Next Article in Journal
Impact of Entrepreneurial Leadership and Bricolage on Job Security and Sustainable Economic Performance: An Empirical Study of Croatian Companies during COVID-19 Pandemic
Next Article in Special Issue
A Sensitivity Analysis for Thermal Performance of Building Envelope Design Parameters
Previous Article in Journal
A Network Approach to Revealing Dynamic Succession Processes of Urban Land Use and User Experience
Previous Article in Special Issue
Effect of Aggregate and Binder Type on the Functional and Durability Parameters of Lightweight Repair Mortars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Thermal Behavior and Energy Efficiency of Modified Concretes in the Tropical Climate: A Systemic Review

1
Department of Civil and Construction Engineering, Faculty of Engineering and Science, Curtin University, CDT 250, Miri 98009, Sarawak, Malaysia
2
Department of Civil Engineering, College of Engineering, Prince Sattam Bin Abdulaziz University, Alkharj 11942, Saudi Arabia
3
Department of Civil Engineering, Faculty of Engineering and IT, Amran University, Amran 9677, Yemen
4
Department of Civil Engineering, Faculty of Engineering, Universiti Malaysia Sarawak, Kota Samarahan 94300, Sarawak, Malaysia
5
Curtin Malaysia Research Institute, Sarawak Biovalley Pilot Plant, Curtin University, Sarawak Malaysia, CDT 250, Miri 98009, Sarawak, Malaysia
6
Polytechnic Institute, Far Eastern Federal University, 690922 Vladivostok, Russia
7
Peter the Great St. Petersburg Polytechnic University, 195251 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(21), 11957; https://doi.org/10.3390/su132111957
Submission received: 9 April 2021 / Revised: 21 May 2021 / Accepted: 25 May 2021 / Published: 29 October 2021
(This article belongs to the Special Issue Thermal Behavior and Energy Efficiency of Buildings)

Abstract

:
Concrete remains the most utilised construction material for building envelopes, which regulate the indoor temperature to achieve human thermal comfort. Often, the energy consumption for building performance appraisal is related to the thermal behaviour of building materials as heating, ventilation, and air conditioning systems all variously contribute to human comfort. Following the development of concrete technology, many types of concrete have been invented to serve several purposes in the construction industry. To clearly understand the concrete type tailored for the specifics of a construction project, the local climate, concrete mechanical properties, and concrete thermal behaviours should be primarily identified to achieve energy efficiency, which also suits the sustainability of global materials. This paper, therefore, reviews the modified concrete thermal behaviours in the tropical climate for more systematic city planning in order to achieve better energy efficiency. Urban heat islands in the tropics and contributing factors, as well as heat transfer mechanisms, are first highlighted. The requirements of concrete thermal behaviour for building envelopes are then discussed through specific heat capacity, thermal conductivity, thermal diffusivity, time lag, and decrement factor in the context of applications and energy consumption in the tropical regions. With a case study, it is found that concrete thermal behaviours directly affect the energy consumption attributed mainly to the use of cooling systems in the tropics. The study can be a reference to mitigating the urban heat island phenomenon in the planning of urban development.

1. Introduction

Population centralisation and increment customarily take place in urban and metropolitan areas in which impermeable construction materials surfaces, notably concrete, are gradually replacing natural green vegetation zones due to various financial opportunities and accelerated infrastructure development. Despite numerous evident revolutions in building technologies, concrete remains the primary material employed in a diversity of construction events. Hence, building and structural engineers expectedly need to acquire more knowledge on concrete than other building and construction materials. Concretes possess comparatively high specific heat that may contribute to extra energy consumption during the building performance stage. Their low albedo surfaces often result in secondary heat to buildings, and thereby promote the urban heat island (UHI) phenomenon [1]. To better appreciate the effects, a 7 °C of variation was discovered between high-density residential areas and adjacent rural areas [2] due to UHI.
Since ancient times, building envelopes act as a sheltering establishment to regulate the indoor temperature and achieve human thermal comfort. Accordingly, building science has been greatly progressed and continuously advances with various innovative materials to form building envelopes. In terms of architecture, a building envelope can be categorised into physical form and orientation, opaque system, transparent system, and shading. Energy consumption is most relevant to the heat gain from opaque systems [3,4,5,6]. Heating and cooling systems are among the main contributors to energy consumption as the building envelope plays an essential role in issues of energy efficiency [7]. Existing studies have shown that some 30% of the power consumption within buildings in tropical climates is spent to cool down the heat from roofs and walls [8,9]. In Europe, 27% of the annual energy consumption was contributed to by residential dwellings alone in 2009 [10].
The thermal bridging of concrete tends to let the heat from outside transfer indoors through the conductivity of building materials [3]. There are many efforts introduced to regulate indoor temperature from outdoor thermal fluctuation, for instance, with lightweight concretes [3,11,12], insulation panels [13], etc. Even so, the unnecessary gross energy consumption for mitigating human thermal discomfort within the building compound, especially in tropic climates, remains unavoidably high, thus resulting in massive energy consumption through cooling systems. Therefore, understanding the thermal behaviours of concrete materials is imperative to assist in better planning and improving the energy expenditure performance of buildings in these regions. Concerning the energy consumption aspect, this paper reviews the concrete thermal behaviours to achieve energy efficiency in building performance with the final aim of reducing the carbon footprint. For the benefits of tropical climates, the concrete thermal behaviours are summarised and discussed in the context of applications relevant to these regions. As concrete strength and thermal insulating properties are the concerned factors in minimizing the energy consumption in the later building performance stage, the discussions are made on these parameters in preliminary study.

2. Tropical Climate and Issues

2.1. Tropical Climate General Description

As one of the climate groups in the Köppen classification, a tropical climate, which is located below 25 latitudes of southern and northern hemispheres, has been defined as having a mean temperature of ≥19 °C during the coldest month. The territories associated with the weather group are those around the Equator, consisting of parts of Australia, Pacific Ocean islands, southern Asia, Central America, parts of South America, and Central Africa. Following the rain intensity of the driest month as exhibited in Table 1, this climate can be further subdivided into the tropical rainforest, tropical monsoon, and savanna climates. Therefore, this weather group is characterised as a two-season climate zone, with distinctively wet and dry seasons throughout the year.

2.2. Urban Heat Island (UHI)

UHI is a phenomenon describing an isolated urban area associated with higher temperatures due to dense human population compared to its neighbouring suburban areas [17]. As a by-product of rapid development and civilisation, the trend of which is continually expanding worldwide, the UHI phenomenon may pose a new environmental threat towards the buildup of global warming. Urban areas with predominantly constructed materials over natural landscapes, aquatic terrains, and vegetation topographies are likely to attain a higher temperature than the surrounding outskirts. A UHI is likely to form in metropolitan areas that are characterised by a high percentage of water resistance, low vegetation, and non-reflective surfaces. In the UHI regions, the temperature is significantly higher during the nighttime, with no wind to dissipate the heat released from building materials back to the atmosphere.
Some consequences from UHI, such as an increase in monthly rainfall, prolonged growing seasons, and the deterioration of air and water qualities, negatively impact human health and daily activities, as well as the surrounding ecosystems.

2.2.1. Factors Contributing to UHI

Asphalt, bricks, and concrete with low albedo effects are the materials most associated with urban and metropolitan areas, which are credited to the presence of a vast amount of infrastructure and the ability to exert significant temperature variation compared with suburban areas [18]. These materials exhibit darker artificial surfaces (as opposed to those of nature) that tend to absorb more heat during the daytime. As there are significant differences in thermal bulk properties (heat capacity and thermal conductivity) and surface radiative properties (albedo and emissivity) between building materials and natural surfaces, a UHI can be formed under these circumstances. Moreover, the lack of vegetation areas due to replacement by buildings and paved-over surfaces, such as parking lots and roads, results in lesser evapotranspiration activity, a natural cooling process performed by green plants.
Furthermore, different building layouts can produce different thermal properties [19]. In urban areas, in which urban canyon effects are created as the result of the existence of high-rise buildings with multiple reflective surfaces, more heat is absorbed that encourages the formation of UHI. Therefore, city geometry also plays a major role in promoting the development of UHI. In addition, the poor use of thermal mass in a building can contribute to UHI by absorbing a lot of heat during the daytime and radiating it out at night, exacerbating energy consumption, energy costs, and comfort liability. The heat generated from the mechanical air conditioning systems to the surrounding environments may cause a rise in the ambient temperature, also known as waste heat. The massive use of air conditioning systems in urban areas generates relatively high waste heat compared to rural areas.

2.2.2. UHI in Tropical Climate

Table 2 summarises the UHI effects in the tropics. From UHI models, it was found that regional temperature could rise by up to 2.26 °C in a two decade time frame in Central Asia [20]. Several studies have also been actively conducted to evaluate the UHI in Malaysia, by means of numerical investigation [21,22,23], satellite imagery and remote sensing [24,25,26], surface energy balance modelling [27,28], weather station monitoring [29,30,31,32,33], etc. Although there is no significant borderline for urban and rural areas in Singapore, a maximum of a 4.01 °C temperature difference was found [34]. The UHI studies in South Asia also have been included in Table 3 of previous published data [35].

2.3. Heat Propagation Mechanisms

Heat propagates from one medium to another when there exists a temperature difference between them. Fundamentally, there are three types of heat transfer mechanisms, namely, conduction, convection, and radiation. Heat is transferred by conduction through a medium, and via air by means of convection and radiation. Without a global movement of the matter, heat can be conducted between molecules or close contact materials. Heat conduction is referred to as the internal thermal energy transfer produced by microscopic particles collisions and the movement of electrons within a body [42]. In a solid body, heat propagates through two mechanisms: the transmission of thermal energy or the vibration of atoms [43]. To reduce heat transfer by conduction, high thermal resistance materials are the most useful. In terms of heat convection, the transferring mechanism is exerted by the fluid mass motion. Heat convection happens when there is a difference in surface temperature from surrounding fluid [42]. Unlike conduction and convection, heat transmission by radiation can be formed by all physical substances via their electromagnetic energy emission due to rotational and vibrational movements of molecules and atoms. A substance can emit thermal radiation at temperatures hotter than 0 °K.

