium-sized pores contained inside
an amorphous silica matrix.
Physicomechanical aspects
Setting characteristics of white cement paste
Standard water of consistency and Setting time for the WOPC-MS hardened composites are shown in Table2;
Fig.3, shows the both initial and nal setting times of WOPC-MS hardened composites as a function of time.
It can be observed that the values of standard water consistency of all admixed cement pastes are slightly higher
than that of WOPC blank (Mix WM0), as more water consistency is required in order to high dispersion of the
MS, which causes an increasing in w/c ratio and speeding up the rate of hydration of cement60. Also, the obtained
results indicate that the addition of MS for cement pastes results in a notable reduction in time taken to set, so
that all setting time values obtained for WOPC-MS hardened composites are shorter than those of the blank Mix
(WM0). e shortest setting time behavior of MS may be attributed to the nucleation eect, the combination
of pozzolanic activity, and the lling eect of MS during the hydration process38,61,62. It was noted that as the
amount of addition of MS increases, both the initial and nal setting times increase owing to the compactness
and accumulation of the small nanoparticle of MS, which would have reduced the penetration rate of water and
the rate of hydration reaction, and consequently increased the setting time. Mix (WM1) has the quickest setting
times when compared to other WOPC composite cement blends containing MS where MS aids as an active
crystal site for the generation of an excess of calcium aluminate hydrate (CAH), calcium silicate hydrate (CSH),
and gel calcium aluminosilicate (CASH) and further CSH stages subsequence on.
Compressive strength (CS)
Figure4 represents the results of the CS values of the hardened neat WOPC pastes, as well as WOPC pastes
modied with various additions (0.25, 0.50, 0.75 and 1%) of (MS) of 6.14nm average particle size. As can be seen
Fig. 1. (a) N2-adsorption/desorption isotherm, (b) EDX analysis, (c) FTIR spectra, (d) TGA thermograms,
and (e) XRD patterns MS-NPs.
Scientic Reports | (2025) 15:2909 5
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Fig. 3. Setting times of the neat WOPC (Mix WMS0) and WOPC-MS composites.
Fig. 2. HR-TEM of MS-NPs.
Scientic Reports | (2025) 15:2909 6
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
clearly from Fig.4, the values of compressive mechanical strength showed a continuous overall development
with the increase of the hydration time for all hardened MS-white ordinary Portland cement pastes. In general,
the hydration of dierent phases in WOPC clinker and the subsequent formation of hydration products such as
ettringite (C63H32), calcium aluminosilicatehydrate (C(A)SH), alumino ferrrite monosulfate hydrate (AFm),
calcium aluminate hydrates (CAH), and calcium silicate hydrates (CSH) are the main causes of the ongoing rise
in strength values. Adding MS improved the strength of the WOPC-MS hardened composites. MS signicantly
inuences the early hydration process (up to 7 days) by enhancing the matrix structure of the WOPC-MS-
hardened composites. Additionally, it contributes to a slower increase in CS values later on by protecting the
un-hydrated parts of cement grains with the initial hydration products, leading to a gradual rate of hydration42.
e WOPC-MS hardened composites with 0.25%, 0.50%, 0.75%, and 1.0% of MS-NPs displayed a similar trend
to the blank (Mix WMS0), but exhibited signicantly higher values of CS, particularly in the initial stages of
hydration (up to 7 days). It is evident that the WMS3 mix, which contains 0.75% MS, demonstrated the highest
strength value aer 28 days compared to the other mixes (WMS0, WMS1, WMS2, and WMS4). e WOPC-
MS composites were strengthened with varying percentages of MS-NPs: 0.25%, 0.50%, 0.75%, and 1.0%. is
improvement comes from the lling eect of integrated mesoporous silica (MS), and the cement matrix becomes
less porous. Furthermore, the results indicated that the extent of reduction in the CS values was heightened as
the percentage of MS added to WOPC rose. is could disrupt the composite microstructure.
Cement pastes hydration
Chemically combined water content (wn)
Figure5 demonstrate the results of Wn for the WOPC-MS hardened composites pastes. Evidently, the Wn
- values obtained for all the tested hardened samples indicate a gradual continuous increase up to the nal
hydration ages studied (28 days). e Wn values obtained for all the evaluated hardened samples demonstrate
a steady and continuous rise until the ultimate hydration age is examined (28 days). e observed outcome is
attributed to the advancement of the hydration process and the spread of calcium silicate hydrates (CSH) and
calcium hydroxide (Ca(OH)2) as essential byproducts of hydration63. e cement pastes composed of Mixes
(WMS1 – WMS4) exhibited signicantly greater values of Wn compared to plain WOPC at all hydration stages.
Furthermore, Mix WMS3 (containing 0.75% MS) resulted in the greatest Wn values among all tested specimens
aer 28 days compared to those of the other mixes (WMS0, WMS1, WMS2 and WMS4). is nding is owing
Fig. 4. Compressive strength values for hardened composites made from WOPC-MS at dierent ages of
hydration.
Scientic Reports | (2025) 15:2909 7
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
to the advancement eect of MS resulting in acceleration of cement hydration process with a stronger and denser
microstructure via the generation of further hydration products like C(A)SH, C-A-H, and C-S-H that deposit
in the pore structure.
Free lime content (FL)
Figure6 presents FL content (CaO, %) for neat WOPC (mix WMS0) and WOPC-MS pastes modied with
various amounts of MS (WMS1, WMS2, WMS3, and WMS4 composites) cured up to 28 days. e free lime
content (FL) increases with increasing curing age; this may be owing to the progress of the hydration reaction,
which liberates free portlandite CH through the hydration time64. e ndings also demonstrate a reduction
in the values of the free lime content (FL %) for the WOPC pastes containing MS nanoparticles during all ages
of hydration; and the reduction in FL contents (CaO, %) is notable with increasing the % of MS. ere are two
opposite processes for all pastes containing MS, the rst one is a slight increase in FL content (CaO, %) due to
liberated free portlandite from cement hydration up to 7 days and the second process is a marked decrease in
FL content (CaO, %) up to 28 days. is decrease in (FL %) values is attributed to MS nanoparticle’s pozzolanic
activity; hence MS can minimize Ca(OH)2 and aid the forming of a denser system containing C-F-H and C-F-
S-H gels, besides further amount of C-S-H.
