Next Article in Journal
Preparation and Properties of Environmentally Friendly Resin-Based Artificial Stones Fabricated from Ceramic Waste
Next Article in Special Issue
Renewable Energies and Architectural Heritage: Advanced Solutions and Future Perspectives
Previous Article in Journal
Cement Kiln Dust (CKD) as a Partial Substitute for Cement in Pozzolanic Concrete Blocks
Previous Article in Special Issue
Evaluation of Building Energy Savings Achievable with an Attached Bioclimatic Greenhouse: Parametric Analysis and Solar Gain Control Techniques
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Insulation Assessment, Emission Analysis, and Correlation Formulation for Indian Region

1
Faculty of Technology, Batman University, Batman 72100, Turkey
2
Computer Engineering Department of Engineering Faculty, Karamanoğlu Mehmetbey University, Karaman 70200, Turkey
3
Department of Mechanical Engineering, Dr B R Ambedkar National Institute of Technology, Jalandhar 144011, India
4
Structural Engineering and Construction Management Department, Future University in Egypt, Cairo 11835, Egypt
5
Faculty of Engineering, Department of Industrial Engineering and Management, Lucian Blaga University of Sibiu, 550024 Sibiu, Romania
6
CSIR-Central Building Research Institute, Roorkee 247667, India
7
Institute of Materials Engineering, Faculty of Engineering, University of Pannonia, 8200 Veszprem, Hungary
8
Savaria Institute of Technology, Faculty of Informatics, ELTE Eötvös Loránd University, 9700 Szombathely, Hungary
*
Author to whom correspondence should be addressed.
Buildings 2023, 13(2), 569; https://doi.org/10.3390/buildings13020569
Submission received: 14 January 2023 / Revised: 5 February 2023 / Accepted: 13 February 2023 / Published: 20 February 2023

Abstract

:
The current study depicts the effects of different insulation materials and fuel types on the cooling and heating performance of buildings situated in hot and dry, warm and humid, composite, and cold climatic conditions in India. Ten different locations chosen from diverse climatic regions were selected, and various potential parameters for expanded polystyrene and extruded polystyrene insulation materials were evaluated. Potential parameters, such as optimal insulation thickness, annual savings, and payback period, were computed for cooling and heating requirements and were found in the ranges of 0.0428–0.891 m, 10.83–19.19 $/m2, and 1.49–2.36 years for cooling, as well as 0.0063–0.1522 m, 0.29–55.92 $/m2, and 0.95–6.52 years for heating, respectively. An emission analysis was also carried out for the estimation of greenhouse gas (GHG) emissions by the engagement of optimal insulation thickness for heating. The GHG emissions from natural gas, coal, and diesel by the employment of various insulating materials were found in the ranges of 5.39–11.28, 9.47–32.68, and 2.26–4.51 kg/m2-year, respectively. A correlation formulation (power) for optimal insulation thickness was also carried out. For checking the preciseness of the developed mathematical models, statistical tools were utilized, and their obtained values in the satisfactory range signified the accurateness of the developed models.

1. Introduction

Energy is vital to human existence and advancement. The increasing demand for energy is driven by various factors, such as the development of new technologies, the impact of climate change, and population growth. However, the majority of the world’s energy needs are currently met by fossil fuels, which are nonrenewable and finite resources. As a result, these energy sources are becoming increasingly scarce, with depletion happening at an alarming rate. This has led to the urgent need to find sustainable and renewable alternatives to fossil fuels in order to meet the growing energy demands and preserve the planet for future generations [1,2,3]. The use of thermal energy systems, such as those found in homes and businesses for temperature regulation (heating and cooling), can often result in a loss in thermal energy to the surrounding environment. To counter this, insulation materials are commonly applied to the exterior walls, windows, ceilings, and floors of these buildings. The use of insulation can greatly reduce energy loss and lead to significant energy savings. By minimizing energy loss in this way, the overall carbon footprint of a building can also be reduced [4,5]. The optimum insulation thickness for the exterior walls of a building can depend on several factors, including the climate, location, and intended use of the building. The R-value, which is a measure of thermal resistance, is typically used to determine appropriate insulation thickness for exterior walls. As a general rule, the R-value required for exterior walls in a building depends on the insulation materials, climate zone, overall cost, and building type [6,7]. Nyers et al. [8] developed an investment-saving method for determining the optimal insulation thickness of external walls in Serbia. A mathematical model was developed considering the steady-state one-dimensional heat conduction of walls made up of bricks and polystyrene. The overall savings and payback terms were evaluated at optimal insulation thickness and found to be 8.5 $/m2 and 1.22 years, respectively. Liu et al. [9] computed the optimal thickness of insulation for external walls in China’s cold winter and hot summer zones. A mathematical model that incorporated heat and moisture transfer was developed and used to estimate the annual energy consumption. Two different types of insulation materials, viz., expanded polystyrene (EPS) and extruded polystyrene (XPS), were compared with lifecycle savings and payback period. The optimal insulation thicknesses for XPS and EPS were found at 0.053–0.069 m and 0.081–0.105 m, respectively. The maximum lifecycle saving and the payback period were found at 16.6–28.5 $/m2 and 1.89–2.56 years, respectively. The results also revealed that the EPS material performed better than XPS, with more lifecycle savings and a shorter payback period. In another study, Yuan et al. [10] considered the effects of both insulation thickness and reflectivity on outer walls in different climatic zones of Japan. The reflectivity and the insulation thickness were supposed to be varied between 0.1 and 0.8 mm and 10 and 100 mm, respectively. The results showed that a low value of reflectivity and thick insulation were preferable for cold regions and vice versa for hot areas. The effect of insulation thickness on enjoinment was studied by Dombayci [11] in Turkey for cold seasons. A one-dimensional steady-state heat conduction model was developed for a sandwiched wall made up of the materials in the following order: plaster, brick, insulation, brick, and plaster. The results revealed that, at optimum insulation thickness, the energy consumption decreased by 46.6%, and the CO2 and SO2 emissions were reduced by 41.53%. Kaynakli et al. [12] evaluated the optimum insulation thickness for different applications of insulation material (internal, sandwiched, and external) by taking into account the indoor and outdoor conditions. The key factors on which the optimum thickness of insulation depended were found to be indoor temperature and indoor and outdoor relative humidity. The results indicated that, for low indoor and outdoor relative humidity (≤0.6), the type of insulation application did not affect insulation thickness. However, insulation application over the outer edge of an external wall performed better than all the other cases in the studied conditions. In a different study, Bolattürk [13] evaluated the optimum thickness of foam for sixteen cities from four different climate zones of Turkey. The fuels considered for heating and cooling in their study were coal, natural gas, diesel, LPG, and electricity. The results revealed that the optimum insulation thickness, energy saving, and life cost were in the ranges of 2–17 cm, 22–79%, and 1.3–4.5 years, respectively, depending on the different climate zones. The results also indicated that high values of optimum insulation thickness were evaluated for cold climatic regions with significantly less of a payback period. Mahlia et al. [14] evaluated the optimum insulation thickness for different materials in the climatic conditions of Malaysia and formed correlations as a function of the thermal conductivities of different insulating materials. The result revealed that a nonlinear mathematical model (polynomial) predicted the results most precisely. This correlation was xopt = a + bk + ck2, where a = 0.0818, b = −2.973, and c = 64.6, and “k” in the correlation was the thermal conductivity (W/m °C) of the insulating material. Sisman et al. [15] conducted a study and determined optimal insulation thickness for cities in four different climatic zones of Turkey for stone wool material for heating. These properties constituted equations that could be used to specify the optimum point based on the NDD variable for both external walls and roofs. The external wall equation was xopt = a × NDDb, where a = 0.001 and b = 0.7533.
India has diversified climatic zones, including hot and dry, warm and humid, composite, cold, and temperate. A few studies have been reported to estimate optimal insulation thicknesses in different climatic zones in India. Sundaram and Bhaskaran [16] optimized insulation thickness for five cities in India, selecting the warm and humid and composite climates. Three different insulating materials attached at the inner sides of outer walls were considered for the study, and various thermo-economic parameters, such as optimum insulation thickness, annual energy cost, annual electric energy consumption, and payback period, were determined. The results revealed that EPS insulation material performed more effectively than the other studied materials in terms of energy saving and payback period. Mishra et al. [17] computed the optimal insulation thickness for Dehradun, located in a cold climatic region of India. Two dissimilar insulation materials, EPS and XPS, were considered for three different walls of dissimilar materials. The heat loss was calculated by the degree day method, and the fuel assumed for heating was natural gas. The results revealed that the optimum insulation thickness and energy saving varied between 5.2 and 7.4 cm and 31.41 and 67.59 $/m2, respectively. Raza and Aggarwal [18] determined optimum insulation thickness using two different methods, viz., the degree day method and annual full load cooling hours operation. The insulation material was considered to be sandwiched between bricks, and three different insulation materials were investigated. The results revealed that the EPS insulation materials using LPG for heating were the most effective combination. In another study, Singh et al. [19] computed the optimal XPS thickness for various climates of India. The study concluded that, at a 25 °C indoor temperature for all the selected zones, the optimum insulation, annual savings, and payback period were found in the ranges of 0.015–0.031 m, 0.33–2.21 $/m2, and 3.9–6.7 years, respectively.
The literature survey clearly indicates the necessity of evaluating the optimal insulation thickness to help reduce the amount of energy needed to heat and cool buildings, thus decreasing the amount of greenhouse gas emissions associated with energy production. Therefore, in this study, optimal insulation thickness is evaluated based on the degree day method for ten locations selected from different climatic zones in India. EPS and XPS insulation materials are investigated, and various fuels, such as natural gas, coal, diesel, and electricity, are considered for heating and cooling. In 2015, the United Nations (UN) defined 17 sustainable development goals that address the most pressing global issues [20]. This study is directly related to Target 12 (Responsible Consumption and Production). It is also indirectly related to Target 7, Target 11, and Target 13. These targets frame pressing issues in the construction, energy, and manufacturing sectors that need to be addressed in the next quarter century.