2.4. Building Envelope for Indoor Thermal Comfort

A building envelope is essential for the restriction of heat gain from the outside environment in achieving human thermal comfort. About 30% of power consumption is due to the heat gain from opaque systems (walls and roofs) in tropical climates [8,9]. Thermal insulation, radiation barriers, and air barrier systems are currently the chief building envelope methods in controlling heat transfer. On one hand, heat can potentially transfer by convection where air gaps are wider than 5 mm in construction materials [44], thereby increasing the building’s operation costs and energy consumption. On the other hand, heat radiation is predominantly due to the exposure of the roof to sunlight. It has been demonstrated by [45] that 50% of a building’s heat load comes from its roof. Studies have shown that reductions of 6% to 7.7% of heat transfer rate [46], 17% of energy load [47], and 70% heat flux [48] can be achieved with radiant barriers. The application of retroreflective coating materials can also reduce heat gain from solar radiation [49]. Hence, minimising the solar absorption is the main objective in the consideration of the building envelope through the employment of high reflectivity (high albedo) and insulating materials.
Therefore, the building envelopes made with concrete materials serves as a thermal barrier to maintain indoor human thermal comfort. If the building envelope is not functioning as a barrier, more energy is needed for HVAC systems in regulating the indoor ambient temperature. It is essential to explore more concrete thermal behaviour in order to minimize energy consumption in achieving human thermal comfort.

3. Concrete Thermal Behaviour

Concretes exhibit low time lag and high thermal conductivity, and therefore are commonly linked to the increasing trend of energy consumption in the tropics [50]. As already stated, a 30% rise in the energy consumption of residential houses as as result of the UHI phenomenon was discovered in a recent survey [51]. As heat propagates through concrete walls, the manner of the heat amplitude drop is referred to as the decrement factor, while the time delay is known as time lag. These parameters relate to the thermal comfort evaluation of the building envelope and in turn concern the energy consumption response of building materials [52]. A low resistance to temperature change is the thermal behaviour preferred in cold regions, while high thermal inertia is the desired condition in tropics [53]. Thus, thermal comfort can be attained with building materials of high time lag and a low decrement factor [54]. Other parameters, such as specific heat, diffusivity, and conductivity, are also imperative. These engineering terminologies are the next subjects of discussion in the assessment of concrete thermal behaviour.

3.1. Thermal Conductivity

Several factors affect the concrete thermal conductivity [55]. For instance, the type of aggregate [56,57], the properties of cementitious materials [58,59,60,61,62,63,64,65,66], moisture content [67,68], design buildup [69,70,71,72,73,74], temperature [75,76], concrete density [12,77,78,79,80], etc. The measurement of concrete thermal conduction is divided into steady-state and transient methods [81]. A constant heat transfer is referred to as a steady-state condition, a description that is relevant to homogenous materials. The required measurement is time-consuming while more accurate results can be commonly obtained. Meanwhile, time and temperature dependencies inherent in the transient method are more suitable for heterogeneous materials with moisture content variation [82]. A previous study quantified that hot wire (transient approach) was the most popular method to determine the thermal conductivity of concrete [81]. For a general perspective of currently available efforts in literature, Table 3 summarises different concretes with their thermal conductivities (k-values), obtained using theoretical model, steady-state, and transient approaches.
Table 3. Concrete thermal conductivity (k-value).
Table 3. Concrete thermal conductivity (k-value).
Ref.Mix Description Measurement Density, kg/m3 k-Value, W/m °CReferred Guideline
[82]Self-Compacting Concrete with PerliteSteady-state hot plate2227
2253
2279
2292
2.09430–40–0.71570–80
2.17930–40–0.76170–80
2.35530–40–0.76170–80
2.43030–40–0.79370–80
ASTM C177
[83]Oil palm shell foamed concrete with fly ash and silica fumeSteady-state hot plate 1156
1192
1354
1409
1506
1594
0.40
0.41
0.50
0.54
0.55
0.57
BS EN12664
[84]Oil palm shell foamed geopolymer concrete with POFA and fly ash Steady-state hot plate 1300–18000.47–0.58 BS EN12664
[85]Polystyrene foamed concrete Steady-state hot plate150
200
250
400
0.0848
0.0864
0.0927
0.1566
Not specified
[86]Lightweight aggregate and glass bead Transient hot wire 18001.1–1.4ASTM D5334
[87]Lightweight aggregate concrete with diatomite and pumice Transient hot wire1500
900
0.44
0.13
ASTM C1113
[88]Modified waste expanded polystyrene lightweight aggregate concrete Transient hot wire876–19560.6–1.99ASTM C1113
[89]Expanded perlite lightweight aggregate concrete with silica fume and fly ash Transient hot wire509
493
485
511
498
483
0.1720
0.1552
0.1558
0.1676
0.1643
0.1472
ASTM C1113
[90]Autoclaved aerated concrete Transient plane source 415
520
630
0.1–0.2Not specified

3.1.1. Theoretical Model

The cubic model is one of the prediction models of concrete thermal conductivity, which accounts for the individual conductivity of cement paste, aggregate, and aggregate volume, as mathematically displayed in Equation (1). It is noted that the model may suffer from discrepancies due to the effects of minerals of different types even though the contained aggregates have the same density, thus needing extra care in the prediction event. For practice, ACI [91] suggests some practical values for thermal conductivity of both normal and lightweight aggregate concretes, which can be readily applied without the experimental data, such as in the following equation:
k c = k p [ V a 2 / 3 V a 2 / 3 V a + ( V a ( k a V a 2 / 3 k p ) + 1 V a 2 / 3 ) ]
where Va is the aggregate volume, kp is the thermal conductivity of cement paste, ka is the thermal conductivity of aggregate, and kc is the thermal conductivity of concrete.
Equation (2) expresses the relationship between thermal conductivity (kc) and density (d) of the concretes. It should be noted that the kc values of concretes that have same densities, but are made of different aggregates, can deviate from the calculated values derived from Equation (2). For instance, the kc values obtained for normal weight and lightweight concretes might be underestimated because the different types of aggregates contained in the concretes have their own specific thermal properties.
k c = 0.072 e 0.00125 d ( S . I .   units )
Practically, concretes contain moisture. Equation (3) is therefore used to provide a more accurate kc value by taking the densities of concrete in moist and oven-dry conditions into account:
k c ( c o r r e c t e d ) = k c [ 1 + ( 6 d m d o ) d o ]  
where kc is the thermal conductivity of the concrete, and dm and do are the densities of the concrete in moist and oven-dry conditions, respectively.

3.1.2. Steady-State Approach

The steady-state boxed method [92] operates by determining the energy transferred from one hot end to the cold end based on the second law of thermodynamics. The k-value can be calculated through the air temperature difference of both sides [68,93]. Meanwhile, a guarded hot plate measures the thermal conductivity with the mean temperature difference between two surfaces [94]. The temperature range can be set within the plates as studied by [82,83]. The steady-state approach is usually used for thermal conductivity measurement of homogenous materials in which the temperature is time -independent. Accurate results can be obtained but it is time consuming. Equation (4) is the Fourier Law of heat conduction for thermal conductivity measurement:
q = k T
where q is the heat flux, k is the thermal conductivity, and ∇T is the gradient of temperature.

3.1.3. Transient Method

On the other hand, the transient method is preferred to measure the thermal conductivity of moist concrete that is made of heterogeneous materials. Hot wire and plane source are the most common methods employed under the transient approach category. The temperature is measured at a specific distance from the hot wire [95]. In various studies on lightweight aggregate concretes, thermal conductivities have been determined using such a setting [88,96,97]. Moreover, the plane source method measures the thermal conductivity of power input and time variation for the transient plane and line sources. Studies relevant to this method for the measurement of thermal conductivity are [90,98].
Thermal conductivity is calculated using Equation (5) below:
k c = γ D ( τ i ) / E ( t )
where kc is the thermal conductivity of the concrete, γ is the constant of the different resistances in the Wheatstone bridge, D(τi) is the theoretical expression of time-dependent increase, and ∆E is the variation in potential across the Transient Plane Source sensor.
Other than heat transfer, the thermal inertia of building materials also affects the indoor temperature. Thermal inertia characterises the reluctance of temperature change, and the abilities of heat absorption and storage in a building. In tropical regions, where thermal comfort is affected by the hot and humid climate, a comfortable interior temperature could be achieved by using concrete walls with high thermal inertia [99]. The heat diffusivity, capacity, time lag, and decrement factor of building materials are all often measured to determine the thermal inertia. The relationship of thermal inertia with material density, specific heat, and thermal conductivity is expressed in Equation (6):
I = ρ c k
where I is the thermal inertia, while ρ, c, and k are the density, specific heat, and thermal conductivity of the material, respectively.
For a well-thermal-insulated concrete, low-rate thermal diffusivity and heat capacity are important. Table 4 lists the thermal inertia obtained for some selected construction materials.
Thermal diffusivity is the heat transmission rate within a material, which is mathematically defined by the k-value divided by specific heat capacity and density for the same unit of pressure. The specific heat capacity is the amount of heat required to raise the unit temperature for a given mass. Concrete, with low thermal diffusivity but high specific heat capacity as compared to other construction materials, is preferable as a building material.
The time lag and the decrement factor relate closely in defining the indoor thermal comfort in a building. The time of the peak of outdoor heat to appear indoors, is referred to as time lag, while the temperature decrease from the peak temperature is known as its decrement factor. These values can be obtained from the calibrated hot box and in situ field tests where temperature changes are measured. For instance, 6.9 h of time lag was reported for a piece of 250 mm thick concrete [100]. The effectiveness of the decrement factor can be achieved with the increment in insulation thickness, which is a direct indicator of cost. For tropic climate applications, it can be generally ascertained that concretes remain, by comparison to other construction materials, in the category of having good performance in terms of the aforementioned thermal properties, although the most optimal material design can be always be examined and enhanced.

3.2. Other Thermal Behaviours

Concrete expands slightly when the temperature rises and contracts when the temperature falls. This expansion is known as thermal expansion, which is usually influenced by the types of aggregate, cement, and water content and the age of the concrete [101,102]. The coefficient of the thermal expansion of concrete is about 10 millionths per degree Celsius (10 × 10−6 °C), indicating that a 1 m long piece of concrete subjected to a temperature increase of 1 °C will expand by 10 × 10−6 m, or 0.010 mm. Problems occur when the heat in the massive concrete structures cannot be dissipated. Temperature changes that cause a thermal differential can potentially induce thermal stress in the concrete and eventually lead to cracking over time. Control joints are an effective way to control cracking by providing space for the expansion and shrinkage of the concrete. Concrete with high melting points are likely to have lower thermal expansions.