Thermal degradation (TD)
Figure7a presents the inuence of higher temperatures (250, 500, and 750°C) for 3h. en, allow them to
cool gradually in the ambient air, changing the CS values for plain WOPC and WOPC-MS composite cement
pastes. e data in the gure shows a signicant increase along thermal treatment upon heating at 250°C in
CS values for all red composites compared to their recorded values aer 28 days of hydration, followed by a
slight decrease upon heating at 500°C, then a sharp reduction for all composites up to 750°C. Really, the severe
boost in CS aer ring at 250°C could be attributed to the internal (self-autoclaving) process created from the
elimination of the physically adhered water molecules in cement structure. Composite pastes, WOPC-MS, were
enhanced with several amounts (0.25, 0.5, 0.75, and 1 mass %) of MS-NPs. ese modied pastes exhibited
higher CS values aer being red at 250°C compared to the CS values of the plain composite (Mix WMS0).
Aer being red to 250, 500, and 750°C for 3h and then rapidly cooled (using tap water), Fig.7b shows the CS
magnitudes for several composites. Similarly, due to the thermal shock experienced by the red samples during
the quick cooling processes, the CS values for all composites signicantly decrease as the exposure temperature
Fig. 5. Chemically combined water contents for hardened composites made from WOPC-MS at dierent ages
of hydration.
Scientic Reports | (2025) 15:2909 8
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
increases from 250 to 750°C. In addition, cracks may appear later on in the red samples. However, compared
to ordinary WOPC, the drop in CS values for all composites incorporating MS-NPs is lower.
e relative compressive strength percentages (RCS%) for composites WOPC-MS0.25, WOPC-MS0.50,
WOPC-MS0.75, and WOPC-MS1.0 that were cooled gradually in the air were, as shown in Figs.8a and b and
115.4, 116.3, 116.9, and 104.7%, respectively, relative to plain WOPC. is suggests that 0.75 mass% MS-NPs
Fig. 7. Compressive strength values for hardened composites made from WOPC-MS red at dierent
temperatures and cooled (a) gradually in air, (b) suddenly in water.
Fig. 6. Free lime contents of neat WOPC (Mix WMS0) and WOPC-MS hardened composites at dierent ages
of hydration.
Scientic Reports | (2025) 15:2909 9
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
is the ideal incorporation of MS-NPs for enhancing TD at 250°C. is development in TD at 250°C for OPC-
0.75% MS-NPs composite paste is related to divers factors like; good dispersion of MS0.75 NPs (6.14nm and
SBET = 187m2/g) through the composite matrix and credible catalytic activity of MS-NPs to create a massive
amount of hydration products such as (C-S-Hs I and II), C-A-S-H, C-A-H, C-F-H, CH and C-F-S-H) that
precipitate in the voids of the cement matrix and generate hardened matrix that possess perfect resistance to
re decay. Obviously, upon exposure to 500°C, the CS magnitudes were slightly reduced for all composite
pastes (Mixes WMS0, WMS1, WMS2, WMS3 and WMS4). e descent in CS magnitudes is mainly attributed
to the deterioration of most of the hydration products (CSHI, CSHII, CASHs, CAHs, CFH, CH, and CFSH.
yet, all composite pastes containing MS-NPs have lower reduction in CS values if compared with that of plain
composite. the percentages of RCS aer ring at 500°C for Mixes WMS0, WMS1, WMS2, WMS3 and WMS4
were 101.2,112.1, 113.7, 114.6 and 103.1%; supported that WOPC-MS composite pastes possess higher TD if
compared with that of plain (mix WMS0).
e massive reduction in CS magnitudes seen in all tested composites aer being red to 750°C can be
attributed to the complete thermal degradation of all binding centers. As a result, it generates many fractures
inside the composite matrix. e percentages of RCS at 750°C for mixes WMS1, WMS2, WMS3 and WMS4
were 16.6, 16.9, 17.0 and 16.7%; supported that WOPC-MS composite pastes possess higher TD if compared
with that of plain (mix WMS0) at high temperature. Mix WMS3 (WOPC-0.75% MS-NPs composite) acquires
the highest compression magnitudes and the eective TD during treating with thermally temperature up to
750°C, if it compared with plain WOPC.
e previously reported results exposed that all the hardened composites modied with diverse amounts of
MS-NPs oered perfect TD (re impedance) compared to other composites (plain WOPC and WOPC-MS),
specially (WOPC-0.75 MS-NPs), which show the opportunity to utilize of these composites in re impedance
application.
Phase composition
Dierential ermo-gravimetric analysis (TGA/DTG)
In order to thoroughly analyze the thermal stability of the hardened neat WOPC and hardened WOPC-MS pastes
with 0.75% MS, their thermal properties were examined aer 7 and 28 days of hydration. Weight loss (TG) and
rst derivative of weight loss (DTG) curves for hardened WOPC in combination with and without MS-NPs are
shown in Fig.9a–d. Table3 shows how much weight each hardened paste lost at dierent temperature ranges. It
also shows the kinetic parameter data from the Coats-Redfern plots for each step of the decomposition process.
e thermal characterization graphs presented in Fig.9a–d clearly demonstrate that each curve comprises three
distinct stages of thermal degradation;
(I) e initial phase within the temperature range of 24–386°C results in a reduction in weight due to the
evaporation of hydrated products resulting from the decomposition of CSAHs, CSHs, and CAHs. e
analysis of the hydrate content in the solidied neat WOPC and WOPC-MS pastes reveals that the in-
clusion of MS-NPs led to a greater bonding of water molecules, mostly due to the enhanced hydrophilic
properties of the MS-NPs.
(II) e second phase, occurring at temperatures between 473 and 609°C, results from pyrolysis. It is caused by
the dehydroxylation of Ca(OH)2. e estimated mass loss for hardened neat WOPC pastes (Mix WMS0) is
4.324%, and 4.399%, at 7 days and 28 days of hydration. For hardened WOPC-MS pastes containing 0.75%
MS, the estimated mass loss is 3.668%, and 4.647% at 7 days and 28 days of hydration, respectively. e
outcomes show the distinct hardened pathway that WOPC followed in the presence of MS-NPs.
Fig. 8. Relative residual strength for hardened composites made from WOPC-MS red at dierent
temperatures and cooled (a) gradually in air, (b) suddenly in water.