2. Materials and Methods

Calculations were performed for cities in 10 different regions (Jaisalmer, Kota, Mumbai, Chennai, Bhubaneswar, New Delhi, Lucknow, Patna, Srinagar, and Shillong) with different climatic characteristics in Indian climate geography, as shown in Figure 1 [21,22]. Köppen–Geiger climatic classification was utilized to understand climatic attributes on a global scale. These climate groups were expressed in three letters. For Köppen classification, the first letter of these climate groups was as follows: equatorial region (A), arid region (B), hot temperate zone (C), snowy region (D), or polar region (E). The second letter in classification was the rainfall position of the region; the third letter was the region’s temperature [21]. Jaisalmer was classified as a tropical and subtropical desert climate (Bwh); Mumbai, Chennai, and Bhubaneswar were classified as tropical savanna climates (Aw); Kota and New Delhi were classified as mid-latitude steppe and desert climates (Bsh); Lucknow and Patna belonged to the monsoon-influenced humid subtropical climate classification (Cwa); Srinagar belonged to the cold desert climate classification (Bwk); and Shillong was classified as a tropical monsoon climate (Am). Aw-Cwa was mid-rainy, and Bwh-Bsh-Bwk was arid and less rainy [22]. Energy need was computed by accepting T ≤ 15 °C as the degree day value for heating (HDD) and T > 24 °C as the degree day value for cooling (CDD). Annual average values between 2007 and 2016 were used as heating and cooling day degree data; Table 1 shows the annual values of the regions. The data on HDD, CDD, and average temperature for the different locations were collected through the database of the JRC Photovoltaic Geographical Information System (PVGIS) [23]. Coal, natural gas, and diesel were utilized for heating in the lifecycle cost analysis (LCA), while electricity was used for cooling.

2.1. Building Wall Model

The largest heat losses in buildings are reported through walls, floors, roofs, and windows, with a total energy loss of nearly 15–35% [24]. Heat losses and gains arising from buildings vary by architecture, location, and structural materials. The highest loss rate in a building is due to exterior walls and, if optimum insulation is provided, 50% to 60% of the energy can be saved [25]. Therefore, insulation for external walls is an investment for maximum energy gain. Walls are made up of a single layer nowadays, but they may also be considered as a construction component made up of two or more layers and elements [26].
The optimal insulation thickness was found in this investigation by considering that heat losses only occurred through outer walls. The wall model, as shown in Figure 2, was composed of interior plaster, insulating material, concrete, and exterior plaster. Table 2 displays the properties of these wall components [27,28].
The overall heat transfer coefficient (U) was determined using Equation (1) [29]:
U = 1 R i + R w + R i z o + R o                     ( W / m 2 K )
In Equation (1), Ri represents the thermal resistance of the inner surface, and Ro represents the thermal resistance of the outer surface. Rw is the thermal resistance of the wall layers that are not insulated, and Rizo is the insulation material’s thermal resistance computed with Equation (2). In this equation, x is the insulation material’s thickness, and k is the insulation material’s thermal conductivity coefficient. The values of k were considered as 0.026 W/m·K and 0.035 W/m·K for XPS and EPS with costs (Ci) of 46.35 $/m3 and 105.94 $/m3, respectively [16].
R i z o = x k