4. Energy Consumption

Energy is utilised in buildings to maintain human comfort. An inefficient building envelope may cause high heat transfer between indoor and outdoor environments. As heat gain and loss are not feasibly eliminated, efforts have been applied to minimise heat transfer to improve energy efficiency. In the study of energy consumption, energy waste can be found from air leakage and insufficient insulation. Previous research also found that lighting systems also contribute to the overall energy consumption, as summarised in Table 5. The heat gain or loss by a building envelope is greatly related to the energy consumption of a building. As shown in Figure 1, the UK and Sri Lanka consumed the most energy in the building performance category.
In general, the energy utilised for maintaining building performance is especially higher in the tropics to achieve human thermal comfort. There are also some solutions embedded with negative impacts, such as the improper height of building and width of road, lack of green areas, inappropriate building geometry, etc. To reduce the energy consumption, some procedures are recommended, namely, natural ventilation, appropriate building materials, and increased green areas.

4.1. Sustainability in Concrete Construction

In the concrete industry, there are several solutions to reduce the carbon footprint, such as phase-change materials [108,109,110], lightweight concrete [11,12,55,111,112,113], or cement-free concrete [59,65]. Some of these reduce the carbon footprint through minimising the use of cement since its production contributes much carbon to the environment, while some solutions provide more long-term effects on building performance or services. Lightweight concrete is one of the alternatives to reduce energy consumption in the building service stage, while phase-change materials may store or release energy according to the surrounding temperature.
As buildings may have a serviceability period of 50 years, where the cumulative saved energy may become greater, lightweight concrete and phase-change materials in concrete are therefore the focus in the current research trend towards sustainability. One of the challenges of lightweight concrete is the reduction in strength as the voids in the concrete matrices increased. A balance is needed in order to satisfy both design specifications of structural strength and thermal insulating properties.

4.2. Prediction Models for Residential Energy Consumption

Electricity is the main source of residential energy consumption. However, these data are captured through automatic meter reading or advanced metering infrastructure and are not available to the public. To circumvent this, the only prediction models for energy consumption are developed as observed in the current research trend, even though several case studies may be available upon permission of authorities. Prediction methods for energy consumption are customarily derived from physics-based (engineering), data-driven inverse modelling (artificial intelligence), or a combination of both approaches [114]. Table 6 shows the available prediction models adopted for residential energy consumption, namely regression analysis, artificial neural networks, the Gaussian model, etc. It is worth noting, however, that all these models do not represent the full capacity or a well-accepted consensus on current or future predictions of energy consumption. This is due to several governing factors, e.g., different consumption patterns for different families, different electrical appliances, different perceptions towards human comfort, etc. Therefore, the existing models are still under development and are far from maturity.

4.3. Concrete Properties in Energy Consumption

Concrete can be divided into structural and non-structural uses. Concrete for structural use has a higher density as it can provide load-bearing characteristics in building envelopes. Despite their densities, concretes may contribute variously to thermal behaviours. Regarding density, the energy consumption of a concrete building can be further discussed. From Table 7, although low thermal properties can be achieved with lower densities, they may not achieve structural requirements in building construction. Depending on the requirements, as long as the designed usage is clear, both types can be applied for building construction.
Placing high expectations on the load-bearing requirement, designers often ignore concrete thermal properties even though they are essential in the consideration of energy consumption at a later serviceability stage. In this regard, thermal conductivity is a contributing factor to the energy consumption of the building performance. The heat from the sun is radiated into the building envelope, mostly onto walls and roofs, and raises the external surface temperature, which relates to the specific heat of concrete.
Furthermore, heat conduction will take place when there is a temperature gradient, from a hotter surface towards the inner wall. Thus, this potentially increases the indoor ambient temperature and cooling loads may be applied to achieve human thermal comfort, yielding a rise in the overall energy consumption level.
For tropical climate, thermal conductivity is correlated to concrete’s compressive strength and density, as described by Equations (7) and (8) [115]. To achieve 17 MPa of structural use for foamed concrete, 1.8 W/mK of thermal conductivity and 200 kg/m3 density are obtained from Equations (7) and (8), respectively. It is thus far unproven that 17 MPa can be achieved with 200 kg/m3 concrete density. Therefore, optimisation in various aspects of concrete produc tion is needed to fit into the climate requirement.
Thermal   conductivity = 0.9558 ln f c 0.8871
Thermal   conductivity = 0.0311 e 0.019 γ
where fc and γ are the compressive strength and density of concrete, respectively.
Table 6. Residential energy consumption prediction.
Table 6. Residential energy consumption prediction.
Ref.Investigated Method/ParameterBuilding Performance/Prediction Model Remark
[116]Sensor measurements collected every 15 minLinear Regression; Feed Forward Neural Network (FFNN); Support Vector Regression (SVR); Least Squares Support Vector Machines
(LS-SVM); Hierarchical Mixture of Experts (HME) with Linear Regression Experts; HME with FFNN Experts; and Fuzzy C-Means with FFNN.
Least squares support vector machine is the best for future electrical consumption prediction
[117]Building global heat loss coefficient (G), the south equivalent surface (SES), and the difference
between the indoor set point temperature and the sol-air temperature
Multiple regression prediction model Three inputs and one output, simplicity, large applicability, good match with the simulations and the energy certification
calculations plus the inclusion of human behaviour correction
[118]Day-ahead electricity load forecastingSeasonal Autoregressive Integrated
Moving Average
Ability to capture a variety of input factors that influence load behaviour and to enable
a dynamic trade-off between them, as well as between different model parameter values
[119]Total energy and total electricity consumptions Artificial Neural Network (ANN) model, Grey models, Regression model, Polynomial
Model, and Polynomial regression model
ANN is the most acceptable forecasting method
[120]Different material thicknesses and insulation propertiesExtreme Learning Machine (ELM) and Artificial Neural Network (ANN)ELM algorithm can achieve the smallest training error and also
normalise the weights
[121]Hourly prediction of building electricity consumptionImproved Particle Swarm Optimisation algorithm (iPSO) and Genetic Algorithm Artificial Neural Networks (ANNs), iPSO-ANN, and GA-ANN iPSO-ANN model has a shorter modelling time
[122]HVAC hot water energy consumptionsChange-point regression model, Gaussian process regression model (GPM), Gaussian Mixture Regression Model (GMM), and
Artificial Neural Network model (ANN)
GMR model shows slightly better statistical performance; insufficient data training of ANN results in an inaccurate prediction
[123]Temperature-dependent change point model selectionChange point modelAble to select the most appropriate temperature-dependent change point model for all 48 cases tested
[124]Daily residential energy useChange-point models for residential energy useEvaluated the differences in energy slope when compared with energy audit data; increasing the thickness of the duct insulation
[125]Heating and cooling loads of residential buildings Geometric Semantic Genetic Programming (GSGP)Integrating a local searcher and linear scaling in GSGP can speed up the convergence of the search process
[126]Whole-building energy Three-parameter change-point regression model36.9% global CV (RMSE) of the initial simulation was improved to 8.8% after a calibrated simulation
[127]Optimal ON/OFF status for home appliancesLightning Search Algorithm (LSA)-based Artificial Neural Network (ANN)Reduce the peak-hour energy consumption
[128]Feasible Time-of-Use (ToU) tariffsGaussian Mixture ModelGrouping half-hour interval flat-rate tariffs within a day into clusters to determine ToU tariffs
[129]Electricity consumption with three consumption profiles Constrained Gaussian Mixture ModelConsumer behaviour evolves over time
depending on the contextual variables
[130]Weather, time of day, and previous consumptionSupport Vector Regression (SVR)Impact of temporal and spatial granularity
Table 7. Concrete properties from previous research.
Table 7. Concrete properties from previous research.
Ref.Type of ConcreteDensity, kg/m3Compressive Strength, MPaThermal Conductivity, W/mKSpecific Heat Capacity, kJ/kgKThermal Diffusivity, ×10−6 m2/sRemark
[131]Foamed concrete --FA = Fly ash, F = foam
FA0-F301346.251.620.086
FA10-F301403.6612.760.088
FA20-F301489.3452.930.089
FA30-F301524.3453.5050.09
FA0-F40855.570.260.0842
FA10-F40908.30.360.0853
FA20-F40980.980.530.0863
FA30-F401010.740.620.0875
[115]Different grades of normal weight concrete with different proportions, water/cement ratios, and superplasticiser contents2205 to 249815.8 to 62.11.61.040.69
2.21.10.87
2.511.06
2.70.951.19
2.41.041
2.91.011.18
2.30.991.02
1.81.160.7
2.31.030.93
2.70.941.2
2.41.110.94
2.51.041.01
3.20.951.34
2.90.921.29
[132]Formcrete650 0.23--
Cement:sand = 2:17000.24
Water/cement = 0.58000.26
9000.28
10000.31
11000.34
12000.39
[133]Concrete with expanded glass granules (GG) --GG300 = expanded glass granules 300 g
GG400-S100 = expanded glass granules 300 g + sand 100 g
GG3006855.80.163
GG4005614.10.141
GG50055840.14
GG400-S1005233.20.138
GG400-S2006453.80.161
GG400-S3006994.50.177

5. Concrete in Tropical Climate

In the tropics, hot and humid are the primary characteristics of the local weather conditions. In the same region, the average temperature ranges from 20 °C to 30 °C throughout the year. In order to reduce the heating and cooling loads for a building, concrete with minimum heat transfer may be applied to maintain indoor thermal comfort. Specifically, concretes with low specific heat and conductivity are highly preferable in this weather. In terms of thermal behaviour at ambient temperature, a high time lag and a low decrement factor [99] are the other two desired performance characteristics. A high time lag and a low decrement factor illustrate that the heat is transferred slowly from outside to insider within a long time frame with low heat dispersion.
Additionally, the construction material thickness and the insulation system have been studied for the consideration of energy consumption [120]. The increase of the material thickness may effectively prevent the heat being transferred from the outside at the expense of an increment in material cost. A standardised single brick thickness (112.5 mm, in Malaysia) wall is usually applied in tropical countries, such that the effectiveness of heat transfer restriction can be optimised by using concrete with low density.
Lightweight concrete is an alternative in achieving both mechanical and functional properties when relating to thermal performance [11,12,55,113]. From non-structural (brick, 5 MPa) to structural uses (17 MPa), the concrete densities can range between 1000 to 1800 or 2000 kg/m3 for foamed concrete [11,134] or lightweight aggregate concrete. These lightweight concretes have relatively low thermal conductivity. Table 8 presents the cooling load for structural and non-structural concrete, with an 8 °C temperature difference of a 1 m2 wall surface. Depending on the concrete’s application, there is a varying range of concrete density and strength that can be applied within the construction industry. The cooling load is calculated according to Equation (9):
Cooling   load = US   ( T o T i )
where US is the transmission area and (ToTi) is the temperature difference between outside and inside the building (8 °C taken from the previous study, [55]).
In the tropics, the hot weather increases the energy consumption of HVAC systems in order to achieve human thermal comfort. The pores, or concrete density, governs the heat transfer of the building envelope. Concerning the heat barrier of concrete, lightweight concrete serves as an alternative in the tropics to reduce energy consumption. The strength and pores of concrete are two contrasting properties and a balance should be achieved to move towards sustainability.