Scientic Reports | (2025) 15:2909 10
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
(III) e third phase is the pyrolysis of free lime and carbonated segments, which ends at 869, 929, 805, and
806°C. e weight loss for hardened neat WOPC pastes (Mix WMS0) at 7 days and 28 days of hydration
and for hardened WOPC-MS pastes containing 0.75% MS at 7 days and 28 days of hydration is 5.414, 6.98,
5.635, and 3.898%, respectively. Following the combustion process, the residual weight for hardened plain
Stage Temp. range, °C Tonset, °C Tmax, °C Wloss, % Rw1000, °C R2A, 1/s Ea, kJ/mol ∆Ha, kJ/mol ∆Sa, J/mol.K ∆Ga, kJ/mol
Mix WMS0-7d
I 32–380 99 175 9.087
80.95
0.93 10.60 25.23 21.50 -228.725 123.97
II 400–609 463 526 4.324 0.98 2.94 × 1010 183.03 176.38 -52.7186 218.50
III 629–879 629 775 5.414 0.99 1.21 × 1007 171.40 162.69 -119.772 288.21
∑ 379.66 ∑ 360.57 ∑ -401.22 ∑ 630.69
Mix WMS3-7d
I 24–386 100 127 8.87
80.60
0.93 2.99 19.67 16.36 -238.297 111.68
II 472–576 472 509 3.668 0.98 2.61 × 1013 221.74 215.23 3.932746 212.16
III 576–805 618 737 5.635 0.99 2.57 × 1008 188.80 180.40 -94.0709 275.41
∑ 430.22 ∑ 412.00 ∑ -328.44 ∑ 599.25
Mix WMS0-28d
I 24–284 98 150 5.776
80.71
0.91 5.70 20.59 17.07 -233.408 115.80
II 492–596 483 520 4.399 0.99 1.15 × 1012 204.56 197.97 -22.187 215.56
III 596–929 608 753 6.98 0.98 271578.4 141.10 132.57 -151.196 287.69
∑ 366.25 ∑ 347.61 ∑ -406.79 ∑ 619.06
Mix WMS3-28d
I 24–369 100 157 9.734
81.36
0.97 9.37 24.33 20.76 -229.406 119.40
II 473–577 469 518 4.647 0.97 1.41 × 1013 219.98 213.40 -1.2992 214.43
III 577–806 601 744 3.898 0.97 4.86 × 1016 353.53 345.08 64.34119 279.64
∑ 597.84 ∑ 579.23 ∑ -166.36 ∑ 613.47
Tab le 3. ermo-analytical data of hardened neat WOPC and hardened WOPC-MS pastes containing 0.75%
MS aer 7 and 28 days of hydration.
Fig. 9. TG/DTG thermograms of hardened neat WOPC pastes (Mix WMS0) at (a) 7 days, (c) 28 days of
hydration and hardened WOPC-MS pastes containing 0.75% MS (Mix WMS3) at (b) 7 days and (d) 28 days of
hydration.
Scientic Reports | (2025) 15:2909 11
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
WOPC pastes (Mix WMS0) at 7 days and 28 days of hydration and hardened WOPC-MS pastes contain-
ing 0.75% MS at 7 days and 28 days of hydration, respectively, equals 80.95, 80.71, 80.60, and 81.36%.
e pyrolysis mechanism and thermal stability of hardened neat WOPC and hardened WOPC-MS pastes can
be approximated from these collected data. e WOPC under investigation, both with and without MS-NPs,
is stable at room temperature and deteriorates progressively when heated. ese rigid pastes can be thermally
broken down into two stages: pyrolysis and dehydration. e dehydration operation of all hardened pastes
occurred in one stage. Aer the water molecules are eliminated, two phases of the thermal breakdown of the
dried pastes take place.
On the other hand, the DTG curve of the maximum weight loss rate (RWloss, %/minute) can be used in two
ways to talk about the default stability for hardened pastes: e term “thermal stability” can refer to either the
energy of activation (Ea) for decomposition reactions or the temperature values at which the rate of weight loss
from thermal degradation is greatest (Tmax, °C). e former is more commonly used to describe kinetic stability.
e starting temperatures of thermal deterioration (Tonset, °C) are another way to describe it. It is evident that as
the hardening time was increased from 7 to 28 days, the thermal stability of the neat WOPC became less stable,
based on the Tmax of the second and third degradation stages, which may be regarded as one thermal stability
factor. On the other hand, by lengthening the hardening period from 7 to 28 days, hardened WOPC-MS pastes
have improved thermal stability.
Based on the kinetic data in Table3, the activation energies for the initial decomposition reaction (I) in
hardened pastes are arranged in the following order: Mix WMS3-7d < Mix WMS0-28d < Mix WMS3-28d < Mix
WMS0-7d. Furthermore, the order of thermal stability between the stages of thermal degradation can be
described as follows: II > III > I for hardened pastes of Mix WMS0-7d, Mix WMS0-28d, and Mix WMS3-7d.
e combustion stage of the hardened WOPC-MS pastes, on the other hand, exhibits the maximum activation
energy aer 28 days of hydration, indicating that more energy is needed to start the thermal decomposition
process at the specied Tmax. e activation energy and the enthalpy of activation follow the same order: WMS3-
28d > mix WMS0-7d > WMS3-7d > WMS0-28d. In contrast, the activation entropies are in the following order:
mix WMS3-28d < mix WMS0-7d < mix WMS3-7d < mix < WMS0-28d. is indicates that by adding MS-NPs
to the WOPC paste, the activation entropy decreased and eventually reached a minimum aer 28 days of
dehydration.
X-ray diraction (XRD)
Figure10 (i) presents the XRD patterns of composites consisting of plain WOPC at dierent time intervals
(7 and 28 days) and aer being subjected to dierent thermal temperatures (250 and 750°C) aer 28 days of
hydration. e diagram displays distinct peaks representing various hydration yields, including C-S-H and C-H,
the primary hydration products. Additionally, there are indicative peaks of some unreacted portions of β-C2S
and C3S. Furthermore, the peaks featured for the calcite (calcium carbonate; CaCO3) at nearly 2Ɵ of 29.32[o 65
are the results of the reaction of CO2 gas with the formed lime (CH). In addition to crystalline silica (quartz) is
detected66. XRD patterns of the sample red at 250 °C shows all phases mentioned earlier as well as a new phase
of CASH formed as a result of the self (internal) autoclaving phenomena of unreacted fractions; and this result
is in match with the ndings of the compressive strength (CS). e XRD pattern demonstrates a noticeable
decrease in the intensity of the prominent peaks of CSH, CH, and CASH due to the thermal degradation of
Fig. 10. XRD-patterns for (i) hardened neat WOPC pastes (Mix WMS0) and (ii) hardened WOPC-MS pastes
containing 0.75% MS (Mix WMS3).