2.2. Building Wall Heating Load

Windows, ceilings, floors, exterior walls, and air infiltration result in heat losses arising in buildings. Calculations were revised in this analysis by assuming that they came only from the outer wall. For the outer wall, the heat loss was computed using Equation (3) [30]:
q = U . Δ T
where ∆T is the temperature difference, and U is the total heat transfer coefficient. The number of degree days (NDD) and U were implemented to compute the yearly heat loss qA (W/m2) of the unit surface:
q A = ( 3600 × 24 ) . N D D . U    
Equation (5) provides the EA (J/m2-year) yearly energy required for heating and was computed by dividing qA by the system efficiency:
E A = 86400 . N D D . U η = 86400 . N D D ( R T W + R i z o ) . η
Equation (6) calculates the mfA (kg/m2-year) fuel amount consumed in a year:
m f A = 86400 . N D D   ( R T W + R i z o ) . H u . . η
Equation (7) calculates the annual energy cost CA,H ($/m2-year) utilized for heating a unit area:
C A , H = 86400 . H D D . C f ( R T W + R i z o ) . η . H u .
As shown in Table 3 [31], Hu is the fuel lower heat value, η is the the fuel efficiency, and Cf is the fuel price. Equation (8) computes the cost of cooling with respect to the amount of energy utilized:
C A , C = 86400 . C D D . C f ( R T W + R i z o ) .   C O P
Here, COP is the cooling system performance coefficient taken as 2.5 [32]. The values of Cf, Hu, and η are listed in Table 3.
Table 3. Considered fuel characteristics [31,33].
Table 3. Considered fuel characteristics [31,33].
FuelChemical EquationCfηHu
CoalC5.85H5.26O1.13S0.008N0.0770.16610 $/kg0.6521.113 × 106 J/kg
Natural gasC1.05H4O0.034N0.0220.1305 $/m30.9334.526 × 106 J/m3
DieselC7.3125 H10.407 O0.04 S0.026 N0.020.69981 $/kg0.8042.911 × 106 J/kg
Electricity-0.08 $/kWh0.993.5990 × 106 J/kWh

2.3. Optimal Insulation Analysis

The purpose was to reduce insulation costs by analyzing the optimal thickness of insulation desired for a building. The building’s overall heating cost was calculated using LCCA by combining the energy price, any insulation price, the future value factor (PWF), and time (N) [34,35]. For the PWF price, the real interest rate (r) was determined based on two different terms using Equation (9), depending on the interest rate (φ) and the inflation rate (i):
If   φ > i   then   r = φ i 1 + i ;   If   i > φ   then   r = i φ 1 + φ
For a ten-year process, current values were calculated using Equation (10) with inflation and interest rates of 4% and 8%, respectively [36,37]:
P W F = ( r + 1 ) N ( r + 1 ) N . r
Equation (11) calculates the total CT ($) cost of an isolated building, where x and Ci are the used insulation material’s thickness (m) and unit price ($/m3), respectively:
C T = C A . P W F + C i . x
For optimal insulation thickness xopt (m), the total cost must be minimal and was determined as follows:
d C T d x = d d x . ( C A . P W F + C i .   x )
d C T d x = 0
x o p t = 86400   . ( N D D .   C f . P W F . k H U . C i . η ) 1 / 2 k . R T W
Equation (15) calculates the cost of the lifecycle ( p b ) following the profit from the investment made. In this equation, the SA represents annual savings:
p b = C T S A

2.4. Environmental Analysis

The steadily rising global population needs greater and greater quantities of energy every day. The related increasing energy demand has been generally utilized to heat houses. This demand is mainly supplied by fossil fuels, which are the widest and cheapest sources of energy, leading to the emission of greenhouse gases and harmful pollutants. Heating costs can be reduced by a certain degree with a certain increment in insulation thickness. The general chemical formula of combustion for fuel is expressed by Equation (16) [38]:
C x + H z + O w + S y + N t + α . A ( O 2 + 3.76 N 2 ) x C O 2 + ( z 2 ) H 2 O + y S O 2 + B . O 2 + E . N 2
Here, the constants (A, B, and E) are calculated as follows:
A = ( x + y + z 4 w 2 )
B = ( x + y + z 4 w 2 ) . ( α 1 )
E = 3.76 α   . ( x + y + z 4 w 4 ) + t 2
SOX and CO emissions are neglected in Equation (16). The emission rates of combustion products resulting from the burning of 1 kg of fuel can be calculated in Equations (20) and (21) [39]:
M C O 2 = x   C O 2 M k g     C O 2 / k g   f u e l  
M S O 2 = y   S O 2 M k g     S O 2 / k g   f u e l  
Here, M (kg/kmol) is the molar weight of the fuel determined as follows:
M = 12 x + z + 16 w + 32 y + 14 t
Depending on the total fuel consumption, emission values can be calculated using Equations (23) and (24):
m C O 2 = 44 . x M m f A
m S O 2 = 64 . y M m f A

3. Results and Discussion

3.1. Optimal Insulation Thickness

Optimal insulation thickness varied with fuel type, insulating material, and climate. Heat losses in buildings displayed a downtrend. Therefore, the heat losses and heat loads of buildings decreased. As a result, the overall amounts of fuel and emissions were reduced all at once. However, the overall cost (fuel and insulation) reduced initially and then climbed again after the minimal figure was reached. The ideal insulation thickness was determined by the point at which the overall cost was the lowest. Investment cost and total cost both increased due to unnecessarily increased insulation thickness. Figure 3 illustrates the relationship between the total cost of insulation thickness and externally insulated wall applications.
Optimal insulation thickness was computed for various fuels and insulating materials using Equation (14). The results can be seen in Table 4 and Table 5 for externally insulated walls in Jaisalmer, Kota, Mumbai, Chennai, Bhubaneswar, New Delhi, Lucknow, and Patna for cooling, as well as in Srinagar and Shillong for heating. Insulation became meaningless because of the initial investment cost and economic parameters in solutions for low numbers of degree days. It can be seen when Table 4 is reviewed that XPS had differences compared to EPS in using electricity for cooling when analyzing the optimal insulation thickness. The lowest thermal insulation thickness values related to using XPS material in building insulation were in New Delhi (Bsh) and Patna (Cwa), which were in the composite climatic zone, by 0.0382 (m). The highest value was in Jaiselmer by 0.0484 (m), which was in a hot and dry region.
New Delhi, Patna, and Jaisalmer had the highest values by 0.0908 (m) for cooling when EPS material was used in building insulation. These values proportionally varied by degree day number. The results for XPS material under a cooling load were 2.20 years on average (the payback period was 1.96 (years) at least and 2.36 (years) at best). The annual savings were 10.83 ($/m2) at least, 17.47 ($/m2) at best, and 13.23 ($/m2) on average. The annual savings rate was 40.28 (%) at least, 46.99 (%) at best, and 44.39 (%) on average. According to the results for the EPS material, the payback period was 1.49 (years) at least, 1.79 (years) at best, and 1.67 years on average. The annual savings were 12.79 ($/m2) at least, 19.94 ($/m2) at best, and 15.38 ($/m2) on average. The annual savings rate was 31.83 (%) at least, 37.38 (%) at best, and 35.22 (%) on average. According to the average results for both materials, 0.0607 (m) was the optimal insulation thickness, 1.93 (years) was the payback period, 14.30 was the annual savings ($/m2), and 39.80 (%) was the annual savings rate. Cities located in a hot climate zone only consumed energy for cooling. It can be seen that serious energy savings could be achieved with products selected in the optimum thickness values from insulation materials ranging from 0.1 cm to 0.12 cm in the current market.
Insulating properties of XPS and EPS materials in the cases of using natural gas, coal, and diesel fuel under a heating load can be seen in Table 5.
The lowest insulation thickness for XPS material was 0.0063 (m) for coal (Shillong); the highest value was 0.0834 (m) for diesel fuel (Srinagar). The lowest and highest values for EPS material for the same regions and types of fuels, respectively, were 0.0168 (m) and 0.1522. Regarding the results for XPS material under a heating load, the payback period was 1.24 (years) at least, 6.52 (years) at best, and 3.74 years on average. The annual savings were 0.29 ($/m2) at least, 51.73 ($/m2) at best, and 12.62 ($/m2) on average. The annual savings rate was 27.06 (%) at least, 90.66 (%) at best, and 61.82 (%) on average. According to the results for EPS material, the payback period was 0.95 (years) at least, 4.85 (years) at best, and 2.80 years on average. The annual savings were 0.68 ($/m2) at least, 55.92 ($/m2) at best, and 14.25 ($/m2) on average. The annual savings rate was 21.15 (%) at least, 78.19 (%) at best, and 51.13 (%) on average.
Regarding the average results in both regions for both materials, 0.0462 (m) was the optimal insulation thickness, the payback period was 3.27 (years), 13.43 was the annual savings ($/m2), and 56.47 (%) was the annual savings rate. The three main parameters that affected the results were fuel components, insulating properties, and the number of degree days. In general terms, variables were close to each other for India; good results for both heating and cooling were obtained.