5.1. Preliminary Study

It is necessary to quantify the energy consumption or energy saved when applying different types of concretes using a cooling load calculation or simulation. Before quantifying the energy consumption, the hypothesis of thermal resistance for lightweight concrete is better than normal concrete. In this preliminary study, the surface temperatures (the internal and external surfaces) of the casted wall sample and the conventional wall sample are recorded for further comparison. Figure 2 shows the experimental setup of the preliminary study.
The design mix of OPS & 50-QD of [55,135] has been adopted for preliminary data collection from tropical climate conditions and was compared to conventional concrete, as shown in Figure 2. The mix contained oil palm shells (OPS) as coarse aggregate and 50% quarry dust as a fine aggregate replacement. Scaled walls measuring 300 × 300 × 102 mm have been constructed. One surface of the wall was exposed to sunlight for 3 hours and another, opposite side was covered by polystyrene.
As shown in Figure 3, the surface temperatures of conventional concrete were found to not have a significant difference where variation was ranged in 1 °C or 2 °C, whereas, for lightweight concrete using OPS may have higher a variation at a particular time. The gradient of the graph for conventional concrete is steeper than the lightweight concrete. This can be related to the specific heat capacity since conventional concrete possesses higher specific heat capacity, which is a disadvantage during nighttime, as the heat will be released to the building’s interior. There is a lack of information on lightweight concrete used in building envelopes as most of its applications have secondary uses in the construction of buildings. Therefore, further investigation on heat transfer should be conducted for lightweight concrete in tropical climate.

5.2. Case Study on a Residential House

The energy consumption for HVAC also related to the thermal properties of concrete, time lag, and decrement factors, and this case study is performed to examine the effects of lightweight concrete in cooling load calculation, assuming that the inner surface temperature is the indoor ambient temperature.
To exemplify the concrete thermal behaviour in a tropic climate, the surface temperature of an external wall of a residential house in Johor Bahru, Malaysia [136] has been used in the case study. The time lag and decrement factor, as well as mechanical properties of various concretes, have been obtained from a previous study [135], as summarised in Table 9. Some assumptions have been made to calculate the total amount of electricity used with different types of concrete, such that the material indoor surface temperature is the indoor ambient temperature. The surface temperature and ambient temperature were collected using thermocouple wire type-T with Graphtec Midi data logger and bad HOBO data logger, respectively. Otherwise, there are no other contributing factors for the concrete’s thermal behavior.
The estimated inner wall surface temperatures are plotted for various concretes, as displayed in Figure 3. The reference line is introduced at 25 °C, as this is the average temperature for human thermal comfort [104]. Whenever the curves are higher than 25 °C, it is assumed that the cooling system is required. In other words, energy is required for the air-conditioning to regulate the drop in temperature. It can be seen that the concrete with 25% lightweight expanded clay (LECA) and 75% oil palm shell (OPS) with the lowest density of 1780 kg/m3 demands relatively low energy for cooling loads. Other concrete walls necessitate more operational time of the cooling system. Although this estimation seems simple and may need further improvement and verification, the exercise illustrates qualitatively the effects of walls made from different concretes on the thermal flow behaviors. Next, the period time that the temperature was above the reference temperature has been measured and divided by the 48 hours of the designated total time frame. Table 9 shows the percentage of the resulting cooling load from which lightweight aggregate concrete with OPS showed a lower cooling load than the other concretes. The pores in the OPS and lightweight aggregates reduce the heat transfer within the concrete, proven by the decrement factor, leading to lower amounts of energy required for the indoor cooling system. The pores relate to the concrete density formation, and more pores means less density and lower heat transfer in concrete. With superior thermal resistance capability, the concrete strength is neglected, as both properties are not able to exist together. However, lightweight aggregate concrete with OPS possesses a lower density with a minimum structural requirement of 17 MPa according to the ACI specification and has better thermal performance. Hence, these two concrete mixes are highly recommended for future study in the consideration of optimal thermal behaviors coupled with mechanical considerations for tropical applications. Figure 4 shows the wall inner surface temperature estimation.

5.3. Further Investigation and Application

In order to obtain more reliable thermal barriers in concrete building envelope design, quantifying the saved energy from various types of concrete can be an useful information in urban development planning. Energy consumption can be theoretically determined by simple cooling load calculation using Equation (9) [55]. Compressive strength is related to thermal conductivity as described in Equations (7) and (8). The correlation of the cooling load (concrete voids) and concrete strength can be proposed in future study.
For more accurate energy saving quantification, numerical investigation is one of the research trends that can be ventured for systematic urban planning. Other than concrete construction, thermal behaviour of dry construction and various types of building (such as commercial buildings) can also be studied for better energy consumption strategies in terms of a building’s performance stage. Development is a continuous activity that may cause environmental issues and systematic planning should be applied to minimize the negative effects.

6. Conclusions

This paper reviews concrete thermal behaviours in the context of a tropical climate. Based on this review study, several conclusions have been drawn as follows:
i.
Temperature differences observed in UHIs of tropical countries have been highlighted.
ii.
The building envelope is used to maintain the indoor temperature from outside temperature fluctuation and achieve human thermal comfort. Building materials are the sources of the heat regulation for the building envelope. Heat is transferred through conduction, convection, and radiation within the envelope before reaching indoors.
iii.
Concrete thermal behaviours, such as thermal conductivity and thermal inertia, in terms of heat diffusivity, heat capacity, time lag, and the decrement factor can be determined and utilised for building materials to minimise heat transfer, and thus reduce energy consumption since residential energy consumption contributes predominantly to the energy consumption of the overall country.
iv.
Residential energy consumption remains the highest amongst all building energy expenditure performances in one tropical case study.
v.
Current residential energy consumption models are still under the development and are far from maturity for real scenario application. Further study is warranted.
vi.
Due to hot and humid weather conditions in the tropics, low specific heat and thermal conductivity are required to minimise the heat transfer through the building envelope. In other words, high time lag and a low decrement factor should be achieved for optimal energy efficiency in this weather condition.
vii.
From a case study of a residential house located in Johor Bahru, Malaysia, a lower density concrete with a low decrement factor displays a superior thermal performance attributed to the lowest cooling load required to maintain the indoor temperature. Reducing the consumed energy paves the way towards the sustainability of human comfort in the tropical zone.

Author Contributions

Data curation, Y.H.L., M.A., A.B.H.K., Y.Y.L. and S.F.K.; formal analysis, Y.H.L., A.B.H.K., Y.Y.L. and S.F.K.; funding acquisition, M.A., R.F., N.V. and Y.V.; investigation, Y.H.L., A.B.H.K., Y.Y.L. and S.F.K.; Methodology, Y.H.L., Y.Y.L. and S.F.K.; Project administration, Y.H.L., M.A., A.B.H.K., R.F., N.V. and Y.V.; resources, Y.H.L., M.A., A.B.H.K., Y.Y.L. and S.F.K.; supervision, Y.H.L., A.B.H.K. and M.A.; validation, Y.H.L., M.A., A.B.H.K., Y.Y.L., S.F.K., R.F., N.V. and Y.V.; visualization, Y.H.L., Y.Y.L. and S.F.K.; writing—original draft, Y.H.L., M.A., A.B.H.K., Y.Y.L. and S.F.K.; writing—review and editing, Y.H.L., M.A., A.B.H.K., Y.Y.L., S.F.K., R.F., N.V. and Y.V. All authors have read and agreed to the published version of the manuscript.

Funding

The experimental works of this research were supported by Curtin University Malaysia and University of Malaysia Sarawak, Malaysia under the Cross Disciplinary Research Grant (F02/CDRG/1818/2019). The research was funded by the Ministry of Science and Higher Education of the Russian Federation as the grant Self-Healing Construction Materials (contract No. 075-15-2021-590 dated 4 June 2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors gratefully acknowledge the financial support given by the Deanship of Scientific Research at Prince Sattam bin Abdulaziz University, Alkharj, Saudi Arabia, Curtin University Malaysia and University of Malaysia Sarawak, Malaysia, Moscow Automobile and Road Construction University, Moscow, Russia; and the cooperation of the Department of Civil Engineering, Faculty of Engineering and IT, Amran University, Yemen, for this research.

Conflicts of Interest

The authors declare no conflict of interest.

Correction Statement

This article has been republished with a minor correction to the reference 18. This change does not affect the scientific content of the article.