Scientic Reports | (2025) 15:2909 12
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
hydration products at an elevated temperature of 750°C. Furthermore, the calcite peaks, composed of calcium
carbonate (CaCO3), vanished at 750°C. is disappearance occurred due to the disintegration of both calcium
hydroxide (CH) and calcium carbonate (CaCO3), resulting in the formation of calcium oxide (CaO)67.
Figure10(ii) displays the XRD patterns of composite made of WOPC-0.75% MS-NPs at diverse times intervals
(7 and 28 days) and aer ring at various thermal temperatures (250 and 750°C) at 28 days of hydration. It is
legibly that all above-mentioned phases are existent like CSHs (as wollastonite and Clinozoisite phases), Calcite,
alite, belite, and CH. Furthermore, as the hydration progressed, the diraction lines corresponding to CSHs and
CFH phases increased with hydration age up to 28 days68,69.
In the Fig.10(ii) as a result of the catalytic activity of 0.75% MS-NPs, new peaks appeared corresponding to
extra phases are seen such as the peaks of CASHs phase and ilavite (CFSH). In addition, the calcite and CH peaks
were not as strong because of the pozzolanic interaction of 0.75% MS-NPs with CH. At a temperature of 250°C,
there was a signicant rise in the intensity of CSH, CH, and CASH peaks in the specimen. is increase can be
attributed to the internal reaction (self-autoclaving) of unreacted phases, namely alite and belite. Whereas, for
the specimen exposed for 750°C, a signicant decline in the intensity of CSH, CH, CFSH and CASH peaks has
been detected as a result of the thermal decadence of such hydration yields69.
Textural and morphological features
Figure11a and b display the microstructure and morphology of the hardened, neat WOPC pastes following 7
and 28 days of hydration. e SEM micrograph aer 7 days reveal the formation of small amounts of (C-S-H;
tobermorite-like) along with the creation of small hexagonal portlandite (CH) crystals. In addition to a
considerable amount of unreacted particles can be well detected as shown in Fig.11a39. A dense matrix was
acquired aer 28 days of curing time as clear in Fig.11b, composed of massive amounts of needle crystals and
brous C-S-H combined with portlandite (CH) which appear as accumulated hexagonal crystals. also, as a result
of the side reaction of atmospheric CO2 gas with CH created through handling of the specimens, a tiny amount
of calcite (CaCO3) as cubic plates crystals can be observed in the cementitious matrix. Moreover, the presence
of vacuum in the cementitious matrix for the sedimentation of other new phases of hydration70. SEM images as
illustrated in Fig.11c for thermally treated WOPC pastes at 250°C aer 28 days of hydration shows the presence
of a compact structure in the form of sheets, ill-crystallized of (CSH) and plates of CASHs that overlapped with
Fig. 11. SEM images of WOPC-hardened cement pastes (Mix WMS0) at 7 days (a); at 28 days (b); aer ring
at 250°C (c); and aer ring at 750°C (d).
Scientic Reports | (2025) 15:2909 13
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
the hexagonal plates of portlandite (CH) liberated from WOPC hydration. e attained microstructure at 250°C
arm the internal (self-autoclaving) reaction of unreacted WOPC particles which cause of the generation of
more amount of hydration phases like (C-S-H, C-H, C-A-H, C-A-S-H), forming a denser structure4. Figure11d
indicates SEM images of thermally treated WOPC pastes at 750°C aer 28 days of hydration showing a whole
thermal degradation for almost hydration yields with dierent micro-cracks appearing43.
Figure12a, b claries the SEM images of hardened composite containing WOPC-0.75 MS-NPs (mix WMS3)
aer 7 and 28 days, respectively. Aer 7 days, the results of the SEM micrographs of this composite are less
dense and include less quantity of hydration yields and a large fraction of unreacted clinker parts if compared to
their SEM micrographs aer 28 days of curing age. Moreover, the hardened composite containing WOPC-0.75
MS-NPs still boost of the mechanical and microstructure properties if compared with that of blank WOPC.
Figure12b, illustrates the presence of excessive hydration products as micro-rods and crystals of brous CSH,
CASHs plates, CFH ne crystals and AFt bers (3CaO.Al2O3.3CaSO4.32H2O); each of them are interweaved
with hexagonal CH crystals. ese bulk phases are responsible for the acquired hard and dense matrix for
WOPC-0.75 MS-NP and this is related to catalytic activity and the pore-lling eect of MS.
SEM images of thermally treated WOPC-0.75 MS-NPs (mix WMS3) pastes at 250°C aer 28 days of
hydration are illustrated in Fig.12c. e SEM micrographs reect a very compact structure if compared to that
of a cement blank (Mix MS0) under the same ring conditions. e micrographs of this composite conrm
the presence of an excessive amount of hydration products like amorphous brous CSH and plates of CAHs,
CASHs beside CFSH ne crystals interweaved with a few hexagonal CH plates resulting in a high density matrix
with minimal porosity, which is ascribed to the internal (self) autoclaving reaction which favorably aects the
mechanical characteristics.
aer exposure to 750°C, Fig.12d revealed the existence of various micro-cracks and the relatively not fully
thermal ruptures for nearly all created phases as a result of still exists small quantity of brous CSH which may
be ascribed to presence of a high alkaline content in WOPC31.
Fig. 12. SEM images of WOPC- MS hardened cement pastes (Mix WMS3) at 7 days (a); at 28 days (b); aer
ring at 250°C (c); and aer ring at 750°C(d).
Scientic Reports | (2025) 15:2909 14
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Conclusion
Positive eect of eco-friendly MS-NPs on the mechanical, physical, re resistivity, and microstructure
characteristics of WOPC pastes is reported in this article. Using eco-friendly MS-NPs was successful due to
essential factors such as amorphousity, nucleating side eects on WOPC, and waste recycling. is type of
WOPC-MS hardened composite has high physico-mechanical properties, low CO2 emission, and low cost,
and it is required in construction projects. e primary inuential factors in the performance of WOPC were
amorphous silicate, color, and particle size.
e results of this investigation can be summarized as follows: -.