3.2. Environmental Analysis

Figure 4 shows changes based on insulation thickness in the annual CO2 and SO2 gases. The CO2 emissions at different values of insulation thickness and for different types of fuels are presented in Figure 4a. For both EPS and XPS materials, the highest and lowest CO2 emissions were observed corresponding to coal and diesel, respectively. Fuel quantity for heating unit volume decreased with increasing insulation thickness. This is because there was an observed decrease in the emission of deleterious gases. According to the average results in two different cities, XPS and EPS respectively became 66% and 52% in the case of providing insulation and optimal insulation thickness. In addition, a decrease was seen in SO2 and CO2 emission values, with more sustainable ecological environmental structures.

3.3. Degree Days and Correlations

There have been some studies that have proposed methods to optimize the thicknesses of insulation materials. For example, Mahlia et al. [14] used the thermal conductivity (k) of different insulation materials to compute the optimal values. However, they did not consider the needed annual energy of buildings. On the contrary, Sisman et al. [15] computed optimal values using NDD but without using the k value of the insulation material. The results have shown that these two proposed methods are promising. However, a question arises here: what if both parameters, the k value of the insulation material and NDD, were used to compute optimal thickness? To investigate this problem, we tried to develop a mathematical model that used both parameters (k and NDD). The model was developed using a multiple linear regression method, which is a machine-learning technique [40,41]. Equation (25) is a general form of the model:
x o p t = a + b . k + c . N D D  
The optimal values of a, b, and c in Equation (25) were configured using the multiple curve-fitting tool in MATLAB for electricity, natural gas, coal, and diesel separately. For simplicity, we called these configurations Model (26), Model (27), Model (28), and Model (29), respectively.
Table 6 shows these configurations and their results. The results were validated by common statistical methods [42]. The R2 values of all these correlations statistically varied from 0.9943 to 0.9598; related values were close to (R ≤ 1). The root mean square error (RMSE) value was 0.0016 at best and 0.0140 at least because of the high harmony between the estimated and computed values. The sum of squares error (SSE) value was 2.209 × 10−5 at best and 1.96 × 10−4 at least. The ideal was zero in these two statistical methods (RMSE ≥ 0) -(SSE ≥ 0). Both the RMSE and SSE values were close to zero, and they gave consistent results. Figure 5 illustrates graphs of the optimal insulation thicknesses of the used insulating (XPS and EPS) materials, as computed by Equation (14) for different fuels.
The results obtained with the configurations of Equation (16) for electricity, natural gas, coal, and diesel (electricity: Model (26); natural gas: Model (27); coal: Model (28); diesel: Model (29)) were compared with previous works [14,15] and the reference values computed for different climate regions [17,19,33,43,44,45]. The comparison results are reported in Table 7. The heat map in Table 7 shows how close the computed thicknesses in each row were to the optimal thickness (reference value). These values in each row are the absolute values of the differences between the computed thicknesses and the reference value, i.e., errors of the model. Red shows the lowest error, green shows the highest error, and yellow shows the average error. Sisman’s model always had the highest error, while the others had similar errors in most cases. The mean errors of our model, Mahlia’s model [14], and Sisman’s model [15] for the instances in which we trained our model (instances 1–28 in Table 7) were 0.0713, 0.807, and 6.3, respectively. The mean errors of these models on the other cases in which we tested our model (cases 29–60 in Table 7) were 1.136, 2.023, and 10.868, respectively. When we considered all the instances in the table, the mean errors were 0.639, 1.456, and 8.737, respectively. The results show that Sisman’s model performs poorly in all the cases we tested, while others (Mahlia’s and our model) performed relatively well. When one compares our model with Mahlia’s model considering their mean errors, one can see that our model is superior to Mahlia’s model. Furthermore, this superiority was not only in the training instances but also in the testing instances.

4. Conclusions and Recommendations

The current study examined how different insulation materials and fuel types affected commercial and domestic building cooling and heating performances in different climatic zones of India. The following conclusions can be drawn from the current study:
  • For sites situated in hot and dry, warm and humid, and composite climatic regions, the ranges of XPS insulation thickness, annual savings, and payback period for these regions were found as 0.0382–0.0484 m, 10.83–17.47 $/m2, and 1.93–2.36 years, respectively. Similarly, these ranges for EPS insulation thickness, annual savings, and payback period were 0.0728–0.0908 m, 12.79–19.9347 $/m2, and 1.49–1.79 years, respectively.
  • For sites situated in cold climatic regions, the ranges of XPS insulation thickness, annual savings, and payback period were 0.0097–0.0834 m, 0.29–51.73 $/m2, and 1.24–6.52 years, respectively. The ranges for EPS insulation thickness, annual savings, and payback period were 0.0168–0.1522 m, 0.68–55.92 $/m2, and 0.95–4.85 years, respectively.
  • The ranges of GHG emissions for XPS material with natural gas, coal, and diesel fuels were 9.67–11.28 kg/m2-year, 27.12–32.68 kg/m2-year, and 2.89–4.51 kg/m2-year, respectively. Similarly, the ranges of GHG emissions for EPS material with natural gas, coal, and diesel fuels were 5.39–9.17 kg/m2-year, 9.47–23.75 kg/m2-year, and 2.26–3.43 kg/m2-year, respectively.
  • The optimal insulation thickness for XPS material was lower than that of EPS material, while the payback period and annual savings for XPS material were greater and lower, respectively, than those of EPS material in all the circumstances.
  • XPS was proved to be more effective than EPS, and the correlations obtained could aid in the determination of optimal insulation thickness for a specific location based on the number of degree days.
The study showed that using insulation on building exterior walls led to significant annual savings and had a payback period of less than five years, making it economically feasible. The study also found that increasing insulation thickness reduced GHG emissions from heating fuels. Future research can determine the best insulation thicknesses for different climate zones and materials.