References

  1. Jhatial, A.A.; Goh, W.I.; Mohamad, N.; Rind, T.A.; Sandhu, A.R. Development of Thermal Insulating Lightweight Foamed Concrete Reinforced with Polypropylene Fibres. Arab. J. Sci. Eng. 2020, 45, 4067–4076. [Google Scholar] [CrossRef]
  2. Shahmohamadi, P.; Che-Ani, A.I.; Maulud, K.N.A.; Tawil, N.M.; Abdullah, N.A.G. The Impact of Anthropogenic Heat on Formation of Urban Heat Island and Energy Consumption Balance. Urban Stud. Res. 2011, 2011, 497524. [Google Scholar] [CrossRef]
  3. Real, S.; Gomes, M.G.; Rodrigues, A.M.; Bogas, J.A. Contribution of structural lightweight aggregate concrete to the reduction of thermal bridging effect in buildings. Constr. Build. Mater. 2016, 121, 460–470. [Google Scholar] [CrossRef]
  4. Korsun, V.I.; Vatin, N.; Korsun, A.; Nemova, D. Physical-Mechanical Properties of the Modified Fine-Grained Concrete Subjected to Thermal Effects up to 200 °C. Appl. Mech. Mater. 2014, 633-634, 1013–1017. [Google Scholar] [CrossRef]
  5. Korsun, V.; Korsun, A.; Volkov, A. Characteristics of mechanical and rheological properties of concrete under heating conditions up to 200 °C. In MATEC Web of Conferences; EDP Sciences: Les Ulis, France, 2013. [Google Scholar] [CrossRef]
  6. Korsun, V.I.; Khon, K.; Ha, V.Q.; Baranov, A.O. Strength and deformations of high-strength concrete under short-term heating conditions up to + 90 °C. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2020. [Google Scholar] [CrossRef]
  7. Garay, R.; Uriarte, A.; Apraiz, I. Performance assessment of thermal bridge elements into a full scale experimental study of a building façade. Energy Build. 2014, 85, 579–591. [Google Scholar] [CrossRef]
  8. Chua, K.J.; Chou, S. Energy performance of residential buildings in Singapore. Energy 2010, 35, 667–678. [Google Scholar] [CrossRef]
  9. Zingre, K.T.; Wan, M.P.; Tong, S.; Li, H.; Chang, V.W.-C.; Wong, S.K.; Toh, W.B.T.; Lee, I.Y.L. Modeling of cool roof heat transfer in tropical climate. Renew. Energy 2015, 75, 210–223. [Google Scholar] [CrossRef]
  10. Lapillonne, B.; Sebi, C.; Pollier, K.; Mairet, N. Energy Efficiency Buildings in the EU Trends—Lessons from the ODYSSEE MURE Project; ADEME: Paris, France, 2012. [Google Scholar]
  11. Amran, M.; Fediuk, R.; Vatin, N.; Lee, Y.H.; Murali, G.; Ozbakkaloglu, T.; Klyuev, S.; Alabduljabber, H. Fibre-Reinforced Foamed Concretes: A Review. Materials 2020, 13, 4323. [Google Scholar] [CrossRef]
  12. Amran, M.; Lee, Y.H.; Vatin, N.; Fediuk, R.; Poi-Ngian, S.; Lee, Y.Y.; Murali, G. Design Efficiency, Characteristics, and Utilization of Reinforced Foamed Concrete: A Review. Crystals 2020, 10, 948. [Google Scholar] [CrossRef]
  13. Amran, Y.M.; El-Zeadani, M.; Lee, Y.H.; Lee, Y.Y.; Murali, G.; Feduik, R. Design innovation, efficiency and applications of structural insulated panels: A review. Structures 2020, 27, 1358–1379. [Google Scholar] [CrossRef]
  14. Bierregaard, R.O., Jr.; Lovejoy, T.E.; Kapos, V.; dos Santos, A.A.; Hutchings, R.W. The biological dynamics of tropical rainforest fragments: A prospective comparison of fragments and continuous forest. Bioscience 1992, 42, 859–866. [Google Scholar] [CrossRef]
  15. McKnight, T.L. Climate zones and types. Phys. Geogr. Landsc. Apprec. 2000, 42, 859–866. [Google Scholar]
  16. Iwai, C.B.; Oo, A.N.; Topark-Ngarm, B. Soil property and microbial activity in natural salt affected soils in an alternating wet–dry tropical climate. Geoderma 2012, 189-190, 144–152. [Google Scholar] [CrossRef]
  17. Masumoto, K. Urban heat islands. In Environmental Indicators; Springer: Dordrecht, The Netherlands, 2015. [Google Scholar] [CrossRef]
  18. Ghazanfari, S.; Naseri, M.; Faridani, F.; Aboutorabi, H.; Farid, A. Evaluating the effects of UHI on climate parameters (A case study for Mashhad, Khorrasan). WSEAS Trans. Environ. Dev. 2009, 5, 508–517. [Google Scholar]
  19. Din, M.F.M.; Lee, Y.Y.; Ponraj, M.; Ossen, D.R.; Iwao, K.; Chelliapan, S. Thermal comfort of various building layouts with a proposed discomfort index range for tropical climate. J. Therm. Biol. 2014, 41, 6–15. [Google Scholar] [CrossRef]
  20. Zong-Ci, Z.; Yong, L.; Jian-Bin, H. Are There Impacts of Urban Heat Island on Future Climate Change? Adv. Clim. Chang. Res. 2013, 4, 133–136. [Google Scholar] [CrossRef]
  21. Morris, K.I.; Chan, A.; Morris, K.J.K.; Ooi, M.C.G.; Oozeer, M.Y.; Abakr, Y.A.; Nadzir, M.S.M.; Mohammed, I.Y.; Al-Qrimli, H.F. Impact of urbanization level on the interactions of urban area, the urban climate, and human thermal comfort. Appl. Geogr. 2017, 79, 50–72. [Google Scholar] [CrossRef]
  22. Ooi, M.; Chan, A.; Ashfold, M.; Morris, K.; Oozeer, M.; Salleh, S.A. Numerical study on effect of urban heating on local climate during calm inter-monsoon period in greater Kuala Lumpur, Malaysia. Urban Clim. 2017, 20, 228–250. [Google Scholar] [CrossRef]
  23. Wang, K.; Aktas, Y.D.; Stocker, J.; Carruthers, D.; Hunt, J.; Malki-Epshtein, L. Urban heat island modelling of a tropical city: Case of Kuala Lumpur. Geosci. Lett. 2019, 6, 4. [Google Scholar] [CrossRef]
  24. Ahmad, S.; Hashim, N.M. Effects of Soil Moisture on Urban Heat Island Occurrences: Case of Selangor, Malaysia. Humanit. Soc. Sci. J. 2007, 2, 132–138. [Google Scholar]
  25. Yusuf, Y.A.; Pradhan, B.; Idrees, M.O. Spatio-temporal Assessment of Urban Heat Island Effects in Kuala Lumpur Metropolitan City Using Landsat Images. J. Indian Soc. Remote Sens. 2014, 42, 829–837. [Google Scholar] [CrossRef]
  26. Rahaman, S.; Jahangir, S.; Haque, S.; Chen, R.; Kumar, P. Spatio-temporal changes of green spaces and their impact on urban environment of Mumbai, India. Environ. Dev. Sustain. 2021, 23, 6481–6501. [Google Scholar] [CrossRef]
  27. Tso, C.; Chan, B.; Hashim, M.A. An improvement to the basic energy balance model for urban thermal environment analysis. Energy Build. 1990, 14, 143–152. [Google Scholar] [CrossRef]
  28. Tso, C. A survey of urban heat island studies in two tropical cities. Atmos. Environ. 1996, 30, 507–519. [Google Scholar] [CrossRef]
  29. Elsayed, I.S.M. Mitigation of the urban heat island of the city of Kuala Lumpur, Malaysia. Middle East J. Sci. Res. 2012, 11, 1602–1613. [Google Scholar] [CrossRef]
  30. Elsayed, I.S.M. International, the Effects of Urbanization on the Intensity of the Urban Heat Island: A Case Study on the City of Kuala Lumpur. Ph.D. Thesis, International Islamic University, Ampang Jaya, Malaysia, 2006. [Google Scholar]
  31. Elsayed, I.S.M. The Effects of Urbanization on the Intensity of the Urban Heat Island of Kuala Lumpur City. Ph.D. Thesis, International Islamic University, Ampang Jaya, Malaysia, 2012. [Google Scholar]
  32. Elsayed, I.S.M. Effects of Population Density and Land Management on the Intensity of Urban Heat Islands: A Case Study on the City of Kuala Lumpur, Malaysia. Appl. Geogr. Inf. Syst. 2012, 8, 267–283. [Google Scholar] [CrossRef]
  33. Elsayed, I.S.M. A Study on the Urban Heat Island of the City of Kuala Lumpur, Malaysia. J. King Abdulaziz Univ. Meteorol. Environ. Arid Land Agric. Sci. 2012, 23, 121–134. [Google Scholar] [CrossRef]
  34. Wong, N.H.; Yu, C. Study of green areas and urban heat island in a tropical city. Habitat Int. 2005, 29, 547–558. [Google Scholar] [CrossRef]
  35. Kotharkar, R.; Ramesh, A.; Bagade, A. Urban Heat Island studies in South Asia: A critical review. Urban Clim. 2018, 24, 1011–1026. [Google Scholar] [CrossRef]
  36. Salleh, S.A.; Abd.Latif, Z.; Mohd, W.M.N.W.; Chan, A. Factors Contributing to the Formation of an Urban Heat Island in Putrajaya, Malaysia. Procedia Soc. Behav. Sci. 2013, 105, 840–850. [Google Scholar] [CrossRef]
  37. Morris, K.I.; Salleh, S.A.; Chan, A.; Ooi, M.C.G.; Abakr, Y.A.; Oozeer, M.Y.; Duda, M. Computational study of urban heat island of Putrajaya, Malaysia. Sustain. Cities Soc. 2015, 19, 359–372. [Google Scholar] [CrossRef]
  38. Amanollahi, J.; Tzanis, C.; Ramli, M.F.; Abdullah, A.M. Urban heat evolution in a tropical area utilizing Landsat imagery. Atmos. Res. 2016, 167, 175–182. [Google Scholar] [CrossRef]
  39. Shaharuddin, A.; Noorazuan, M.H.; Takeuchi, W.; Noraziah, A. The effects of Urban Heat Islands on Human Comfort: A case of Klang Valley Malaysia. Glob. J. Adv. Pure Appl. Sci. 2014, 2, 268–293. [Google Scholar]
  40. Amorim, M.C.D.C.T.; Dubreuil, V. Intensity of Urban Heat Islands in Tropical and Temperate Climates. Climate 2017, 5, 91. [Google Scholar] [CrossRef]
  41. Molina, L.E.T.; Morales, S.; Carrión, L.F. Urban Heat Island Effects in Tropical Climate. In Vortex Dynamics Theories and Applications; Harun, Z., Ed.; IntechOpen: London, UK, 2020. [Google Scholar] [CrossRef]