1. e results obtained from the experiments proved the suitability of recycled agricultural waste for improving
the properties of WOPC pastes for the preparation of construction materials with outstanding durability and
mechanical characteristics as well as re resistivity.
2. Inclusion of mesoporous silica SiO2 nanoparticles (0.75 mass %) in WOPC pastes motivate the thermal sta-
bility.
3. e composite WMS3 (WOPC–0.75 MS) provides numerous advantages from both an economic and en-
vironmental perspective. Among all the studied nanocomposites, this mix had the greatest CS values, the
shortest setting time compared to plain, and the highest RS% aer ring, making it the best choice for gen-
eral construction purposes.
4. e massive types of hydration products in the WOPC matrix were armed via SEM, XRD, and TG/DTG
techniques, especially aer a 0.75% addition of MS-NP.
Data availability
Availability of data and materials:-All data generated or analyzed during this study are available upon request
from all authors of this paper.
Received: 26 April 2024; Accepted: 14 January 2025
References
1. Shehata, N., Mohamed, O. A., Sayed, E. T., Abdelkareem, M. A. & Olabi, A. G. Geopolymer concrete as green building materials:
Recent applications, sustainable development and circular economy potentials. Sci. Total Environ. 836, 155577 (2022).
2. Feng, P. et al. e signicance of dispersion of nano-SiO2 on early age hydration of cement pastes. Mater. \& Des. 186, 108320
(2020).
3. Mohamed, O. A., Farghali, A. A., Eessaa, A. K. & El-Shamy, A. M. Cost-eective and green additives of pozzolanic material derived
from the waste of alum sludge for successful replacement of Portland cement. Sci. Rep. 12, 20974 (2022).
4. Ramadan, M., El-Gamal, S. M. A. & Selim, F. A. Mechanical properties, radiation mitigation and re resistance of OPC-recycled
glass powder composites containing nanoparticles. Constr. Build. Mater. 251, 118703 (2020).
5. Igami, R. et al. Full-scale observation of drying-induced microstructure change in hardened cement paste by water and 2-propanol
1H NMR relaxometry. Cem. Concr Res. 186, 107698 (2024).
6. Hanus, M. J. & Harris, A. T. Nanotechnology innovations for the construction industry. Prog Mater. Sci. 58, 1056–1102 (2013).
7. Amin, M., Tayeh, B. A., Kandil, M. A., Agwa, I. S. & Abdelmagied, M. F. Eect of rice straw ash and palm leaf ash on the properties
of ultrahigh-performance concrete. Case Stud. Constr. Mater. 17, e01266 (2022).
8. Hakeem, I. Y. et al. Eects of nano sized sesame stalk and rice straw ashes on high-strength concrete properties. J. Clean. Prod. 370,
133542 (2022).
9. Pandey, A. & Kumar, B. Eects of rice straw ash and micro silica on mechanical properties of pavement quality concrete. J. Build.
Eng. 26, 100889 (2019).
10. Agwa, I. S., Omar, O. M., Tayeh, B. A. & Abdelsalam, B. A. Eects of using rice straw and cotton stalk ashes on the properties of
lightweight self-compacting concrete. Constr. Build. Mater. 235, 117541 (2020).
11. Qin, Z. et al. Biomass and biofuels in China: Toward bioenergy resource potentials and their impacts on the environment. Renew.
Sustain. Energy Rev. 82, 2387–2400 (2018).
12. Alex, J., Dhanalakshmi, J. & Ambedkar, B. Experimental investigation on rice husk ash as cement replacement on concrete
production. Constr. Build. Mater. 127, 353–362 (2016).
13. Kang, S. H., Hong, S. G. & Moon, J. e use of rice husk ash as reactive ller in ultra-high performance concrete. Cem. Concr Res.
115, 389–400 (2019).
14. Praveenkumar, T. R., Vijayalakshmi, M. M. & Meddah, M. S. Strengths and durability performances of blended cement concrete
with TiO2 nanoparticles and rice husk ash. Constr. Build. Mater. 217, 343–351 (2019).
15. Zunino, F. & Lopez, M. A methodology for assessing the chemical and physical potential of industrially sourced rice husk ash on
strength development and early-age hydration of cement paste. Constr. Build. Mater. 149, 869–881 (2017).
16. Padhi, R. S., Patra, R. K., Mukharjee, B. B. & Dey, T. Inuence of incorporation of rice husk ash and coarse recycled concrete
aggregates on properties of concrete. Constr. Build. Mater. 173, 289–297 (2018).
17. Shehab, H. K., Eisa, A. S. & Wahba, A. M. Mechanical properties of y ash based Geopolymer concrete with full and partial cement
replacement. Constr. Build. Mater. 126, 560–565 (2016).
18. Bommisetty, J., Keertan, T. S., Ravitheja, A. & Mahendra, K. Eect of waste ceramic tiles as a partial replacement of aggregates in
concrete. Mater. Today Proc. 19, 875–877 (2019).
19. Xu, W. et al. Comparative study of water-leaching and acid-leaching pretreatment on the thermal stability and reactivity of biomass
silica for viability as a pozzolanic additive in cement. Mater. (Basel). 11, 1697 (2018).
20. Khan, W., Shehzada, K., Bibi, T., Islam, S. U. & Khan, S. W. Performance evaluation of Khyber Pakhtunkhwa Rice Husk Ash (RHA)
in improving mechanical behavior of cement. Constr. Build. Mater. 176, 89–102 (2018).
21. Momotaz, H. et al. Properties of the interfacial transition zone in rubberised concrete–an investigation using nano-indentation
and EDS analysis. J. Build. Eng. 77, 107405 (2023).
22. Dheyaaldin, M. H. & Kianmehr, P. Sustainable utilization of nanomaterials in reactive powder composite: state of the Art Review.
J. Build. Eng. 111047 (2024).
23. Shumuye, E. D. et al. Inuence of novel hybrid nanoparticles as a function of admixture on responses of engineered geopolymer
composites: A review. J. Build. Eng. 108782 (2024).
Scientic Reports | (2025) 15:2909 15
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
24. El-Sawy, A. A., Heikal, M., Ibrahim, S. M. & Mohamed, O. A. Experimental and theoretical study on the physico-mechanical
characteristics and radiation shielding capability of hardened alumina sludge waste-cement pastes containing MnFe2O4-
nanoparticles. Discov Appl. Sci. 6, 567 (2024).