Author Contributions

Conceptualization, M.A.K., A.Y. and A.S.; Data curation, M.A.K., A.Y. and A.S.; Investigation, M.A.K., A.Y., A.S. and R.K. Methodology, M.A.K., A.Y. and R.K., Software, M.A.K., A.Y., A.S. and R.K., Visualization, M.A.K., A.S. and T.S., Writing—original draft, M.A.K., A.Y. and A.S. Writing—review and editing, M.A.K., A.M., D.D., T.A. and T.S., Validation, A.Y. and A.S., Resources, A.M., D.D., T.A. and T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Data available on request.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Amtropical monsoon climate q A annual heat loss in unit area (J/m2-year)
Awtropical savanna climate r actual interest rate
Bskcold semi-arid (steppe) climate R i inside air film thermal resistances (m2K/W)
C A . C annual cooling energy cost ($/m2-year) R i z o thermal resistance of insulation layer (m2K/W)
C A . H annual heating energy cost ($/m2-year) R o outside air film thermal resistance (m2K/W)
C i cost of insulation in ($/m3) R T W sum of Ri.Rw.Ro (m2K/W)
CDD cooling degree days (°C-days) R w total thermal resistance of wall materials without insulation (m2K/W)
C f price of fuel ($/kg; $/m3) S A annual savings ($/m2)
Csahot summer mediterranean climateSO2sulfur dioxide
Csbwarm summer mediterranean climateUoverall heat transfer coefficient (W/m2K)
C T total cost ($) x thickness of insulation material (m)
CO2carbon dioxide x o p t optimum insulation thickness (m)
Dbfwarm summer humid continental climateXPSextruded polystyrene
E A annual energy requirement (J/m2-year)
EPSexpanded polystyreneGreek letters
g inflation rate η heating system efficiency
HDD heating degree days (°C-days)ΔTtemperature difference (°C)
H u heating value of fuel (J/kg; J/m3; J/kwh)
i interest rateSubscripts
k thermal conductivity of insulation material (W/m K)Aannual
LCAlifecycle cost analysisCcooling
Mmolar weight of fuelHheating
m f A amount of fuel consumed per year (kg/m2-year)iinside
Nlifetime (years)izoinsulation
NDD number of (degree days (°C-days)ooutside
p b payback period (years)optoptimum
PWFpresent worth factorttotal
q heat loss (MJ m2 year−1)wwall

References

  1. Kallioğlu, M.A.; Ercan, U.; Avcı, A.S.; Fidan, C.; Karakaya, H. Empirical Modeling between Degree Days and Optimum Insulation Thickness for External Wall. Energy Sources Part A Recovery Util. Environ. Eff. 2020, 42, 1314–1334. [Google Scholar] [CrossRef]
  2. Altun, A.F. Determination of Optimum Building Envelope Parameters of a Room Concerning Window-to-Wall Ratio, Orientation, Insulation Thickness and Window Type. Buildings 2022, 12, 383. [Google Scholar] [CrossRef]
  3. Wang, H.; Huang, Y.; Yang, L. Integrated Economic and Environmental Assessment-Based Optimization Design Method of Building Roof Thermal Insulation. Buildings 2022, 12, 916. [Google Scholar] [CrossRef]
  4. Shaik, S.; Gorantla, K.; Ghosh, A.; Arumugam, C.; Maduru, V.R. Energy Savings and Carbon Emission Mitigation Prospective of Building’s Glazing Variety, Window-to-Wall Ratio and Wall Thickness. Energies 2021, 14, 8020. [Google Scholar] [CrossRef]
  5. Skaar, C.; Gaarder, J.-E.; Bunkholt, N.S.; Sletnes, M. Energy Upgrading of Basement Exterior Walls: The Good, the Bad and the Ugly. Buildings 2023, 13, 133. [Google Scholar] [CrossRef]
  6. Kallioğlu, M.A.; Sharma, A.; Chinnasamy, V.; Chauhan, R.; Singh, T. Optimum Insulation Thickness Assessment of Different Insulation Materials for Mid-Latitude Steppe and Desert Climate (BSH) Region of India. Mater. Today Proc. 2020, 44, 4421–4424. [Google Scholar] [CrossRef]
  7. Xu, H.; Ding, J.; Li, T.; Mu, C.; Gu, X.; Wang, R. A Study on Optimum Insulation Thickness in Walls of Chinese Solar Greenhouse for Energy Saving. Agronomy 2022, 12, 1104. [Google Scholar]
  8. Nyers, J.; Kajtar, L.; Tomić, S.; Nyers, A. Investment-Savings Method for Energy-Economic Optimization of External Wall Thermal Insulation Thickness. Energy Build. 2015, 86, 268–274. [Google Scholar] [CrossRef]
  9. Liu, X.; Chen, Y.; Ge, H.; Fazio, P.; Chen, G. Determination of Optimum Insulation Thickness of Exterior Wall with Moisture Transfer in Hot Summer and Cold Winter Zone of China. Procedia Eng. 2015, 121, 1008–1015. [Google Scholar] [CrossRef] [Green Version]
  10. Yuan, J.; Farnham, C.; Emura, K.; Alam, M.A. Proposal for Optimum Combination of Reflectivity and Insulation Thickness of Building Exterior Walls for Annual Thermal Load in Japan. Build. Environ. 2016, 103, 228–237. [Google Scholar] [CrossRef]
  11. Dombaycı, Ö.A. The Environmental Impact of Optimum Insulation Thickness for External Walls of Buildings. Build. Environ. 2007, 42, 3855–3859. [Google Scholar] [CrossRef]
  12. Kaynakli, O.; Bademlioglu, A.H.; Ufat, H.T. Determination of Optimum Insulation Thickness for Different Insulation Applications Considering Condensation. Teh. Vjesn. 2018, 25, 32–42. [Google Scholar]
  13. Bolattürk, A. Determination of Optimum Insulation Thickness for Building Walls with Respect to Various Fuels and Climate Zones in Turkey. Appl. Therm. Eng. 2006, 26, 1301–1309. [Google Scholar] [CrossRef]
  14. Mahlia, T.M.I.; Taufiq, B.N.; Masjuki, H.H. Correlation between Thermal Conductivity and the Thickness of Selected Insulation Materials for Building Wall. Energy Build. 2007, 39, 182–187. [Google Scholar] [CrossRef]
  15. Sisman, N.; Kahya, E.; Aras, N.; Aras, H. Determination of Optimum Insulation Thicknesses of the External Walls and Roof (Ceiling) for Turkey’s Different Degree-Day Regions. Energy Policy 2007, 35, 5151–5155. [Google Scholar] [CrossRef]
  16. Shanmuga Sundaram, A.; Bhaskaran, A. Optimum Insulation Thickness of Walls for Energy-Saving in Hot Regions of India. Int. J. Sustain. Energy 2014, 33, 213–226. [Google Scholar] [CrossRef]
  17. Mishra, S.; Usmani, J.A.; Varshney, S. Energy Saving Analysis in Building Walls through Thermal Insulation System. Int. J. Eng. Res. Appl. 2012, 2, 128–135. [Google Scholar]
  18. Raza, A.; Aggarwal, V. Evaluation of Optimum Insulation Thickness for Cost Saving of Wall Design in India. Available online: https://www.researchsquare.com/article/rs-1565499/v1.pdf (accessed on 13 October 2022).
  19. Singh, H.K.; Prakash, R.; Shukla, K.K. Economic Insulation Thickness of External Walls in Hot and Composite Regions of India. In Proceedings of the Global Conference on Renewable Energy GCRE2016, Patna, India, 4–6 March 2016; p. 6. [Google Scholar]
  20. Clark, H.; Wu, H. The Sustainable Development Goals: 17 Goals to Transform Our World. In Furthering Work United Nations; UN: New York, NY, USA, 2016; pp. 36–54. [Google Scholar]
  21. Kottek, M.; Grieser, J.; Beck, C.; Rudolf, B.; Rubel, F. World Map of the Köppen-Geiger Climate Classification Updated. Meteorol. Z. 2006, 15, 259–263. [Google Scholar] [CrossRef]
  22. Rubel, F.; Brugger, K.; Haslinger, K.; Auer, I. The climate of the European Alps: Shift of very high resolution Köppen-Geiger climate zones 1800–2100. Meteorol. Z. 2017, 26, 115–125. [Google Scholar] [CrossRef]
  23. JRC Photovoltaic Geographical Information System (PVGIS)—European Commission. Available online: https://re.jrc.ec.europa.eu/pvg_tools/en/#MR (accessed on 15 February 2022).
  24. Szokolay, S. Introduction to Architectural Science; Routledge: London, UK, 2012. [Google Scholar]
  25. Shehadi, M. Energy Consumption Optimization Measures for Buildings in the Midwest Regions of USA. Buildings 2018, 8, 170. [Google Scholar] [CrossRef] [Green Version]
  26. Kon, O.; Caner, İ. The Effect of External Wall Insulation on Mold and Moisture on the Buildings. Buildings 2022, 12, 521. [Google Scholar] [CrossRef]
  27. Aditya, L.; Mahlia, T.M.I.; Rismanchi, B.; Ng, H.M.; Hasan, M.H.; Metselaar, H.S.C.; Muraza, O.; Aditiya, H.B. A Review on Insulation Materials for Energy Conservation in Buildings. Renew. Sustain. Energy Rev. 2017, 73, 1352–1365. [Google Scholar] [CrossRef]
  28. Standard, T. 825 (TS 825). In Thermal Insulation in Buildings; Official Gazette (23725): Ankara, Turkey, 1999; Volume 14. [Google Scholar]
  29. Nellis, G.; Klein, S. Heat Transfer; Cambridge University Press: Cambridge, UK, 2008; ISBN 1-139-47661-0. [Google Scholar]
  30. Qian, X.; Lee, S.W.; Yang, Y. Heat Transfer Coefficient Estimation and Performance Evaluation of Shell and Tube Heat Exchanger Using Flue Gas. Processes 2021, 9, 939. [Google Scholar] [CrossRef]
  31. Statistics, E.; Central Statistics Office, Ministry of Statistics and Programme Implementation. Govt. India New Delhi (22nd Issue) 2015. Available online: www.mospi.gov (accessed on 8 February 2022).
  32. Cengel, Y. Heat and Mass Transfer: Fundamentals and Applications; McGraw-Hill Higher Education: Columbus, OH, USA, 2014. [Google Scholar]
  33. Kurekci, N.A. Determination of Optimum Insulation Thickness for Building Walls by Using Heating and Cooling Degree-Day Values of All Turkey’s Provincial Centers. Energy Build. 2016, 118, 197–213. [Google Scholar] [CrossRef]
  34. Llantoy, N.; Chàfer, M.; Cabeza, L.F. A Comparative Life Cycle Assessment (LCA) of Different Insulation Materials for Buildings in the Continental Mediterranean Climate. Energy Build. 2020, 225, 110323. [Google Scholar] [CrossRef]
  35. Hasan, A. Optimizing Insulation Thickness for Buildings Using Life Cycle Cost. Appl. Energy 1999, 63, 115–124. [Google Scholar] [CrossRef]
  36. Inflation Rate in India 1984–2024. Available online: https://www.statista.com/statistics/271322/inflation-rate-in-india/ (accessed on 16 August 2020).
  37. SBI—Loans, Accounts, Cards, Investment, Deposits, Net Banking—Personal Banking. Available online: https://sbi.co.in/ (accessed on 23 February 2021).
  38. Yildiz, A.; Gurlek, G.; Erkek, M.; Ozbalta, N. Economical and Environmental Analyses of Thermal Insulation Thickness in Buildings. J. Therm. Sci. Technol. 2008, 28, 25–34. [Google Scholar]
  39. Ozel, M. Cost Analysis for Optimum Thicknesses and Environmental Impacts of Different Insulation Materials. Energy Build. 2012, 49, 552–559. [Google Scholar] [CrossRef]
  40. Asadi, S.; Amiri, S.S.; Mottahedi, M. On the Development of Multi-Linear Regression Analysis to Assess Energy Consumption in the Early Stages of Building Design. Energy Build. 2014, 85, 246–255. [Google Scholar] [CrossRef]
  41. Chapra, S.C.; Canale, R.P. Numerical Methods for Engineers; McGraw-Hill Higher Education: Boston, MA, USA, 2010. [Google Scholar]
  42. Qian, X.; Lee, S.; Soto, A.; Chen, G. Regression Model to Predict the Higher Heating Value of Poultry Waste from Proximate Analysis. Resources 2018, 7, 39. [Google Scholar] [CrossRef] [Green Version]
  43. Alghoul, S.K.; Gwesha, A.O.; Naas, A.M. The Effect of Electricity Price on Saving Energy Transmitted from External Building Walls. Energy Res. J. 2016, 7, 1–9. [Google Scholar] [CrossRef] [Green Version]
  44. Alsayed, M.F.; Tayeh, R.A. Life Cycle Cost Analysis for Determining Optimal Insulation Thickness in Palestinian Buildings. J. Build. Eng. 2019, 22, 101–112. [Google Scholar] [CrossRef]
  45. Ucar, A.; Balo, F. Effect of Fuel Type on the Optimum Thickness of Selected Insulation Materials for the Four Different Climatic Regions of Turkey. Appl. Energy 2009, 86, 730–736. [Google Scholar] [CrossRef]
Figure 1. Selected Indian cities’ geographic locations.
Figure 1. Selected Indian cities’ geographic locations.
Buildings 13 00569 g001
Figure 2. The exterior wall model in the study and the characteristics of its components.
Figure 2. The exterior wall model in the study and the characteristics of its components.
Buildings 13 00569 g002
Figure 3. Cost–insulation thickness relationship for EPS and XPS insulation materials: (a) Mumbai, (b) Srinigar, (c) Chennai, and (d) Patna.
Figure 3. Cost–insulation thickness relationship for EPS and XPS insulation materials: (a) Mumbai, (b) Srinigar, (c) Chennai, and (d) Patna.
Buildings 13 00569 g003
Figure 4. Fuel gas emission–insulation thickness: (a) CO2 (Shillong) (b) SO2-Coal (Shillong and Srinagar).
Figure 4. Fuel gas emission–insulation thickness: (a) CO2 (Shillong) (b) SO2-Coal (Shillong and Srinagar).
Buildings 13 00569 g004
Figure 5. Optimal insulation thickness and k-NDD values: (a) electricity; (b) natural gas; (c) coal; (d) diesel.
Figure 5. Optimal insulation thickness and k-NDD values: (a) electricity; (b) natural gas; (c) coal; (d) diesel.
Buildings 13 00569 g005
Table 1. Cooling and heating values of the selected cities [23].
Table 1. Cooling and heating values of the selected cities [23].
No.CityLatitude
(Degree)
Altitude
(m)
HDD (°C/Year)CDD (°C/Year)Average Temp. (°C/Year)ProcessClimate
1Jaisalmer29.90 N22543256527.38CoolingHot dry
2Kota25.20 N27187212326CoolingHot dry
3Mumbai19.07 N140209226.82CoolingWarm humid
4Chennai13.08 N6.70248527.82CoolingWarm humid
5Bhubaneswar20.29 N581189526.24CoolingWarm humid
6New Delhi28.61 N216267179124.34CoolingComposite
7Lucknow26.85 N123144184224.89CoolingComposite
8Patna25.59 N5375179125.02CoolingComposite
9Srinagar34.08 N158521092613.14HeatingCold
10Shillong25.57 N1525815016.04HeatingCold
Table 2. Construction of walls used in the analysis [27,28].
Table 2. Construction of walls used in the analysis [27,28].
Wall TypeDensity
(kg·m−3)
Thickness (m)Thermal Conductivity
(W/m·K)
Thermal Resistance
(m2K/W)
Thermal Resistance
RTW (Total Wall) (m2K/W)
Inner plaster12000.020.850.02300.548
Hollow brick7500.130.450.2880
Outer plaster11000.020.850.0230
R inside---0.1670
R outside---0.0450
Table 4. Results of various insulation materials for cooling.
Table 4. Results of various insulation materials for cooling.
MaterialsXopt (m)Annual Savings ($/m2)Payback Period (Years)Annual Savings Rate (%)
Jaisalmer
XPS0.048417.471.9640.28
EPS0.090819.941.4931.83
Kota
XPS0.042813.632.1643.71
EPS0.080915.821.6434.66
Mumbai
XPS0.042413.362.1843.99
EPS0.080215.541.6534.89
Chennai
XPS0.047516.771.9940.84
EPS0.089119.191.5132.29
Bhubaneswar
XPS0.040111.702.2945.88
EPS0.075413.731.7436.46
New Delhi
XPS0.038210.832.3646.99
EPS0.072812.791.7937.38
Lucknow
XPS0.038911.252.3346.44
EPS0.074113.251.7736.92
Patna
XPS0.038210.832.3646.99
EPS0.072812.791.7937.38
Table 5. Results of various insulation materials for heating.
Table 5. Results of various insulation materials for heating.
FuelMaterialsXopt (m)Annual Savings ($/m2)Payback Period (Years)Annual Savings Rate (%)
Srinagar
Natural gasXPS0.02424.363.2760.39
EPS0.04835.632.4748.80
CoalXPS0.01882.623.8667.70
EPS0.03873.622.9155.29
DieselXPS0.083451.731.2427.06
EPS0.152255.920.9521.15
Shillong
Natural gasXPS0.00970.695.4983.69
EPS0.02271.254.1170.57
CoalXPS0.00630.296.5290.66
EPS0.01680.684.8578.19
DieselXPS0.046516.052.0341.44
EPS0.087318.421.5432.79
Table 6. Developed model regression constants.
Table 6. Developed model regression constants.
FuelModelabcR2SSERMSE
Electricity26−0.10394.1611.823 × 10−50.99433.443 × 10−50.0016
Natural gas27−0.059292.0611.549 × 10−50.96043.08 × 10−50.0055
Coal28−0.050791.6891.329 × 10−50.95982.209 × 10−50.0047
Diesel29−0.15096.0893.934 × 10−50.96611.96 × 10−40.0140
Table 7. Relative error statistical results of reference results with estimated models.
Table 7. Relative error statistical results of reference results with estimated models.
LowMidHigh
CaseLocationReferenceRef. ValueRelative Error (Er) %
Developed ModelPrevious Models
CountryCityReferenceClimateNDDProcessFuelMaterialk
(W/m·K)
xopt (m)Developed Eq. for Corresponding FuelMahlia et al. [14]Sisman et al. [15]
1IndiaJaisalmerPresent StudyBwh2565CoolingElectricityXPS0.02600.04840.05360.00476.6422
2IndiaJaisalmerPresent StudyBwh2565CoolingElectricityEPS0.0350.09080.02610.37363.0736
3IndiaKotaPresent StudyBsh2123CoolingElectricityXPS0.0260.04280.00330.12556.4946
4IndiaKotaPresent StudyBsh2123CoolingElectricityEPS0.0350.08090.00640.29692.965
5IndiaMumbaiPresent StudyAw2092CoolingElectricityXPS0.0260.04240.00060.13616.4819
6IndiaMumbaiPresent StudyAw2092CoolingElectricityEPS0.0350.08020.00480.29082.9555
7IndiaChennaiPresent StudyAw2485CoolingElectricityXPS0.0260.04750.04290.01416.6034
8IndiaChennaiPresent StudyAw2485CoolingElectricityEPS0.0350.08910.02380.36163.0534
9IndiaBhubaneswarPresent StudyAw1895CoolingElectricityXPS0.0260.04010.03270.20136.3431
10IndiaBhubaneswarPresent StudyAw1895CoolingElectricityEPS0.0350.07540.0110.24562.9053
11IndiaNew DelhiPresent StudyAw1791CoolingElectricityXPS0.0260.03820.03420.2616.3874
12IndiaNew DelhiPresent StudyAw1791CoolingElectricityEPS0.0350.07280.0210.21872.8764
13IndiaLucknowPresent StudyCwa1842CoolingElectricityXPS0.0260.03890.02770.23836.4096
14IndiaLucknowPresent StudyCwa1842CoolingElectricityEPS0.0350.07410.01570.23242.8898
15IndiaPatnaPresent StudyCwa1791CoolingElectricityXPS0.0260.03820.03420.2616.3874
16IndiaPatnaPresent StudyCwa1791CoolingElectricityEPS0.0350.07280.0210.21872.8764
17IndiaSrinagarPresent StudyBwk2109HeatingNatural gasXPS0.0260.02420.11420.990612.189
18IndiaSrinagarPresent StudyBwk2109HeatingNatural gasEPS0.0350.04830.05770.17765.6081
19IndiaSrinagarPresent StudyBwk2109HeatingCoalXPS0.0260.01880.12511.562315.9773
20IndiaSrinagarPresent StudyBwk2109HeatingCoalEPS0.0350.03870.06060.46987.2474
21IndiaSrinagarPresent StudyBwk2109HeatingDieselXPS0.0260.08340.08370.42242.827
22IndiaSrinagarPresent StudyBwk2109HeatingDieselEPS0.0350.15220.04610.62631.0971
23IndiaShillongPresent StudyAm815HeatingNatural gasXPS0.0260.00970.28663.966115.0769
24IndiaShillongPresent StudyAm815HeatingNatural gasEPS0.0350.02270.1221.50575.8699
25IndiaShillongPresent StudyAm815HeatingCoalXPS0.0260.00630.37226.646323.7534
26IndiaShillongPresent StudyAm815HeatingCoalEPS0.0350.01680.14032.38578.2825
27IndiaShillongPresent StudyAm815HeatingDieselXPS0.0260.04650.15110.03592.3537
28IndiaShillongPresent StudyAm815HeatingDieselEPS0.0350.08730.07990.34850.7863
29LibyaTripoli[43]Csa492CoolingElectricityEPS0.0370.0690.14520.1270.5453
30TurkeyAdana[33]Csa874HeatingNatural gasXPS0.0240.0690.94620.30931.3823
31TurkeyAdana[33]Csa874HeatingNatural gasEPS0.0350.0690.61760.17571.3823
32TurkeyAdana[33]Csa874HeatingNatural gasGlass wool0.050.0241.38742.94385.849
33TurkeyAdana[33]Csa874HeatingNatural gasRock wool0.0480.0350.51931.51243.6965
34TurkeyAdana[33]Csa874HeatingNatural gasPolyurethane0.0170.051.21430.00142.2875
35TurkeyAdana[33]Csa874HeatingCoalXPS0.0310.0480.72530.07742.4245
36TurkeyAdana[33]Csa874HeatingCoalEPS0.0440.0171.06713.47378.6692
37TurkeyAdana[33]Csa874HeatingCoalGlass wool0.0610.0311.05983.54274.3025
38TurkeyAdana[33]Csa874HeatingCoalRock wool0.0590.0440.37451.98332.7358
39TurkeyAdana[33]Csa874HeatingCoalPolyurethane0.0220.0611.03310.21871.6947
40PalestineJericho[44]Csa1989CoolingElectricityPolystyrene0.0380.0590.53250.05274.1762
41PalestineHebron[44]Dfb456CoolingElectricityPolystyrene0.0380.0221.84041.82313.5769
42PalestineJerusalem[44]Csa768CoolingElectricityPolystyrene0.0380.0490.39130.26752.0433
43PalestineTulkarem[44]Csa1066CoolingElectricityPolystyrene0.0380.0570.29130.08962.3492
44PalestineGaza[44]Bsh1097CoolingElectricityPolystyrene0.0380.0620.19620.00172.1463
45PalestineBethelem[44]Csa971CoolingElectricityPolystyrene0.0380.0530.35610.17192.3573
46PalestineJenin[44]Csa1399CoolingElectricityPolystyrene0.0380.0680.17160.08662.4454
47PalestineNablus[44]Csa854CoolingElectricityPolystyrene0.0380.0520.34120.19442.1064
48TurkeyAğrı[45]Dsb4423HeatingCoalXPS0.0310.02611.31230.981520.3638
49TurkeyAğrı[45]Dsb4423HeatingNatural gasXPS0.0310.03141.32840.647116.7578
50TurkeyAydın[45]Csa1213HeatingCoalXPS0.0310.00227.040822.50894.6422
51TurkeyAydın[45]Csa1213HeatingNatural gasXPS0.0310.0053.67819.343541.0826
52TurkeyElazığ[45]Dsa2653HeatingCoalXPS0.0310.0190.93831.72218.9686
53TurkeyElazığ[45]Dsa2653HeatingNatural gasXPS0.0310.01821.51081.841619.8463
54TurkeyKocaeli[45]Cfa1786HeatingCoalXPS0.0310.01131.23943.576823.9209
55TurkeyKocaeli[45]Cfa1786HeatingNatural gasXPS0.0310.01062.0443.87925.5666
56India-[19]Hot and humid1288CoolingElectricityXPS0.0360.0281.4761.08916.8622
57India-[19]Hot and dry1111CoolingElectricityXPS0.0360.0311.13230.88695.3529
58India -[19]Composite1121CoolingElectricityXPS0.0360.0311.13820.88695.396
59IndiaDehradun[17]Cfa3587HeatingNatural gasXPS0.0330.0570.12780.05197.3542
60IndiaDehradun[17]Cfa3587HeatingNatural gasEPS0.0310.0730.17580.29155.5231
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kallioğlu, M.A.; Yılmaz, A.; Sharma, A.; Mohamed, A.; Dobrotă, D.; Alam, T.; Khargotra, R.; Singh, T. Optimal Insulation Assessment, Emission Analysis, and Correlation Formulation for Indian Region. Buildings 2023, 13, 569. https://doi.org/10.3390/buildings13020569

AMA Style

Kallioğlu MA, Yılmaz A, Sharma A, Mohamed A, Dobrotă D, Alam T, Khargotra R, Singh T. Optimal Insulation Assessment, Emission Analysis, and Correlation Formulation for Indian Region. Buildings. 2023; 13(2):569. https://doi.org/10.3390/buildings13020569

Chicago/Turabian Style

Kallioğlu, Mehmet Ali, Ahmet Yılmaz, Ashutosh Sharma, Ahmed Mohamed, Dan Dobrotă, Tabish Alam, Rohit Khargotra, and Tej Singh. 2023. "Optimal Insulation Assessment, Emission Analysis, and Correlation Formulation for Indian Region" Buildings 13, no. 2: 569. https://doi.org/10.3390/buildings13020569

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