  42. Bi, Z. Applications—Heat Transfer Problems. In Finite Element Analysis Applications; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar] [CrossRef]
  43. Dahoo, P.R.; Khettab, M.; Chong, C.; Girard, A.; Pougnet, P. Impact of Voids in Interconnection Materials. In Embedded Mechatronic Systems 2; Elsevier: Amsterdam, The Netherlands, 2015; pp. 79–106. [Google Scholar] [CrossRef]
  44. Šadauskiene, J.; Buska, A.; Burlingis, A.; Bliudzius, R.; Gailius, A. The Effect of Vertical Air Gaps to Thermal Transmittance of Horizontal Thermal Insulating Layer. J. Civ. Eng. Manag. 2009, 15, 309–315. [Google Scholar] [CrossRef]
  45. Alyasari, H.I.; Ameen, R.F.M.; Altaweel, M.D. Thermal Performance through the Use of Radiant Barrier and Phase Change Material in Concrete Flat Roofs. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2018. [Google Scholar] [CrossRef]
  46. Mohamed, H.; Chang, J.D.; Alshayeb, M. Effectiveness of High Reflective Roofs in Minimizing Energy Consumption in Residential Buildings in Iraq. Procedia Eng. 2015, 118, 879–885. [Google Scholar] [CrossRef]
  47. Asadi, S.; Hassan, M.M. Evaluation of the thermal performance of a roof-mounted radiant barrier in residential buildings: Experimental study. J. Build. Phys. 2013, 38, 66–80. [Google Scholar] [CrossRef]
  48. Michels, C.; Lamberts, R.; Guths, S. Evaluation of heat flux reduction provided by the use of radiant barriers in clay tile roofs. Energy Build. 2008, 40, 445–451. [Google Scholar] [CrossRef]
  49. Zhang, Y.; Long, E.; Li, Y.; Li, P. Solar radiation reflective coating material on building envelopes: Heat transfer analysis and cooling energy saving. Energy Explor. Exploit. 2017, 35, 748–766. [Google Scholar] [CrossRef]
  50. Field, C.B.; Barros, V.R.; Mastrandrea, M.D.; Mach, K.J.; Abdrabo, M.K.; Adger, N.; Yohe, G.W. Summary for policymakers. In Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2014; pp. 1–32. [Google Scholar]
  51. Capozzoli, A.; Gorrino, A.; Corrado, V. A building thermal bridges sensitivity analysis. Appl. Energy 2013, 107, 229–243. [Google Scholar] [CrossRef]
  52. Kontoleon, K.; Eumorfopoulou, E. The influence of wall orientation and exterior surface solar absorptivity on time lag and decrement factor in the Greek region. Renew. Energy 2008, 33, 1652–1664. [Google Scholar] [CrossRef]
  53. Ruivo, C.R.; Ferreira, P.M.; Vaz, D.C. On the error of calculation of heat gains through walls by methods using constant decrement factor and time lag values. Energy Build. 2013, 60, 252–261. [Google Scholar] [CrossRef]
  54. Yumrutaş, R.; Ünsal, M.; Kanoğlu, M. Periodic solution of transient heat flow through multilayer walls and flat roofs by complex finite Fourier transform technique. Build. Environ. 2005, 40, 1117–1125. [Google Scholar] [CrossRef]
  55. Lee, Y.; Chua, N.; Amran, M.; Lee, Y.Y.; Kueh, A.H.; Fediuk, R.; Vatin, N.; Vasilev, Y. Thermal Performance of Structural Lightweight Concrete Composites for Potential Energy Saving. Crystals 2021, 11, 461. [Google Scholar] [CrossRef]
  56. Makul, N.; Fediuk, R.; Amran, M.; Zeyad, A.; Murali, G.; Vatin, N.; Klyuev, S.; Ozbakkaloglu, T.; Vasilev, Y. Use of Recycled Concrete Aggregates in Production of Green Cement-Based Concrete Composites: A Review. Crystals 2021, 11, 232. [Google Scholar] [CrossRef]
  57. Tolstoy, A.; Lesovik, V.; Fediuk, R.; Amran, M.; Gunasekaran, M.; Vatin, N.; Vasilev, Y. Production of Greener High-Strength Concrete Using Russian Quartz Sandstone Mine Waste Aggregates. Materials 2020, 13, 5575. [Google Scholar] [CrossRef] [PubMed]
  58. Amran, Y.M.; Alyousef, R.; Alabduljabbar, H.; Alaskar, A.; Alrshoudi, F. Properties and water penetration of structural concrete wrapped with CFRP. Results Eng. 2020, 5, 100094. [Google Scholar] [CrossRef]
  59. Amran, M.; Murali, G.; Khalid, N.H.A.; Fediuk, R.; Ozbakkaloglu, T.; Lee, Y.H.; Haruna, S.; Lee, Y.Y. Slag uses in making an ecofriendly and sustainable concrete: A review. Constr. Build. Mater. 2021, 272, 121942. [Google Scholar] [CrossRef]
  60. Sabapathy, L.; Mohammed, B.S.; Al-Fakih, A.; A Wahab, M.M.; Liew, M.S.; Amran, Y.H.M. Acid and Sulphate Attacks on a Rubberized Engineered Cementitious Composite Containing Graphene Oxide. Materials 2020, 13, 3125. [Google Scholar] [CrossRef] [PubMed]
  61. Amran, M.; Debbarma, S.; Ozbakkaloglu, T. Fly ash-based eco-friendly geopolymer concrete: A critical review of the long-term durability properties. Constr. Build. Mater. 2021, 270, 121857. [Google Scholar] [CrossRef]
  62. Amran, Y.M.; Soto, M.G.; Alyousef, R.; El-Zeadani, M.; Alabduljabbar, H.; Aune, V. Performance investigation of high-proportion Saudi-fly-ash-based concrete. Results Eng. 2020, 6, 100118. [Google Scholar] [CrossRef]
  63. Amran, M.; Murali, G.; Fediuk, R.; Vatin, N.; Vasilev, Y.; Abdelgader, H. Palm Oil Fuel Ash-Based Eco-Efficient Concrete: A Critical Review of the Short-Term Properties. Materials 2021, 14, 332. [Google Scholar] [CrossRef]
  64. Zeyad, A.M.; Johari, M.A.M.; Alharbi, Y.R.; Abadel, A.A.; Amran, Y.M.; Tayeh, B.A.; Abutaleb, A. Influence of steam curing regimes on the properties of ultrafine POFA-based high-strength green concrete. J. Build. Eng. 2021, 38, 102204. [Google Scholar] [CrossRef]
  65. Amran, Y.M.; Alyousef, R.; Alabduljabbar, H.; El-Zeadani, M. Clean production and properties of geopolymer concrete; A review. J. Clean. Prod. 2020, 251, 119679. [Google Scholar] [CrossRef]
  66. Siddika, A.; Amin, R.; Rayhan, A.; Islam, S.; Al Mamun, A.; Alyousef, R.; Amran, Y.H.M. Performance of sustainable green concrete incorporated with fly ash, rice husk ash, and stone dust. Acta Polytech. 2021, 61, 279–291. [Google Scholar] [CrossRef]
  67. Gomes, M.G.; Flores-Colen, I.; Manga, L.; Soares, A.; de Brito, J. The influence of moisture content on the thermal conductivity of external thermal mortars. Constr. Build. Mater. 2017, 135, 279–286. [Google Scholar] [CrossRef]
  68. Taoukil, D.; El Bouardi, A.; Sick, F.; Mimet, A.; Ezbakhe, H.; Ajzoul, T. Moisture content influence on the thermal conductivity and diffusivity of wood-concrete composite. Constr. Build. Mater. 2013, 48, 104–115. [Google Scholar] [CrossRef]
  69. Mugahed Amran, Y.H. Determination of Structural Behavior of Precast Foamed Concrete Sandwich Panel. Ph.D. Thesis, Universiti Putra Malaysia (UPM), Ampang Jaya, Malaysia, 2016. [Google Scholar]
  70. Amran, Y.H.M.; Alyousef, R.; Alabduljabbar, H.; Alrshoudi, F.; Rashid, R.S.M. Influence of slenderness ratio on the structural performance of lightweight foam concrete composite panel. Case Stud. Constr. Mater. 2019, 10, e00226. [Google Scholar] [CrossRef]
  71. Amran, Y.H.M.; Rashid, R.S.M.; Hejazi, F.; Ali, A.A.A.; Safiee, N.A.; Bida, S.M. Structural Performance of Precast Foamed Concrete Sandwich Panel Subjected to Axial Load. KSCE J. Civ. Eng. 2017, 22, 1179–1192. [Google Scholar] [CrossRef]
  72. Amran, Y.M.; Rashid, R.S.; Hejazi, F.; Safiee, N.A.; Ali, A.A. Response of precast foamed concrete sandwich panels to flexural loading. J. Build. Eng. 2016, 7, 143–158. [Google Scholar] [CrossRef]
  73. Saheed, S.; Aziz, F.; Amran, M.; Vatin, N.; Fediuk, R.; Ozbakkaloglu, T.; Murali, G.; Mosaberpanah, M. Structural Performance of Shear Loaded Precast EPS-Foam Concrete Half-Shaped Slabs. Sustainability 2020, 12, 9679. [Google Scholar] [CrossRef]
  74. Rudenko, A.; Biryukov, A.; Kerzhentsev, O.; Fediuk, R.; Vatin, N.; Vasilev, Y.; Klyuev, S.; Amran, M.; Szelag, M. Nano- and Micro-Modification of Building Reinforcing Bars of Various Types. Crystals 2021, 11, 323. [Google Scholar] [CrossRef]
  75. Pásztory, Z.; Horváth, T.; Glass, S.V.; Zelinka, S. Experimental investigation of the influence of temperature on thermal conductivity of multilayer reflective thermal insulation. Energy Build. 2018, 174, 26–30. [Google Scholar] [CrossRef]
  76. Vosteen, H.-D.; Schellschmidt, R. Influence of temperature on thermal conductivity, thermal capacity and thermal diffusivity for different types of rock. Phys. Chem. Earth Parts A B C 2003, 28, 499–509. [Google Scholar] [CrossRef]
  77. Fediuk, R.; Amran, M.; Vatin, N.; Vasilev, Y.; Lesovik, V.; Ozbakkaloglu, T. Acoustic Properties of Innovative Concretes: A Review. Materials 2021, 14, 398. [Google Scholar] [CrossRef] [PubMed]
  78. Fedyuk, R.; Baranov, A.; Amran, Y.M. Effect of porous structure on sound absorption of cellular concrete. Constr. Mater. Prod. 2020, 3, 5–18. [Google Scholar] [CrossRef]
  79. Amran, Y.H.M. Influence of structural parameters on the properties of fibred-foamed concrete. Innov. Infrastruct. Solut. 2020, 5, 1–18. [Google Scholar] [CrossRef]