25. Mohtasham Moein, M., Rahmati, K., Saradar, A., Moon, J. & Karakouzian, M. A critical review examining the characteristics of
modied concretes with dierent nanomaterials. Mater. (Basel). 17, 409 (2024).
26. Franco-Luján, V. A., Montejo-Alvaro, F., Ramírez-Arellanes, S., Cruz-Martínez, H. & Medina, D. I. Nanomaterial-Reinforced
Portland-Cement-based materials: A review. Nanomaterials 13, 1383 (2023).
27. Chen, Y., Li, X. & Du, H. A review of high temperature properties of cement based composites: Eects of nano materials. Mater.
Today Commun. 35, 105954 (2023).
28. Xu, J., Shen, W., Corr, D. J. & Shah, S. P. Eects of nanosilica on cement grain/C–S--H gel interfacial properties quantied by
modulus mapping and nanoscratch. Mater. Res. Express. 6, 45061 (2019).
29. Sa, B., Aknouche, H., Mechakra, H., Aboutaleb, D. & Bouali, K. Incorporation mode eect of Nano-silica on the rheological and
mechanical properties of cementitious pastes and cement mortars. in IOP Conference Series: Earth and Environmental Science vol.
143 12015 (2018).
30. Setiati, N. R. Eects of additional nanosilica of compressive strength on mortar. in IOP Conference Series: Materials Science and
Engineering vol. 223 12065 (2017).
31. Abdelzaher, M. A. & Shehata, N. Hydration and synergistic features of nanosilica-blended high alkaline white cement pastes
composites. Appl. Nanosci. 12, 1731–1746 (2022).
32. Al-Shemy, M. T. et al. Silver nanoparticles incorporated with superior silica nanoparticles-based rice straw to maximize biogas
production from anaerobic digestion of landll leachate. J. Environ. Manage. 365, 121715 (2024).
33. Firozi, A. & Dost, G. A. Investigation of nano structured mesoporous silica eect on compressive strength properties of normal
concrete. QUID Investig Cienc. y Tecnol 1106–1111 (2017).
34. Horszczaruk, E., Mijowska, E., Cendrowski, K., Mijowska, S. & Sikora, P. Eect of incorporation route on dispersion of mesoporous
silica nanospheres in cement mortar. Constr. Build. Mater. 66, 418–421 (2014).
35. Seroka, N. S., Taziwa, R. T. & Khotseng, L. Extraction and synthesis of silicon nanoparticles (SiNPs) from sugarcane bagasse ash:
A mini-review. Appl. Sci. 12, 2310 (2022).
36. Azat, S., Korobeinyk, A. V., Moustakas, K. & Inglezakis, V. J. Sustainable production of pure silica from rice husk waste in
Kazakhstan. J. Clean. Prod. 217, 352–359 (2019).
37. El-Sakhawy, M., Adel, A. M., Diab, M. A. & Al-Shemy, M. Facile methods for the preparation of micro- and mesoporous amorphous
silica from rice husk. Biomass Convers. Biorenery. 1–10. https://doi.org/10.1007/s13399-020-01112-2 (2020).
38. Zhang, A., Ge, Y., Yang, W., Cai, X. & Du, Y. Comparative study on the eects of nano-SiO2, nano-Fe2O3 and nano-NiO on
hydration and microscopic properties of white cement. Constr. Build. Mater. 228, 116767 (2019).
39. Shafeek, A. M., Khedr, M. H., El-Dek, S. I. & Shehata, N. Inuence of ZnO nanoparticle ratio and size on mechanical properties
and whiteness of White Portland Cement. Appl. Nanosci. 10, 3603–3615 (2020).
40. Hazem, M. M., Hashem, F. S., El-Gamal, S. M. A. & Amin, M. S. Mechanical and microstructure characteristics development of
hardened oil well cement pastes incorporating y ash and silica fume at elevated temperatures. J. Taibah Univ. Sci. 14, 155–167
(2020).
41. Abo-El-EneinS. A., El-HosinyF. I., El-GamalS. M. A., AminM. S. & Ramadandan M. Gamma radiation shielding, re resistance
and physicochemical characteristics of Portland cement pastes modied with synthesized Fe2O3 and ZnO nanoparticles. Constr.
Build. Mater. 173, 687–706 (2018).
42. Abo-El-Enein, S. A., El-Gamal, S. M. A., Aiad, I. A., Azab, M. M. & Mohamed, O. A. Early hydration characteristics of oil well
cement pastes admixed with newly prepared organic admixture. HBRC J. 14, 207–214 (2018).
43. Mohamed, O. A., El-Gamal, S. M. A. & Farghali, A. A. Utilization of alum sludge waste for production of eco-friendly blended
cement. J. Mater. Cycles Waste Manag. 24, 949–970 (2022).
44. Al-shemy, M. T., Al-sayed, A. & Dacrory, S. Fabrication of sodium alginate / graphene oxide / nanocrystalline cellulose scaold for
methylene blue adsorption: Kinetics and thermodynamics study. Sep. Purif. Technol. 290, 120825 (2022).
45. Adel, A. M. et al. Fabrication of packaging paper sheets decorated with alginate/oxidized nanocellulose–silver nanoparticles bio-
nanocomposite. Int. J. Biol. Macromol. 181, 612–620 (2021).
46. Sotomayor, F. J., Cychosz, K. A. & ommes, M. Characterization of Micro/Mesoporous materials by Physisorption: Concepts and
Case studies. Acc. Mater. Surf. Res. 3, 34–50 (2018).
47. ommes, M. et al. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution
(IUPAC Technical Report). Pure Appl. Chem. 87, 1051–1069 (2015).
48. Ayinla, R. T. et al. A review of technical advances of recent palm bio-waste conversion to activated carbon for energy storage. J.
Clean. Prod. 229, 1427–1442 (2019).
49. Toro, R. G. et al. Study of the Eect of Titanium Dioxide Hydrosol on the Photocatalytic and Mechanical properties of Paper sheets.
Mater. (Basel). 13, 1326 (2020).
50. Umeda, J. & Kondoh, K. High-purication of amorphous silica originated from rice husks by combination of polysaccharide
hydrolysis and metallic impurities removal. Ind. Crops Prod. 32, 539–544 (2010).
51. Zhuravlev, L. T. & Potapov, V. V. Density of silanol groups on the surface of silica precipitated from a hydrothermal solution. Russ
J. Phys. Chem. A. 80, 1119–1128 (2006).
52. Shahnani, M., Mohebbi, M., Mehdi, A., Ghassempour, A. & Aboul-Enein, H. Y. silica microspheres from rice husk: A good
opportunity for chromatography stationary phase. Ind. Crops Prod. 121, 236–240 (2018).
53. Nassar, M. Y., Ahmed, I. S. & Raya, M. A. A facile and tunable approach for synthesis of pure silica nanostructures from rice husk
for the removal of ciprooxacin drug from polluted aqueous solutions. J. Mol. Liq. 282, 251–263 (2019).
54. Santana Costa, J. A. & Paranhos, C. M. Systematic evaluation of amorphous silica production from rice husk ashes. J. Clean. Prod.
192, 688–697 (2018).
55. Hagab, R. H., Kotp, Y. H. & Eissa, D. Using nanotechnology for enhancing phosphorus fertilizer use eciency of peanut bean
grown in sandy soils. J. Adv. Pharm. Educ. Res. 8, 59–67 (2018).
56. Adel, A. M., El-Shall, F. N., Diab, M. A. & Al-Shemy, M. T. Biogenic silver-doped mesoporous silica nanoparticles for multifunctional
eco-designed textile printing. Biomass Convers. Biorenery. https://doi.org/10.1007/s13399-022-03643-2 (2022).
57. Prabha, S., Durgalakshmi, D., Rajendran, S. & Lichtfouse, E. Plant-derived silica nanoparticles and composites for biosensors,
bioimaging, drug delivery and supercapacitors: A review. Environ. Chem. Lett. 19, 1667–1691 (2021).
58. Hu, Y., Bai, S., Wu, X., Tan, S. & He, Y. Biodegradability of mesoporous silica nanoparticles. Ceram. Int. 47, 31031–31041 (2021).
59. Zhang, B. et al. Biodegradable hybrid mesoporous silica nanoparticles for gene/chemo-synergetic therapy of breast cancer. J.
Biomater. Appl. 33, 1382–1393 (2019).
60. Ojeda-Farias, O. et al. Dynamics of concentrated white portland cement suspensions using multispeckle diusing wave
spectroscopy. Constr. Build. Mater. 323, 126407 (2022).
61. Hussain, M., Sikandar, M. A., Nasir, A. & Ahmed, W. Eect of SiO2 coated leuco-dye based thermochromic pigment on the
properties of Portland cement pastes. J. Build. Eng. 35, 102019 (2021).
62. Yu, R., Spiesz, P. & Brouwers, H. J. H. Eect of nano-silica on the hydration and microstructure development of Ultra-high
performance concrete (UHPC) with a low binder amount. Constr. Build. Mater. 65, 140–150 (2014).
Scientic Reports | (2025) 15:2909 16
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
63. Abo-El-Enein, A., El-Gamal, S. M. A., Aiad, S. A., Azab, I. M., Mohamed, A. & M. & Eect of newly prepared admixture on the
Rheology and early hydration characteristics of the Oil Well Cement Pastes. Benha J. Appl. Sci. 1, 1–11 (2016).
64. El-Gamal, S. M. A., Abo-El-Enein, S. A., Ahmed, N. E. D. & Abdelnaby, A. A. Eect of novel binaphthol admixture on the hydration
characteristics and mechanical properties of OPC pastes. Adv. Cem. Res. 31, 331–342 (2019).
65. Mohsen, A., Abdel-Gawwad, H. A. & Ramadan, M. Performance, radiation shielding, and anti-fungal activity of alkali-activated
slag individually modied with zinc oxide and zinc ferrite nano-particles. Constr. Build. Mater. 257, 119584 (2020).
66. Heikal, M., Ali, A. I., Ibrahim, B. & Toghan, A. Electrochemical and physico-mechanical characterizations of y ash-composite
cement. Constr. Build. Mater. 243, 118309 (2020).
67. Heikal, M. Eect of temperature on the structure and strength properties of cement pastes containing y ash alone or in
combination with limestone. Ceram. Silik. 50, 167 (2006).
68. Heikal, M., Zaki, M. E. A. & Ibrahim, S. M. Characterization, hydration, durability of nano-Fe2O3-composite cements subjected
to sulphates and chlorides media. Constr. Build. Mater. 269, 121310 (2021).
69. Heikal, M., Helmy, I. M., Awad, S. & Ibrahim, N. S. Performance of silica-nano-particles on the physicochemical, and microscopic
characteristics of blended and composite cement. Ceramics–Silikáty 64, 320–337 (2020).
70. Selim, F. A., Hashem, F. S. & Amin, M. S. Mechanical, microstructural and acid resistance aspects of improved hardened Portland
cement pastes incorporating marble dust and ne kaolinite sand. Constr. Build. Mater. 251, 118992 (2020).
Acknowledgements
Not applicable.
Author contributions
(1) Abdallah A. Aziz: Writing- Original dra preparation, Methodology, Investigation, Soware, Writing-
Reviewing and Editing, Visualization. (2) H. F. Nassar*: Conceptualization, Resources, Investigation, Super-
vision.3. M.T. Al-Shemy: Conceptualization, Writing- Original dra preparation, Soware Supervision, Vis-
ualization, Validation, Investigation, Writing- Reviewing and Editing.4. O. A. Mohamed*: Conceptualization,
Writing- Original dra preparation, Methodology, Supervision, Investigation, Soware, Writing- Reviewing and
Editing, Visualization.
Funding
Open access funding provided by e Science, Technology & Innovation Funding Authority (STDF) in cooper-
ation with e Egyptian Knowledge Bank (EKB).
Declarations
Competing interests
e authors declare no competing interests.
Additional information
Correspondence and requests for materials should be addressed to H.F.N. or O.A.M.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access is article is licensed under a Creative Commons Attribution 4.0 International License, which
permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and
indicate if changes were made. e images or other third party material in this article are included in the article’s
Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included
in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or
exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy
of this licence, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2025
Scientic Reports | (2025) 15:2909 17
| https://doi.org/10.1038/s41598-025-86798-y
www.nature.com/scientificreports/
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
This research investigates the impact of incorporating low-cost MnFe2O4 spinel nanoparticles (MF-NPs) at varying concentrations (0.5, 1, and 2 mass%) on the mechanical and physical properties of blended cement pastes. These pastes were produced by replacing different proportions (5, 10, and 15 mass%) of ordinary Portland cement (OPC) with activated alumina sludge waste (AAS), to promote sustainability. Also, the research examined the gamma radiation shielding effectiveness of certain hardened composites against a ¹³⁷Cs gamma radiation source with an energy of 661.64 keV using a NaI (Tl) detector (Oxford Model) with 3″ × 3″, amplifier and 16 k multi-channel analyzer. The linear attenuation coefficient (LAC) of all the studded samples were calculated theoretically using a Monte Carlo code MCNP-5 code. The gamma radiation shielding properties were analyzed in depth using a Monte Carlo code MCNP-5 simulation model. The theoretical and experimental results for LAC were found to be in complete agreement. Phy-X/PSD software was applied to estimate the mass attenuation coefficient (MAC) for gamma radiation at various energies, as well as the effective atomic number (Zeff), mean free path (MFP), half-value layer (HVL), and tenth-value layer (TVL). The findings demonstrated that the addition of 0.5% MnFe2O4 nanoparticles (MF-NPs) to blended cement pastes exhibited the best physical and mechanical characteristics, as well as the most effective gamma radiation shielding.
Article
Full-text available
To understand the microstructural changes in the hardened cement paste during the drying, hardened white Portland cement pastes were D-dried or dried under 11, 33, 40, and 75% relative humidity, and then, pastes were impregnated with 2-propanol or water under vacuum. Measurement of 1 H NMR relaxometry was employed for the pastes before and after impregnation, and the full-scale microstructural changes of the hardened cement paste after drying were successfully visualized. It was observed that the volume of the coarse pores outside the agglomeration of calcium-silicate-hydrate (C-S-H) increased, and the volume of the C-S-H interlayer spaces and gel pores decreased as the drying relative humidity decreased. Thus, it was evidenced that interlayer space is decreasing during drying. It was also confirmed that 2-propanol can enter some of the interlayer space of C-S-H after harsh drying.
Article
Full-text available
The movement of the construction industry towards sustainable development has drawn attention to the revision of concrete. In addition to reducing pollution, the use of nano-materials should lead to the provision of higher quality concrete in terms of regulatory items (workability, resistance characteristics, durability characteristics, microstructure). The present study investigates 15 key characteristics of concrete modified with nano-CaCO3, nano-clay, nano-TiO2, and nano-SiO2. The results of the study showed that nanomaterials significantly have a positive effect on the hydration mechanism and the production of more C-S-H gel. The evaluation of resistance characteristics also indicates the promising results of these valuable materials. The durability characteristics of nano-containing concrete showed significant improvement despite high dispersion. Concrete in coastal areas (such as bridges or platforms), concrete exposed to radiation (such as hospitals), concrete exposed to impact load (such as nuclear power plants), and concrete containing recycled aggregate (such as bricks, tiles, ceramics) can be effectively improved by using nanomaterials. It is hoped that the current review paper can provide an effective image and idea for future applied studies by other researchers.
Article
Full-text available
Portland cement (PC) is a material that is indispensable for satisfying recent urban requirements , which demands infrastructure with adequate mechanical and durable properties. In this context, building construction has employed nanomaterials (e.g., oxide metals, carbon, and industrial/agro-industrial waste) as partial replacements for PC to obtain construction materials with better performance than those manufactured using only PC. Therefore, in this study, the properties of fresh and hardened states of nanomaterial-reinforced PC-based materials are reviewed and analyzed in detail. The partial replacement of PC by nanomaterials increases their mechanical properties at early ages and significantly improves their durability against several adverse agents and conditions. Owing to the advantages of nanomaterials as a partial replacement for PC, studies on the mechanical and durability properties for a long-term period are highly necessary.
Article
Full-text available
The current study concerns the valorization of waste rice hulls into value-added materials. To achieve this goal, first, the extraction route of mesoporous silica nanoparticles (SiO2NPs) from rice hulls using citric acid as a carbon template was compared with that prepared by conventional methods of incineration only. Both routes were performed at different temperatures of 600 and 700°C, and the fabricated SiO2NPs were studied by N2 physisorption isotherm, X-ray diffraction (XRD), Fourier transform infrared (FTIR), dynamic light scattering (DLS), and thermogravimetric (TG) analyses. The results show the influence of both routes and temperature on the porous structure, surface area, and charge of as-fabricated SiO2NPs. Furthermore, the physico-chemical features of as-prepared mesoporous SiO2NPs were utilized, in a facile and green way, to prepare silver NPs with high anti-microbial activity. The results were validated by the tools of UV-visible, FTIR, XRD, and XPS spectroscopy. Finally, the fabricated silver-doped mesoporous SiO2NPs were used as auxiliary additives through one-pot printing of different types of fabric (i.e., polyester (synthetic fabric), cotton (natural fabric), and cotton/polyester (blended fabric)). The EDX mapping of the pigment-printed fabrics with Ag-doped SiO2NPs exhibited the homogeneous distribution of Ag and Si atoms together with C and O throughout the fabric matrix. Interestingly, the pigment-printed fabrics with Ag-doped SiO2NPs demonstrated superior antibacterial activity, ultraviolet protection factor (UPF), and color characteristics in comparison to the conventional dyeing component. Graphical Abstract
Article
Elevated temperatures pose a serious potential risk to the mechanical and durability properties of cement-based composites, resulting in irreversible degradation, and eventual structural collapse. Various methods have been utilized to prevent deterioration and improve the thermal stability of cement-based composites when exposed to elevated temperatures. Furthermore, nano-reinforcement have more evident advantages than conventional materials, which fail to stop crack initiation at a smaller scale. This study summarizes the behaviors and changes of concrete at three phase of temperature, as well as a new and coherent spalling mechanisms. The predicted spalling temperature for each type of spalling is closely related to the changes of physic-chemical performance in concrete. The incorporation of various fibers and supplementary cementing materials effectively reduce the risk of the first two type of spalling. For thermo-chemical spalling, incorporating nanomaterials shows excellent potential in thermal stability of calcium silicate hydrate owing to their distinctive large surface area, flexibility in structure and size, reactivity, and easy functionalization. Moreover, molecular dynamic simulations are applied to further investigate the interaction between structure and mechanical properties of nano-reinforced composites. Limitations and developments are also discussed as topics for future research. The current work done could be used as the basis of the further research and the guidance of sustainable design.