  80. Lesovik, V.; Chernysheva, N.; Fediuk, R.; Amran, M.; Murali, G.; de Azevedo, A.R. Optimization of fresh properties and durability of the green gypsum-cement paste. Constr. Build. Mater. 2021, 287, 123035. [Google Scholar] [CrossRef]
  81. Asadi, I.; Shafigh, P.; Bin Abu Hassan, Z.F.; Mahyuddin, N.B. Thermal conductivity of concrete—A review. J. Build. Eng. 2018, 20, 81–93. [Google Scholar] [CrossRef]
  82. Gandage, A.S.; Rao, V.V.; Sivakumar, M.; Vasan, A.; Venu, M.; Yaswanth, A. Effect of Perlite on Thermal Conductivity of Self Compacting Concrete. Procedia Soc. Behav. Sci. 2013, 104, 188–197. [Google Scholar] [CrossRef]
  83. Alengaram, U.J.; Al Muhit, B.A.; bin Jumaat, M.Z.; Jing, M.L.Y. A comparison of the thermal conductivity of oil palm shell foamed concrete with conventional materials. Mater. Des. 2013, 51, 522–529. [Google Scholar] [CrossRef]
  84. Liu, M.Y.J.; Alengaram, U.J.; Jumaat, M.Z.; Mo, K.H. Evaluation of thermal conductivity, mechanical and transport properties of lightweight aggregate foamed geopolymer concrete. Energy Build. 2014, 72, 238–245. [Google Scholar] [CrossRef]
  85. Sayadi, A.A.; Tapia, J.V.; Neitzert, T.R.; Clifton, G.C. Effects of expanded polystyrene (EPS) particles on fire resistance, thermal conductivity and compressive strength of foamed concrete. Constr. Build. Mater. 2016, 112, 716–724. [Google Scholar] [CrossRef]
  86. Yun, T.S.; Jeong, Y.J.; Han, T.-S.; Youm, K.-S. Evaluation of thermal conductivity for thermally insulated concretes. Energy Build. 2013, 61, 125–132. [Google Scholar] [CrossRef]
  87. Topçu, I.B.; Uygunoğlu, T. Properties of autoclaved lightweight aggregate concrete. Build. Environ. 2007, 42, 4108–4116. [Google Scholar] [CrossRef]
  88. Demirboğa, R.; Kan, A. Thermal conductivity and shrinkage properties of modified waste polystyrene aggregate concretes. Constr. Build. Mater. 2012, 35, 730–734. [Google Scholar] [CrossRef]
  89. Demirboǧa, R.; Gül, R. Thermal conductivity and compressive strength of expanded perlite aggregate concrete with mineral admixtures. Energy Build. 2003, 35, 1155–1159. [Google Scholar] [CrossRef]
  90. Jin, H.-Q.; Yao, X.-L.; Fan, L.-W.; Xu, X.; Yu, Z.-T. Experimental determination and fractal modeling of the effective thermal conductivity of autoclaved aerated concrete: Effects of moisture content. Int. J. Heat Mass Transf. 2016, 92, 589–602. [Google Scholar] [CrossRef]
  91. Cavanaugh, K.; Speck, J.F. Guide to Thermal Properties of Concrete and Masonry Systems Reported; ACI Committee 122, Concrete; ACI Committee: Farmington Hills, MI, USA, 2002. [Google Scholar]
  92. ASTM:C1363-97. Standard Test Method for the Thermal Performance of Building Assemblies; Means of a Hot Box Apparatus 1, C1363-97; ASTM: West Conshohocken, PA, USA, 1997. [Google Scholar]
  93. Belkharchouche, D.; Chaker, A. Effects of moisture on thermal conductivity of the lightened construction material. Int. J. Hydrog. Energy 2016, 41, 7119–7125. [Google Scholar] [CrossRef]
  94. ASTM International. Standard Test Method for Steady-State Heat Flux Measurements and Thermal Transmission Properties by Means of the Guarded-Hot-Plate; ASTM Stand. C177-13; ASTM: West Conshohocken, PA, USA, 2013. [Google Scholar] [CrossRef]
  95. Maglić, K.D.; Cezairliyan, A.; Peletsky, V.E. Compendium of Thermophysical Property Measurement Methods: Volume 2 Recommended Measurement Techniques and Practices; Springer: Berlin/Heidelberg, Germany, 1992. [Google Scholar] [CrossRef]
  96. Wongkeo, W.; Thongsanitgarn, P.; Pimraksa, K.; Chaipanich, A. Compressive strength, flexural strength and thermal conductivity of autoclaved concrete block made using bottom ash as cement replacement materials. Mater. Des. 2012, 35, 434–439. [Google Scholar] [CrossRef]
  97. Collet, F.; Pretot, S. Thermal conductivity of hemp concretes: Variation with formulation, density and water content. Constr. Build. Mater. 2014, 65, 612–619. [Google Scholar] [CrossRef]
  98. Bederina, M.; Marmoret, L.; Mezreb, K.; Khenfer, M.; Bali, A.; Quéneudec, M. Effect of the addition of wood shavings on thermal conductivity of sand concretes: Experimental study and modelling. Constr. Build. Mater. 2007, 21, 662–668. [Google Scholar] [CrossRef]
  99. Fathipour, R.; Hadidi, A. Analytical solution for the study of time lag and decrement factor for building walls in climate of Iran. Energy 2017, 134, 167–180. [Google Scholar] [CrossRef]
  100. Baggs, D. Thermal Mass and Its Role in Building Comfort and Energy Efficiency, Technical Guide; EcoSpecifier: Cannon Hill, Australia, 2013. [Google Scholar]
  101. Marshall, A. The thermal properties of concrete. Build. Sci. 1972, 7, 167–174. [Google Scholar] [CrossRef]
  102. Kodur, V. Properties of Concrete at Elevated Temperatures. ISRN Civ. Eng. 2014, 2014, 468510. [Google Scholar] [CrossRef]
  103. Ahuja, A.; Mosalam, K.M. Evaluating energy consumption saving from translucent concrete building envelope. Energy Build. 2017, 153, 448–460. [Google Scholar] [CrossRef]
  104. Gan, W.; Cao, Y.; Jiang, W.; Li, L.; Li, X. Energy-Saving Design of Building Envelope Based on Multiparameter Optimization. Math. Probl. Eng. 2019, 2019, 5261869. [Google Scholar] [CrossRef]
  105. Shehadi, M. Energy Consumption Optimization Measures for Buildings in the Midwest Regions of USA. Buildings 2018, 8, 170. [Google Scholar] [CrossRef]
  106. Ghabra, N.; Rodrigues, L.; Oldfield, P. The impact of the building envelope on the energy efficiency of residential tall buildings in Saudi Arabia. Int. J. Low Carbon Technol. 2017, 12, 411–419. [Google Scholar] [CrossRef]
  107. Jannat, N.; Hussien, A.; Abdullah, B.; Cotgrave, A. A Comparative Simulation Study of the Thermal Performances of the Building Envelope Wall Materials in the Tropics. Sustainability 2020, 12, 4892. [Google Scholar] [CrossRef]
  108. Bentz, D.P.; Turpin, R. Potential applications of phase change materials in concrete technology. Cem. Concr. Compos. 2007, 29, 527–532. [Google Scholar] [CrossRef]
  109. Adesina, A. Use of phase change materials in concrete: Current challenges. Renew. Energy Environ. Sustain. 2019, 4, 9. [Google Scholar] [CrossRef]
  110. Drissi, S.; Ling, T.-C.; Mo, K.H. Thermal performance of a solar energy storage concrete panel incorporating phase change material aggregates developed for thermal regulation in buildings. Renew. Energy 2020, 160, 817–829. [Google Scholar] [CrossRef]
  111. Amran, Y.H.M.; Alyousef, R.; Alabduljabbar, H.; Khudhair, M.H.R.; Hejazi, F.; Alaskar, A.; Alrshoudi, F.; Siddika, A. Performance properties of structural fibred-foamed concrete. Results Eng. 2020, 5, 100092. [Google Scholar] [CrossRef]
  112. Tay, L.T.; Lee, Y.Y.; Lee, Y.H.; Kueh, A.B.H. Compressive and Flexural Strengths of Mortar with Silica Aerogel Powder. In Proceedings of the International Conference on Civil, Offshore and Environmental Engineering; Springer: Singapore, 2021. [Google Scholar] [CrossRef]
  113. Lee, Y.H.; Lim, M.H.; Lee, Y.L.; Lee, Y.Y.; Tan, C.S.; Mohammad, S.; Ma, C.K. Compressive strength of lightweight foamed concrete with charcoal as a sand replacement. Indian J. Eng. Mater. Sci. 2018, 25, 98–108. [Google Scholar] [CrossRef]
  114. Do, H.; Cetin, K.S. Residential Building Energy Consumption: A Review of Energy Data Availability, Characteristics, and Energy Performance Prediction Methods. Curr. Sustain. Energy Rep. 2018, 5, 76–85. [Google Scholar] [CrossRef]
  115. Talebi, H.R.; Kayan, B.A.; Asadi, I.; Hassan, Z.F.B.A. Investigation of thermal properties of normal weight concrete for different strength classes. J. Environ. Treat. Tech. 2020, 8, 908–914. [Google Scholar]
  116. Edwards, R.E.; New, J.; Parker, L.E. Predicting future hourly residential electrical consumption: A machine learning case study. Energy Build. 2012, 49, 591–603. [Google Scholar] [CrossRef]
  117. Catalina, T.; Iordache, V.; Caracaleanu, B. Multiple regression model for fast prediction of the heating energy demand. Energy Build. 2013, 57, 302–312. [Google Scholar] [CrossRef]
  118. Jetcheva, J.G.; Majidpour, M.; Chen, W.-P. Neural network model ensembles for building-level electricity load forecasts. Energy Build. 2014, 84, 214–223. [Google Scholar] [CrossRef]
  119. Farzana, S.; Liu, M.; Baldwin, A.; Hossain, U. Multi-model prediction and simulation of residential building energy in urban areas of Chongqing, South West China. Energy Build. 2014, 81, 161–169. [Google Scholar] [CrossRef]
  120. Naji, S.; Keivani, A.; Shamshirband, S.; Alengaram, U.J.; Jumaat, M.Z.; Mansor, Z.; Lee, M. Estimating building energy consumption using extreme learning machine method. Energy 2016, 97, 506–516. [Google Scholar] [CrossRef]
  121. Li, K.; Hu, C.; Liu, G.; Xue, W. Building’s electricity consumption prediction using optimized artificial neural networks and principal component analysis. Energy Build. 2015, 108, 106–113. [Google Scholar] [CrossRef]
  122. Zhang, Y.; O’Neill, Z.; Dong, B.; Augenbroe, G. Comparisons of inverse modeling approaches for predicting building energy performance. Build. Environ. 2015, 86, 177–190. [Google Scholar] [CrossRef]
  123. Paulus, M.T.; Claridge, D.E.; Culp, C. Algorithm for automating the selection of a temperature dependent change point model. Energy Build. 2015, 87, 95–104. [Google Scholar] [CrossRef]
  124. Perez, K.X.; Cetin, K.; Baldea, M.; Edgar, T.F. Development and analysis of residential change-point models from smart meter data. Energy Build. 2017, 139, 351–359. [Google Scholar] [CrossRef]
  125. Castelli, M.; Trujillo, L.; Vanneschi, L.; Popovič, A. Prediction of energy performance of residential buildings: A genetic programming approach. Energy Build. 2015, 102, 67–74. [Google Scholar] [CrossRef]
  126. Kim, K.H.; Haberl, J.S. Development of methodology for calibrated simulation in single-family residential buildings using three-parameter change-point regression model. Energy Build. 2015, 99, 140–152. [Google Scholar] [CrossRef]
  127. Ahmed, M.S.; Mohamed, A.; Homod, R.Z.; Shareef, H. Hybrid LSA-ANN Based Home Energy Management Scheduling Controller for Residential Demand Response Strategy. Energies 2016, 9, 716. [Google Scholar] [CrossRef]
  128. Li, R.; Wang, Z.; Gu, C.; Li, F.; Wu, H. A novel time-of-use tariff design based on Gaussian Mixture Model. Appl. Energy 2016, 162, 1530–1536. [Google Scholar] [CrossRef]
  129. Melzi, F.N.; Same, A.; Zayani, M.H.; Oukhellou, L. A Dedicated Mixture Model for Clustering Smart Meter Data: Identification and Analysis of Electricity Consumption Behaviors. Energies 2017, 10, 1446. [Google Scholar] [CrossRef]
  130. Jain, R.K.; Smith, K.M.; Culligan, P.J.; Taylor, J.E. Forecasting energy consumption of multi-family residential buildings using support vector regression: Investigating the impact of temporal and spatial monitoring granularity on performance accuracy. Appl. Energy 2014, 123, 168–178. [Google Scholar] [CrossRef]
  131. Habsya, C.; Diharjo, K.; Setyono, P.; Satwiko, P. Physical, mechanical and thermal properties of lightweight foamed concrete with fly ash. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2018. [Google Scholar] [CrossRef]
  132. Mydin, M.A.O. Effective thermal conductivity of foamcrete of different densities. Concr. Res. Lett. 2011, 2, 181–189. [Google Scholar]
  133. Bumanis, G.; Bajare, D.; Korjakins, A. Mechanical and Thermal Properties of Lightweight Concrete Made from Expanded Glass. J. Sustain. Arch. Civ. Eng. 2013, 2, 19–25. [Google Scholar] [CrossRef]
  134. Amran, Y.M.; Rashid, R.S.; Hejazi, F.; Safiee, N.A.; Ali, A.A. Structural behavior of laterally loaded precast foamed concrete sandwich panel. Int. J. Civ. Environ. Struct. Constr. Archit. Eng. 2016, 10, 255–263. [Google Scholar]
  135. Chua, N. Thermal Performance of Lightweight Concrete; Curtin University Malaysia: Sarawak, Malaysia, 2020. [Google Scholar]
  136. Lee, Y.Y.; Lee, Y.H.; Mohammad, S.; Shek, P.N.; Ma, C.K. Thermal characteristics of a residential house in a new township in Johor Bahru. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2017. [Google Scholar] [CrossRef]
Figure 1. Energy utilisation in the US, UK, and Sri Lanka [2].
Figure 1. Energy utilisation in the US, UK, and Sri Lanka [2].
Sustainability 13 11957 g001
Figure 2. Experimental setup of the preliminary study.
Figure 2. Experimental setup of the preliminary study.
Sustainability 13 11957 g002
Figure 3. Comparison of surface temperatures for OPS & 50-QD and conventional concrete under three hours of natural sunlight exposure.
Figure 3. Comparison of surface temperatures for OPS & 50-QD and conventional concrete under three hours of natural sunlight exposure.
Sustainability 13 11957 g003
Figure 4. Wall inner surface temperature estimation.
Figure 4. Wall inner surface temperature estimation.
Sustainability 13 11957 g004
Table 1. Tropical climate.
Table 1. Tropical climate.
Tropical Rainforest [14]Tropical Monsoon [15]Savanna [16]
Location North and south latitudinal ranges of 5–10 degrees from the equator Between the latitude of 10 degrees north and the Tropic of CancerBetween 10° and 20° north–south latitude
Average temperature 21–30 °C27.1 °C ± 3.6 °C20–30 °C
Total precipitation 2540 mm 3409.2 mm700–1000 mm
Table 2. Summary of UHI effects in tropics.
Table 2. Summary of UHI effects in tropics.
Ref. Country AreaYear of Assessment Method Temperature Difference, °C
[34]Singapore SingaporeJuly–September, 2002Temperature traverse AT–4.01
[36]Malaysia Putrajaya1999–2009Satellite ST–6.75
[37]Malaysia Putrajaya2012 Numerical simulation AT–3.10
[38]MalaysiaKlang2000 & 2010 Satellite ST–21.50 & ST–22.70
[39]Malaysia Kuala Lumpur2008–2009 Satellite & GIS ST–14.50
[33]Malaysia Kuala Lumpur2004Temperature traverse & weather stationsAT–5.50
[40]Brazil Presidente Prudente2016Weather stationAT–2.40 (mean), 4.10 (monthly), 6.40 (hourly)
[41]USPuerto Rico2008–2018Weather station 2.06–3.04
AT = ambient temperature, ST = surface temperature.
Table 4. Thermal inertia of selected construction materials [100].
Table 4. Thermal inertia of selected construction materials [100].
MaterialDensity
kg/m3
Specific Heat,
kJ/kg.K
Thermal Mass,
kJ/m3.K
Time Lag
HourThickness, mm
Water 10004.1864186--
Concrete 22400.92020606.9250
AAC5001.1005507.0200
Brick 17000.92013605.0125
Sand stone 20000.9001800--
Compressed FC sheet 17000.9001530--
Earth wall (Adobe) 15500.83713009.2250
Rammed earth 20000.837167310.3250
Compressed earth blocks 20800.837174010.5250
Table 5. Study on energy performance in numerous areas.
Table 5. Study on energy performance in numerous areas.
Ref. Location Concrete Property/Building Model Energy ConsumptionRemark
[103]Berkeley Translucent concrete Reduction of 18% HVAC energy in office room compared to a completely sunlight-deprived room Reduce 50% of lighting energy
[104]Chengdu, Sichuan Normal concrete, different parametersTwo sets of solutions were selected from 1000 available solutions based on low energy consumption on HVAC and lighting systems. Case study on the library building
[105]Midwest regions of USA20 cm hollow concrete block (HCB) when combined with fluorescent lights and double pane heat-absorbing glazing for windowsReduction of cooling loads Higher initial cost
[106]Saudi Arabia Thermal block and 32 mm glazing 77% difference for energy consumptionLow shading coefficient
[107]Tropics Different shape/space factors Energy consumption increases with surface area and volume of the buildingDecrement factor and time lag affect the building thermal performance
Table 8. Cooling load of several wall materials without covers.
Table 8. Cooling load of several wall materials without covers.
Mix Design U-Value, W/m2KCooling Load, W/m2 Remarks
Structural lightweight concrete with expanded clay and oil palm shells [55,135]6.62–7.18 55.5–60.16 New proposed equation:
U c = 0.904 × σ c 0.7
Uc = transmittance value and σc = concrete compressive strength, applicable to concrete strength ≥ 17 MPa
Normal structural concrete 21.96175.68
Clay brick 7.6961.52100 mm thickness
Concrete block 11.1188.88100 mm thickness
Non-structural foamed concrete [83]3.92–5.29 31.36–42.32102 mm thickness
Table 9. Concrete thermal behavior [135].
Table 9. Concrete thermal behavior [135].
SpecimenDensity, kg/m3Compressive Strength, MPaDecrement Factor (f)Time Lag (ϕ), min.Cooling Load, %
Conventional248056.10.94145100
U-POFA236055.30.91345100
G-POFA234052.40.89760100
25-LECA & 75-OPS178017.20.78612065
OPS & 50-QD188019.30.83410529
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, Y.H.; Amran, M.; Lee, Y.Y.; Kueh, A.B.H.; Kiew, S.F.; Fediuk, R.; Vatin, N.; Vasilev, Y. Thermal Behavior and Energy Efficiency of Modified Concretes in the Tropical Climate: A Systemic Review. Sustainability 2021, 13, 11957. https://doi.org/10.3390/su132111957

AMA Style

Lee YH, Amran M, Lee YY, Kueh ABH, Kiew SF, Fediuk R, Vatin N, Vasilev Y. Thermal Behavior and Energy Efficiency of Modified Concretes in the Tropical Climate: A Systemic Review. Sustainability. 2021; 13(21):11957. https://doi.org/10.3390/su132111957

Chicago/Turabian Style

Lee, Yeong Huei, Mugahed Amran, Yee Yong Lee, Ahmad Beng Hong Kueh, Siaw Fui Kiew, Roman Fediuk, Nikolai Vatin, and Yuriy Vasilev. 2021. "Thermal Behavior and Energy Efficiency of Modified Concretes in the Tropical Climate: A Systemic Review" Sustainability 13, no. 21: 11957. https://doi.org/10.3390/su132111957

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop