Next Article in Journal
A Multi-Step Ensemble Approach for Energy Community Day-Ahead Net Load Point and Probabilistic Forecasting
Next Article in Special Issue
BIM Integration with XAI Using LIME and MOO for Automated Green Building Energy Performance Analysis
Previous Article in Journal
Factoring Building Refurbishment and Climatic Effect into Heat Demand Assessments and Forecasts: Case Study and Open Datasets for Germany
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sensitivity Assessment of Building Energy Performance Simulations Using MARS Meta-Modeling in Combination with Sobol’ Method

Institute of Energy Efficiency and Sustainable Building (E3D), RWTH Aachen University, 52074 Aachen, Germany
*
Authors to whom correspondence should be addressed.
Energies 2024, 17(3), 695; https://doi.org/10.3390/en17030695
Submission received: 21 October 2023 / Revised: 22 January 2024 / Accepted: 23 January 2024 / Published: 31 January 2024
(This article belongs to the Special Issue Energy Consumption and Environmental Quality in Buildings)

Abstract

:
Large discrepancies can occur between building energy performance simulation (BEPS) outputs and reference data. Uncertainty and sensitivity analyses are performed to discover the significant contributions of each input parameter to these discrepancies. Variance-based sensitivity analyses typically require many stochastic simulations, which is computationally demanding (especially in the case of the large number of input parameters involved in the analysis). To overcome these impediments, this study proposes a reliable meta-model-based sensitivity analysis, including validation, Morris’ method, multivariate adaptive regression splines (MARS) meta-modeling, and Sobol’ method, to identify the most influential input parameters on BEPS prediction (annual energy consumption) at the early building design process. A hypothetical building is used to analyze the proposed methodology. Six statistical metrics are applied to verify and quantify the accuracy of the model. It is concluded that the cooling set-point temperature and g-value of the window are the most influential input parameters for the analyzed case study.

1. Introduction

Over the past decades, building energy performance simulation (BEPS) has been particularly popular as a major approach to estimating the energy consumption of buildings and, subsequently, reducing the accompanying greenhouse gas emissions. However, the energy performance of buildings depends on a vast number of factors, including the building’s characteristics, HVAC systems, control and maintenance, weather conditions, sociological factors, and occupants’ behavior [1,2]. Due to the uncertainty in these parameters of a model, huge and sometimes serious discrepancies can arise between BEPS predictions provided by different dynamic BEPS tools and reference data. This mismatch is commonly referred to as the “Energy Performance Gap,” which can be classified into three different groups: the gap between first-principle energy models and measurements, the gap between machine learning approaches and measurements, and the gap between predictions and display certificates in legislation [3,4].
Due to the proliferation of numerous BEPS tools, an acceptable degree of confidence in BEPS prediction has become a salient preoccupation for the building simulation community. Validation procedures help to ensure the accuracy and plausibility of the simulation results. The purpose of the validation is to evaluate the quality and reliability of the predicted and calculated performance simulations and to identify errors in the simulation outputs. Subsequently, by using calibration approaches, the uncertainties can be reduced or eliminated, and eventually, simulation results can be aligned with reference data to the greatest extent. Three different kinds of techniques are applied to the validation of dynamic thermal models of buildings: empirical validation, analytical verification, and software-to-software comparative testing [5,6,7].
To quantify the accuracy of building models, i.e., to characterize the discrepancies between simulation results and real data, statistical performance metrics are used. There are several types of statistical metrics that have been commonly used in the literature [8,9,10,11,12,13,14,15,16,17,18,19,20,21,22], including the Mean Absolute Error (MAE), Mean Bias Error (MBE), Normalized Mean Bias Error (NMBE), Root Mean Squared Error (RMSE), Coefficient of Variance of Root Mean Squared Error (CV-RMSE), and Coefficient of Determination (R2). A large number of previous studies have used the mentioned validation techniques and statistical criteria to evaluate the level of prediction accuracy of energy performance models [23,24,25,26,27,28,29,30,31,32,33].
To discover the most influential building model parameters on the simulation outputs (i.e., parameters that contribute enormously to the discrepancies between the energy performance simulation output and the reference data), uncertainty and sensitivity analyses are conducted [34,35,36,37,38]. Table 1 summarizes the validation methods, sensitivity analysis and statistical performance indices used in the reviewed literature in the context of BEPS.
Sensitivity analysis (SA) methods may be categorized into two major groups: local sensitivity analysis (LSA) and global sensitivity analysis (GSA) [49,50]. LSA is based on one-factor-at-a-time (OAT) approaches, where the inference about the model output variation is made by varying only one model input parameter at a time, and all other parameters are kept constant at their nominal values (or baseline points). Although LSA is computationally efficient [51] and easy to comprehend, it is predicated on the assumption that the model is linear (or at least additive) [52]. Furthermore, it cannot grasp mutual interactions between input parameters and is not feasible to explore multidimensional parameter spaces [49]. Therefore, the robustness and reliability of LSA cannot be guaranteed.
The GSA, on the other hand, investigates the model output where all the input parameters chosen for analysis are varied over the whole parameter space simultaneously. Therefore, GSA imparts further information regarding the correlation between input parameters and may be more reliable. Various methods have been proposed for GSA. The most commonly used approaches are a regression method [53], a screening-based method (such as Morris’ method) [34,54,55], and a variance-based method (such as the Sobol’ method and Fourier Amplitude Sensitivity Test (FAST)) [35,48,55,56]. Due to the non-linear, non-monotonic behavior of sophisticated building models, GSA has been widely implemented to rank the most important parameters [55,57]. However, the greatest shortcoming of GSA lies in its computationally demanding evaluations, especially in the early stages of the building design process, in which many design parameters are explored and a quick evaluation of the building performance is highly desirable. To implement GSA methods, a large number of stochastic simulations are typically required. For computationally expensive models, comprehensive GSA can be time-intensive and hence increase their cost [51,58].
To tackle the computational burdens of GSA for intricate models such as those used in dynamic BEPS (i.e., physics-based models), this study proposes a metamodeling-based SA method. Meta-models (also referred to as surrogate models or emulators) are approximate, fast models that can be substituted for highly sophisticated, time-consuming models (such as detailed engineering-based energy models) with acceptable reliability [59,60,61]. Thus, for a large number of simulation runs, meta-models are less computationally demanding in comparison with original, complex building models. A variety of statistical meta-modeling approaches have been employed in the context of BEPS, including regression analysis [62,63,64], support vector regression (SVR) [65,66,67], multivariate adaptive regression splines (MARS) [59,68,69], gaussian process regression (GPR) [60,61], Random Forest (RF) [69], neural networks [59,62], and ensemble models [63].
Rodríquez et al. [70] applied a regression method to calculate sensitivity indices using a case study of a dwelling with three ranges of weather data and three levels of occupancy. A sample size of 200 was investigated to create the set of parameters. Bucking et al. [71] proposed a methodology to identify the influential variables (design and operation parameters) on the building’s performance (annual electricity consumed). A back-tracking search explored and ranked 8 of 26 variables that have a significant impact on the energy consumption in a net-zero energy house using 6000 EnergyPlus simulation runs. Kim [72] developed two meta-models (gaussian process emulator and polynomial chaos expansion) in MATLAB to alleviate the computational burden of BEPS tools, using a five-story office building located in South Korea. The uncertainty results were propagated by 200 simulation runs using the EnergyPlus model, which has 47 input parameters. Hester et al. [64] presented a probabilistic framework using a regression-based energy meta-model to estimate the energy consumption of single-family buildings at the early stages of the design process. A total of 17 building attributes were selected, and Monte Carlo simulations were carried out to evaluate the variability in the meta-model output (energy consumption).
An extensive review of validation, uncertainty, and sensitivity analysis studies in the context of BEPS reveals several substantial research gaps. The choice of uncertain parameters for performing the SA is often based on either expert knowledge or existing previous studies, which can seriously impede an accurate and reliable SA. Furthermore, running stochastic simulations with a large number of samples in BEPS tools is very time-consuming or even impractical in the case of complex models. Finally, the application of the MARS meta-model in performing variance-based SA (in particular the Sobol’ method) has rarely been carried out in the literature, and further studies are required. This study proposes a novel, four-stage SA framework, including validation, Morris’ method, meta-modeling, and Sobol’ method. The proposed framework improves the robustness and accuracy of the SA in the early building design process, considering a large set of uncertainties (105 uncertain parameters), to support robust decision-making under input uncertainties. The MARS machine learning algorithm is applied to develop a meta-model for the complex building model, and its prediction accuracy is scrutinized by comparison with three additional meta-models, namely Polynomial Regression (PR), Random Forest (RF), and Support Vector Regression (SVR) based on the radial basis function (RBF) kernel, using several statistical metrics.
The proposed methodology is applied to answer the main research question of this study: Which uncertain input parameters (in the analyzed case) drive the majority of the model output variation? The remainder of this paper is structured as follows: Section 2 introduces the proposed methodology. In Section 3, the investigated case study, the simulation tool used, and the weather data are presented. The results and discussion are described in Section 4. Finally, Section 5 gives concluding remarks.

2. Methodology

The methodology in this study is comprised of two constituent parts. The first part involves the validation approach, which is applied to evaluate and quantify the accuracy of dynamic BEPS results.
Two-time series datasets are created: reference and synthetic datasets. The reference dataset is built from the simulation results of the investigated case study (see Section 3) for one year, containing values of heating and cooling loads at an hourly resolution. To conduct the validation and evaluate the discrepancies, a virtual building is modeled. The virtual building uses the same simulation setup; however, the values of the input parameters, i.e., the thermophysical and optical properties of the building envelope, infiltration rate, internal heat gains, and HVAC system, are randomly varied using Latin Hypercube sampling (LHS). For each set of input parameters, the probability density function is given as a uniform distribution. The simulation results of the virtual building (heating and cooling loads) are collected in a synthetic dataset. The major advantage of the synthetic dataset over the measurement dataset is that there are no measurement errors, and all required information regarding the parameters of the building is attainable.
The main aim of creating the synthetic dataset is to assess the deviation of the simulation results and, subsequently, to inspect the most influential input parameters on the simulation outputs by SA, which will be discussed in the second part of the methodology.
Both qualitative and quantitative validations are performed. In quantitative validation, statistical metrics are applied to ascertain and quantify the mismatch between the simulation results of reference and synthetic datasets (see Section 2.1). The validation approach used in the first part of the methodology is schematically illustrated in Figure 1.
The second part of the methodology investigated in this study presents the uncertainty and sensitivity analysis, consisting mainly of three sequential stages: screening-based SA (Morris’ method), MARS meta-modeling, and variance-based SA (Sobol’ method).
First, the screening-based SA by Morris [73] is performed to determine important input parameters and reduce the number of parameters that are later used in the metamodel-based SA. Since this study investigates a large number of parameters, implementing Morris’ method prior to performing variance-based SA (Sobol’ method) can substantially reduce the computational cost by selecting key parameters influencing BEPS results. Note that in this study, the term “computational cost” is interpreted as the number of times the investigated model has to be evaluated. 105 uncertain input parameters, including thermophysical properties, infiltration rate, internal heat gains, and HVAC system, are selected. Table A1 (Appendix A) shows the list of input parameters included in the analysis and the assigned uncertainty ranges for each parameter. Due to the large number of input parameters involved in this study, an uncertainty of ±10% is considered for all parameters. It is worth mentioning that the proposed SA framework can also consider parameters with different uncertainty ranges. To build samples of input parameter sets, the factorial sampling plan proposed by Morris [73] is used. All input parameters are uniformly distributed, and it is assumed that there is no correlation between input parameters. Two statistical measures from Morris’ method are used to analyze the influence of each parameter on model outputs (annual heating, cooling, and total energy consumption): the mean of the absolute value μ* and the standard deviation σ of the elementary effects (see Section 2.2). A total of 44 uncertain parameters are selected from an original set of 105.
The main objective of the SA in this study is applying the Sobol’ method [74], which is one of the most robust and accurate variance-based SAs [55,75]. However, it requires a large number of samples and simulations, which incur considerable computational costs. Before conducting the Sobol’ method, meta-models are implemented in the second stage of the SA, which can be evaluated faster and reduce the computational cost of Sobol’ indices in comparison with the direct simulation of the original BEPS model. To generate a dataset for developing meta-models, Latin Hypercube sampling (LHS) in combination with Monte Carlo simulation is applied in this study.
LHS is a particular form of stratified sampling method and is commonly used for uncertainty and sensitivity analysis in computationally demanding models. Due to its conceptual simplicity and efficient stratification over the range of each uncertain parameter, LHS is very popular [76]. The samples (N = 400) for each of the selected uncertain parameters (k = 44) are created using their probability distributions and stored in a matrix X as:
X = x 1,1 x 2,1 x k , 1 x 1,2 x 2,2 x k , 2 x 1 , N x 2 , N x k , N
After generating the sample matrix, 400 Monte Carlo simulations (one simulation per row of the sample matrix) are carried out in Dymola (deterministic simulation) using a Functional Mock-up Interface (FMI) for co-simulation [77] to calculate the simulation output Y for each sample as:
Y = y 1   y 2   y N =   f ( x 1,1 , x 2,1 , , x k , 1 ) f ( x 1,2 , x 2,2 , , x k , 2 ) f ( x 1 , N , x 2 , N , , x k , N )
The FMI provides a definition of an interface standard to facilitate co-simulation and integrate the models into different modeling environments. The models can be encapsulated into a Functional Mock-up Unit (FMU) and used in other simulation tools.
These simulation runs constitute the database, which is randomly partitioned into two sets for training and testing meta-models. The training set (70% of simulation runs) is used to develop meta-models. The remaining 30% of simulation runs are used as a testing set to assess the accuracy and reliability of the developed meta-models.
Four commonly used meta-models, i.e., MARS, PR, RF, and SVR based on the RBF kernel, are assessed. The performances of the developed meta-models are compared by calculating six metrics: MBE, NMBE, RMSE, CVRMSE, R2, and GOF (see Section 2.1).
In the third stage of the SA, the variance-based SA by Sobol’ is applied to impart more quantitative information concerning parameter significance. In the Sobol’ method, the contribution of each uncertain parameter to the total variance of the output is evaluated by first-order ( S i ) and total-order ( S T i ) Sobol’ indices using the Sobol’ sequence sampling (see Section 2.4). Approximately 23,040 simulation runs are carried out using the MARS model as a meta-model of the original complex Dymola model. The most sensitive parameters to the simulation model output (annual total energy consumption) are identified. The uncertainty and sensitivity analysis approach used in the second part of the methodology is shown in Figure 2.

2.1. Statistical Performance Metrics

Six statistical metrics are used to verify and quantify the accuracy of the model. Table 2 presents the list of metrics used in this study.
MBE indicates the average of the model errors and is calculated using Equation (3), where r is the reference (or original) value, p is the predicted value, and n is the total number of data points. A value of MBE close to zero is desirable. NMBE is calculated by dividing the MBE by the mean of the reference values ( r ¯ ) and is defined according to Equation (4). RMSE indicates the square root of the average of the squared model errors and is introduced in Equation (5). CVRMSE is a normalized form of RMSE and is calculated by dividing the RMSE by the mean of the reference values ( r ¯ ) and is defined according to Equation (6).
The CVRMSE and the NMBE are dimensionless metrics. ASHRAE Guideline 14 [78], Measurement and Verification of the Federal Energy Management Program (FEMP) [79], and International Performance Measurement and Verification Protocol (IPMVP®) [80] prescribe two metrics, NMBE and CVRMSE, for the evaluation of the model’s accuracy. The Goodness of Fit (GOF) proposed by ASHRAE 1051-RP [81] combines the two normalized metrics NMBE and CV (RMSE) into a single metric, which is calculated according to Equation (7). The R2 indicates how well prediction values fit reference values and is described according to Equation (8). The value of R2 ranges between 0 and 1. The closer R2 is to 1, the greater the accuracy of the prediction model.

2.2. Morris’ Method

A generalized model is given by a mathematical function according to Equation (9):
Y = f ( X )
where X is a k × N matrix with N samples of k input parameters under consideration (see Equation (1)), and Y is a vector of the corresponding model output (see Equation (2)).
To detect the most important input parameters from among a large number, the screening-based SA proposed by Morris in combination with a factorial sampling plan is implemented. The aim is to determine the subset of parameters that can be fixed at their baseline point values without considerably decreasing the output variance. Morris’ method is generally used to discover which input parameters may be considered as (a) negligible or unimportant, (b) linear and additive, or (c) non-linear or involved in interactions with other parameters [82,83]. The considerable advantage of the screening-based SA by Morris is that the assessment of the influence of input parameters can be carried out with a small sample size. In cases of large numbers of parameters, Morris’ method is a felicitous screening-based technique prior to performing variance-based SA to alleviate the computational cost [83].
The samples are representative of the parameter space Ω, which is discretized by dividing the range of each parameter into a number of p levels with a predetermined distance of equal size ∆. Each input parameter varies once with step ∆ that builds t trajectories of (k + 1) points in Ω. Evaluation of the influence of each parameter on the model output is carried out by calculating a so-called elementary effect (EE), according to Equation (10).
E E i = f x 1 , , x i 1 , x i ± , , x k f ( x 1 , , x k )  
The computation of each elementary effect requires two sample points. Therefore, the number of required simulation runs (N) is given by N = t × (k + 1), where t is the number of elementary effects per input parameter and k is the number of input parameters. The values between 4 and 10 are recommended for t [34,73]. Due to the large number of parameters in this study (k = 105), the number of trajectories t = 4 is used. Therefore, 4 × (105 + 1) = 424 simulation runs are required to perform the screening-based SA by Morris.
The sensitivity of the model output to each input parameter is analyzed by two statistical measures: the mean value (μ) and the standard deviation (σ) of the elementary effects, as defined in Equations (11) and (12):
μ i = 1 t j = 1 t E E i j
σ i = 1 t 1 j = 1 t E E i j μ i 2
where E E i j represents the elementary effects of the input parameter i computed along trajectory j ( E E i j , i = 1, 2, …, k and j = 1, 2, …, t). The mean value of elementary effects (μ) indicates a measure of the overall effect (sensitivity) of an input parameter, while the standard deviation of elementary effects (σ) indicates a measure of interactions of an input parameter with other parameters, i.e., non-linear effects of an input parameter. These two sensitivity measures are basically applied as indicators of which parameters can be considered important [73]. Campolongo et al. [82] proposed using the mean of the absolute value of elementary effects (μ*), as presented in Equation (13), since positive and negative values of elementary effects can cancel each other out (especially where input parameters are involved in interaction effects).
μ i * = 1 t j = 1 t E E i j  
The μ* is used as an indicator to rank parameters in order of importance.

2.3. Sobol’ Method

Sobol’ method is one of the variance-based GSAs proposed by Sobol’ [74]. It is based on the decomposition of the total model output variance and can consider the main effects as well as the interaction effects of parameters (uncertainty in single parameters as well as uncertainty in combinations of parameters). The key aim of this method is that the contribution of each uncertain parameter to the total variance of the model output is quantified by specific order sensitivity indices, including first-order Sobol’ index ( S i ), second-order Sobol’ index ( S i j ), …, and total-order Sobol’ index ( S T i ).
The variance of a model output V(Y) can be defined as a sum of the variances of first-order and higher-order functions:
V Y = j = 1 k V i + j = 1 k i < j k V i , j + + V 1,2 , , k
where V i represents the variance contribution of first-order interaction effects for the input parameter i and V i , j indicates the variance contribution of second-order interaction effects between two parameters.
The first-order Sobol’ sensitivity index ( S i ) is written as follows:
S i = V X i E X ~ i Y X i V ( Y )
where E X ~ i Y X i represents the mean of the model output Y taken over all values of the sample matrix of input parameters X when the parameter X i remains fixed and all other input parameters are changed within the defined range. The term X ~ i indicates a matrix of all input parameters, but X i . The first-order index S i is applied to determine the most influential parameters on the variance of the model output. However, it does not describe the interactions of an input parameter with other parameters. Thus, a low S i does not express that the parameter X i is noninfluential, since there can be higher order interaction effects.
The total-order Sobol’ sensitivity index ( S T i ) considers all interaction effects of the input parameter X i with other parameters (first and all higher order interaction effects) and can be defined as:
S T i = E X ~ i V X i Y X ~ i V ( Y )
The total-order index ( S T i ) is used to determine unimportant (negligible) parameters, i.e., parameters without significantly affecting the variance of the model output.
The sampling from parameter distributions is designed using the Sobol’ sequence, which is a quasi-random low-discrepancy sampling technique, to generate uniform samples of the parameter space Ω. Carrying out variance-based SA by Sobol’ method for a model with k input parameters entails a computational cost of N × (2k + 2) simulation runs, where N is the sample size used for the Monte Carlo simulation [49]. In this study, to perform the robust SA, 23,040 simulation runs are executed using the MARS meta-model.

2.4. Meta-Modeling

The PR is an extension of the linear model that accommodates non-linear relationships between input parameters and simulation outputs. Since polynomial functions of parameters are included in the regression model, it is known as polynomial regression. In this study, the second-order polynomial model is applied, which can be described in Equation (17):
Y ^ = β 0 + i = 1 k β i x i + i = 1 k β i i x i 2 + i = 1 k j = 1 ,   i < j k β i j x i x j  
where Y ^ is the predicted value, x is the input parameter, and β is the regression coefficient. For k parameters, the second-order polynomial model of Equation (17) will have p = (k + 1) × (k + 2)/2 terms [84].
The RF is a tree-based machine learning algorithm that applies trees as building blocks to construct prediction models [85]. The RF model breaks down the dataset into tree-structure forms and uses averaging to estimate the output of all trees. In this study, the number of trees in the forest was set at 600.
The Support Vector Machine (SVM) is an artificial intelligence (AI) algorithm developed by Vapnik [86]. The goal of the SVM is to minimize the generalization error bound and, furthermore, to obtain generalized performance. That is, the SVM attempts to minimize the deviation (ε) of the prediction function from the target. In this study, the RBF is used as a kernel function [87].
The MARS [88] is a nonparametric, automated regression-based machine learning algorithm that uses labeled samples to model the complex non-linear relationship between input parameters (independent variables) and simulation outputs (dependent variables) and, subsequently, to predict outputs (targets). It can be considered a generalization of stepwise linear regression and has an outstanding ability to tackle high-dimensional problems with a large number of input parameters [62,68,69]. The MARS algorithm can be used to identify the parameters that have a considerable influence on simulation outputs, which is essential in the early stages of the building design process.
MARS uses spline basis functions of the form x v t + and t x v + , defined according to Equations (18) and (19).
x v t + = x v t     i f   x v > t   0           o t h e r w i s e
t x v + = t x v     i f   x v < t   0           o t h e r w i s e
where x v is the v-th independent variable (input parameter), t is the knot of the spline, and the subscript “+” means (.)+ takes the positive part of the inner argument. Each function is piecewise linear, with a knot at the value t. Figure 3 presents the spline basis functions (t = 0.5).
The general expression of the MARS model can be described according to Equation (20):
Y ^ = f ^ X = β 0 + m = 1 M β m h m X
where Y ^ is the dependent variable (output) predicted by the MARS model by means of a function f ^ X , β 0 is an initial constant, β m is the coefficient of m-th basis function, h m X is the m-th basis function, which can be a single spline basis function or a product of two or more such functions, and M is the number of basis functions, hm. The basis functions h m X take the form as follows:
h m X = k = 1 K M s k , m ( x v ( k , m ) t k , m )
where K M is the number of knots (number of interaction order in m-th basis function), s k , m has the value of either +1 or −1, which determines the right/left sense of the associated step function, and t k , m indicates the knot location on each of the corresponding variables.

3. Case Study

A hypothetical building with one single thermal zone (shoebox-shaped building) and one singular south-facing window without shading is considered an initial case study. The building has dimensions of 2 m in width, 5 m in length, and a floor-to-ceiling height of 3 m, totaling its interior volume at 30 m3. The window is centered in the southern wall and has a total surface area of 3 m2 (1.5 m × 2 m), as shown in Figure 4. All surfaces are modeled as exterior walls. The material properties of the building envelope used in this study can be found in Table 3. The same boundary conditions are applied to all surfaces, and an initial temperature of 20 °C is defined for the interior air volume and wall instances. The infiltration rate of 0.5 air changes per hour is assumed for the building air volume. No heat recovery is considered. The internal heat gains are set to 200 W. The radiative and convective portions of internal heat gains are taken as 60% and 40%, respectively.
The building is assumed to be heated and cooled by an ideal HVAC system. The maximum heating power of the HVAC system is limited to 1000 W, while the maximum cooling power is capped at 1200 W. The heating set-point temperature is 20 °C, and the cooling set-point temperature is set to 28 °C. The proposed case study is modeled in Dymola using the Modelica® language (Version 3.2.2) [89] and simulated for an entire year using a time step resolution of one hour. The investigated building model is controlled based on the heating/cooling set-point temperature.
For this analysis, the simulations are carried out using the weather data of Denver-Stapleton (Denver, CO, USA), which is a cool-dry climate (zone 5B) according to the ANSI/ASHRAE/IES Standard 90.1 [90]. The original weather data were provided in the EnergyPlus Weather (*.epw) format, which is converted to the Modelica timeseries script (*.mos) before running the simulation in Dymola. The climate data consists of hourly values of solar radiation and meteorological conditions for a 1-year period. Figure 5 shows the outdoor dry bulb temperature and the global horizontal irradiance (GHI) for one year.

4. Results and Discussion

This section presents the results of the application of the proposed framework with respect to the investigated case study of a building with one single thermal zone.

4.1. Validation of Deterministic Simulations

The scope of this section is to assess the results of two deterministic simulations (reference and synthetic), i.e., to reveal the inconsistencies between the two simulation results. In the reference simulation, the aforementioned case study was used to carry out the simulation. In the synthetic simulation, the virtual building model, which has different input parameters related to the thermophysical and optical properties of the building envelope, infiltration rate, internal heat gains, and HVAC system, was simulated.
Both deterministic simulations are performed for one year at an hourly resolution using the same simulation setup.
Figure 6 compares the results of heating and cooling loads. It can be qualitatively observed that there are discrepancies between reference and synthetic models for both heating and cooling load predictions. There are no heating loads from March to November. The maximum heating load predictions have values of 581.7 W and 534.5 W (in January) for the reference model and synthetic model, respectively. The cooling load predictions have maximum values of 906.8 W and 844.9 W (in July) for the synthetic model and reference model, respectively.
To quantify and ascertain the discrepancies between reference and synthetic models, six statistical metrics, described in Section 2.1, were used. The results of the investigated performance metrics for the cooling load are shown in Table 4. It indicates deviations in reference and synthetic simulation results. To discover the most important model parameters that contribute enormously to these discrepancies, a meta-model-based SA was conducted.

4.2. Sensitivity Analysis Using Morris’ Method

The screening-based SA using Morris’ method was performed to determine the most influential input parameters on annual heating, cooling, and total consumption. A total of 105 uncertain parameters, including thermophysical and optical properties, infiltration rate, internal heat gains, and HVAC system, were selected. The factorial sampling plan proposed by Morris was applied to build 424 samples of the input parameter.
Figure 7 illustrates the results of Morris’ sensitivity analysis of building energy simulations for the investigated case study.
Each point in scatter plots indicates the impact of each input parameter included in the investigation on the selected simulation output. The horizontal axis represents the mean of the absolute value of the elementary effects (μ*), which indicates the overall importance of an input parameter. The vertical axis represents the standard deviation of the elementary effects (σ), which describes the total interactions of an input parameter with other parameters involved in the analysis (non-linear effects of parameters). It is also worth pointing out that according to Equation (11) the elementary effect has the same unit as the simulation output, which is the annual energy consumption (kWh/year) in this study. The ratio σ/μ* is used as an indicator for the assessment of the linearity (and monotonicity) of parameters. The ratio σ/μ* = 0 is considered for the linear behavior of parameters. Three slope lines (σ/μ* = 0.1, 0.5, and 1) can be plotted in the scatter plot (Figure 7) to classify parameters into four zones: almost linear (below σ/μ* = 0.1), monotonic (0.1 < σ/μ* < 0.5), almost monotonic (0.5 < σ/μ* < 1.0), and non-linear/non-monotonic (σ/μ* > 1) [54].
The mean of the absolute value of the elementary effects (μ*) is used for ranking the input parameters in accordance with their importance. The higher the μ* value, the more important (and sensitive) the parameter. Figure 8 shows the ranking of 44 input parameters from an initial set of 105 for the simulation output of annual total consumption. Table 5 shows the selected input parameters as well as their rank with the corresponding μ* and parameter index. In this analysis, cooling set-point temperature was identified as the most influential parameter with the highest value of μ*, having a monotonic effect on the annual total consumption, followed by the g-value of the window, window area, and heating set-point temperature.

4.3. Meta-Modeling

In this section, four meta-models (MARS, PR, RF, and SVR) based on the RBF kernel were developed to evaluate their performance. LHS, in combination with Monte Carlo simulation, was used to generate a dataset for developing meta-models for annual total consumption.
Trading off the accuracy and the computational cost, 400 samples for each of the 44 uncertain input parameters (see Table 5) were created using uniform probability distributions. It is assumed that there is no correlation between input parameters. These simulation runs constitute the database, which is randomly partitioned into two sets for training (70% of simulation runs) and testing (30% of simulation runs) meta-models.
Six statistical metrics were applied to compare the performance of the developed meta-models. Table 6 shows that MARS considerably outperforms all other meta-models with respect to annual total consumption predictions.
The R2 values for MARS and PR are very close together at 0.99 and 0.97, respectively. However, the MBE value for MARS (−0.73 kWh) differs significantly from that for the PR model (−13.32 kWh). It can be interpreted that there may be a bias (i.e., systematic error) in the predicted values of the PR model. The negative value of MBE describes that the average of predicted values of the meta-model during the time period is higher than the average of Dymola values, which indicates the overprediction of the meta-model. Figure 9 graphically visualizes the values of CVRMSE, GOF, and R2 for each meta-model examined, which confirms the superior performance of the MARS algorithm.
Figure 10 shows a scatter plot created between the MARS meta-model and the Dymola model for the whole generated sample. The x-axis represents Dymola simulation results, and the y-axis represents MARS predictions. The red line indicates the perfect fit line (y = x). It can be shown that the investigated MARS model provides an accurate prediction for the annual total consumption and is in very good agreement. The MARS is considered a meta-model of the original Dymola model for further SA, which is described in the next section.

4.4. Sensitivity Analysis Using Sobol’ Method

The Saltelli’s extension of the Sobol’ sequence was used to generate uniform samples of parameter space Ω. With a sample size of N = 256 and k = 44 uncertain input parameters (see Table 5), 23,040 evaluation runs (N × (2k + 2)) were carried out using the MARS meta-model to perform the variance-based SA using the Sobol’ method. The most sensitive parameters to the model output (annual total consumption) are shown in Figure 11 as total effects ST from the Sobol’ method. The total sensitivity index ST indicates the total effects (first-order and all parameter interaction effects) of parameters on model output and can be considered a measure of negligible input parameters.
The ranking of the top 16 most influential input parameters (from an initial set of 105) with the corresponding total sensitivity index ST is given in Table 7. The remaining parameters exhibit very low ST values, i.e., they have no significant impact on the total variance of the model output.
It is observed that the first eight input parameters, i.e., cooling set-point temperature, g-value of the window, window area, heating set-point temperature, radiative portion of the internal heat gain, albedo, convective portion of the internal heat gain, and solar absorptance of the ceiling, are consistent with the SA from Morris’ method (Table 5).
Nevertheless, there is considerable disagreement concerning the solar absorptance of the southern wall, which is ranked higher than by Morris’ method. It can be inferred that the solar absorptance of the southern wall is involved in interaction effects with other input parameters or non-linear behavior.

5. Conclusions and Future Work

5.1. Conclusions

BEPS is a function of many input parameters. Due to large uncertainties and the complex non-linear behavior of parameters, a considerable mismatch can occur between BEPS output and reference data. Furthermore, carrying out stochastic simulations with a vast number of samples in BEPS tools is very time-consuming. In this paper, a novel four-stage SA framework based on the MARS meta-model is proposed to explore significant uncertain input parameters that the BEPS output (annual energy consumption) is most sensitive to and to provide an overview of critical parameters in a holistic context. The major contribution of this study can be formulated as follows:
  • In the first stage, two deterministic simulations are carried out. The comparative testing validation method is used to assess the deviation of the simulation results.
  • In the second stage, the screening-based SA using Morris’ method is performed to identify the most important input parameters, which have a significant influence on the model output (annual energy consumption). In this stage, the number of uncertain parameters in the investigated case study is reduced from 105 to 44 prior to performing variance-based SA. The selected input parameters are used for constructing the meta-model.
  • In the third stage, meta-modeling techniques are applied to tackle the computational costs of the variance-based SA, especially in cases of large numbers of time-consuming simulation runs. LHS, in combination with Monte Carlo simulation, is used to propagate uncertainties by running 400 simulation runs in the Dymola model and to generate a database for training and testing meta-models. A comparison study of the four meta-models, MARS, PR, RF, and SVR (based on the RBF kernel), is conducted. The results show that MARS is the most accurate meta-model in this study and provides a very precise prediction for the annual energy consumption compared to the original Dymola model.
  • Finally, in the fourth stage, the variance-based SA using Sobol’ method is performed as a final, robust stage of SA. MARS is applied to implement uncertainty propagation based on Monte Carlo simulations (23,040 simulation runs). Running the MARS meta-model can be implemented faster and reduce the computational cost of Sobol’ sensitivity indices in comparison with the direct simulation of complex building models.
Due to using meta-modeling techniques and implementing two GSAs (Morris and Sobol’ methods), the proposed framework provides the fast, reliable SA of highly sophisticated building models, which is required in the early building design process for robust decision-making. In addition, this framework considers a large set of uncertain parameters, which increases the reliability of uncertainty and sensitivity analysis.
It can be concluded that the cooling set-point temperature has a strong influence on the annual total consumption with the total-order Sobol’ index of S T = 0.76 for the examined case study. Other highly influential parameters are those related to window parameters, i.e., the g-value and area of the window. In addition to this, this framework highlights that performing the screening-based SA (Morris’ method) and MARS meta-model can reduce the computational cost of Sobol’ sensitivity indices dramatically in comparison with the direct simulation of the original, complex building model. Therefore, the interpretation and evaluation of model outputs can be made faster and more robust.

5.2. Future Work

The proposed framework is based on weather data, which represents a cool-dry climate. The SA results could significantly differ if other weather data (such as hot-dry climate) is used. Further research work is required to improve the framework concerning several weather data points. Furthermore, MARS, in combination with Sobol’ method, is used in this study to explore the impacts of uncertain input parameters on annual energy consumption. In future work, other meta-models in combination with different variance-based SAs (such as FAST) will be investigated to compare with the developed SA framework. However, machine learning algorithms are often uninterpretable black-box models and lack transparency. Indeed, for these algorithms, it is extremely hard to explain how predictions were made. Further research is required to apply explainable artificial intelligence (XAI) approaches, such as Local Interpretable Model-Agnostic Explanations (LIME) [91] and Shapley Additive exPlanation (SHAP) [92], to improve the interpretability of machine learning-based systems.

Author Contributions

Conceptualization, A.N., J.F. and C.v.T.; methodology, A.N.; software, A.N.; validation, A.N.; formal analysis, A.N.; investigation, A.N.; resources, A.N. and J.F.; data curation, A.N.; writing—original draft preparation, A.N.; writing—review and editing, J.F. and C.v.T.; visualization, A.N.; supervision, J.F. and C.v.T.; project administration, J.F.; funding acquisition, J.F. and C.v.T. All authors have read and agreed to the published version of the manuscript.

Funding

Parts of this research were funded by the German Federal Ministry for Economic Affairs and Climate Action under grant number 03ET1570B.

Data Availability Statement

Data are available upon request.

Acknowledgments

The authors would like to thank all project partners and contributors to the research project SimQuality (grant number 03ET1570B).

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. List of input parameters and assigned uncertainty ranges used for the screening-based sensitivity analysis by Morris’ method.
Table A1. List of input parameters and assigned uncertainty ranges used for the screening-based sensitivity analysis by Morris’ method.
IndexParameterBaseline PointParameter Range
(Lower and Upper Bounds)
Wall south
1Conductivity layer 1 (plaster)0.140 W/(m·K)[0.126, 0.154]
2Density layer 1 (plaster)800 kg/m3[720, 880]
3Specific heat capacity layer 1 (plaster)1500 J/(kg·K)[1350, 1650]
4Thickness layer 1 (plaster)0.005 m[0.0045, 0.0055]
5Conductivity layer 2 (insulation)0.040 W/(m·K)[0.036, 0.044]
6Density layer 2 (insulation)50 kg/m3[45, 55]
7Specific heat capacity layer 2 (insulation)1000 J/(kg·K)[900, 1100]
8Thickness layer 2 (insulation)0.100 m[0.09, 0.11]
9Conductivity layer 3 (concrete)1.200 W/(m·K)[1.08, 1.32]
10Density layer 3 (concrete)2000 kg/m3[1800, 2200]
11Specific heat capacity layer 3 (concrete)1000 J/(kg·K)[900, 1100]
12Thickness layer 3 (concrete)0.200 m[0.18, 0.22]
13Internal convective heat transfer coefficient2.5 W/(m2·K)[2.25, 2.75]
14External convective heat transfer coefficient8 W/(m2·K)[7.2, 8.8]
15Solar absorptance coefficient of external wall surface0.6[0.54, 0.66]
16Longwave emission coefficient of internal wall surface0.9[0.81, 0.99]
Wall west
17Conductivity Layer 1 (plaster)0.140 W/(m·K)[0.126, 0.154]
18Density layer 1 (plaster)800 kg/m3[720, 880]
19Specific heat capacity layer 1 (plaster)1500 J/(kg·K)[1350, 1650]
20Thickness layer 1 (plaster)0.005 m[0.0045, 0.0055]
21Conductivity layer 2 (insulation)0.040 W/(m·K)[0.036, 0.044]
22Density layer 2 (insulation)50 kg/m3[45, 55]
23Specific heat capacity layer 2 (insulation)1000 J/(kg·K)[900, 1100]
24Thickness layer 2 (insulation)0.100 m[0.09, 0.11]
25Conductivity layer 3 (concrete)1.200 W/(m·K)[1.08, 1.32]
26Density layer 3 (concrete)2000 kg/m3[1800, 2200]
27Specific heat capacity layer 3 (concrete)1000 J/(kg·K)[900, 1100]
28Thickness layer 3 (concrete)0.200 m[0.18, 0.22]
29Internal convective heat transfer coefficient2.5 W/(m2·K)[2.25, 2.75]
30External convective heat transfer coefficient8 W/(m2·K)[7.2, 8.8]
31Solar absorptance coefficient of external wall surface0.6[0.54, 0.66]
32Longwave emission coefficient of internal wall surface0.9[0.81, 0.99]
Wall north
33Conductivity Layer 1 (plaster)0.140 W/(m·K)[0.126, 0.154]
34Density layer 1 (plaster)800 kg/m3[720, 880]
35Specific heat capacity layer 1 (plaster)1500 J/(kg·K)[1350, 1650]
36Thickness layer 1 (plaster)0.005 m[0.0045, 0.0055]
37Conductivity layer 2 (insulation)0.040 W/(m·K)[0.036, 0.044]
38Density layer 2 (insulation)50 kg/m3[45, 55]
39Specific heat capacity layer 2 (insulation)1000 J/(kg·K)[900, 1100]
40Thickness layer 2 (insulation)0.100 m[0.09, 0.11]
41Conductivity layer 3 (concrete)1.200 W/(m·K)[1.08, 1.32]
42Density layer 3 (concrete)2000 kg/m3[1800, 2200]
43Specific heat capacity layer 3 (concrete)1000 J/(kg·K)[900, 1100]
44Thickness layer 3 (concrete)0.200 m[0.18, 0.22]
45Internal convective heat transfer coefficient2.5 W/(m2·K)[2.25, 2.75]
46External convective heat transfer coefficient8 W/(m2·K)[7.2, 8.8]
47Solar absorptance coefficient of external wall surface0.6[0.54, 0.66]
48Longwave emission coefficient of internal wall surface0.9[0.81, 0.99]
Wall east
49Conductivity Layer 1 (plaster)0.140 W/(m·K)[0.126, 0.154]
50Density layer 1 (plaster)800 kg/m3[720, 880]
51Specific heat capacity layer 1 (plaster)1500 J/(kg·K)[1350, 1650]
52Thickness layer 1 (plaster)0.005 m[0.0045, 0.0055]
53Conductivity layer 2 (insulation)0.040 W/(m·K)[0.036, 0.044]
54Density layer 2 (insulation)50 kg/m3[45, 55]
55Specific heat capacity layer 2 (insulation)1000 J/(kg·K)[900, 1100]
56Thickness layer 2 (insulation)0.100 m[0.09, 0.11]
57Conductivity layer 3 (concrete)1.200 W/(m·K)[1.08, 1.32]
58Density layer 3 (concrete)2000 kg/m3[1800, 2200]
59Specific heat capacity layer 3 (concrete)1000 J/(kg·K)[900, 1100]
60Thickness layer 3 (concrete)0.200 m[0.18, 0.22]
61Internal convective heat transfer coefficient2.5 W/(m2·K)[2.25, 2.75]
62External convective heat transfer coefficient8 W/(m2·K)[7.2, 8.8]
63Solar absorptance coefficient of external wall surface0.6[0.54, 0.66]
64Longwave emission coefficient of internal wall surface0.9[0.81, 0.99]
Ceiling
65Conductivity Layer 1 (plaster)0.140 W/(m·K)[0.126, 0.154]
66Density layer 1 (plaster)800 kg/m3[720, 880]
67Specific heat capacity layer 1 (plaster)1500 J/(kg·K)[1350, 1650]
68Thickness layer 1 (plaster)0.005 m[0.0045, 0.0055]
69Conductivity layer 2 (insulation)0.040 W/(m·K)[0.036, 0.044]
70Density layer 2 (insulation)50 kg/m3[45, 55]
71Specific heat capacity layer 2 (insulation)1000 J/(kg·K)[900, 1100]
72Thickness layer 2 (insulation)0.100 m[0.09, 0.11]
73Conductivity layer 3 (concrete)1.200 W/(m·K)[1.08, 1.32]
74Density layer 3 (concrete)2000 kg/m3[1800, 2200]
75Specific heat capacity layer 3 (concrete)1000 J/(kg·K)[900, 1100]
76Thickness layer 3 (concrete)0.200 m[0.18, 0.22]
77Internal convective heat transfer coefficient2.5 W/(m2·K)[2.25, 2.75]
78External convective heat transfer coefficient8 W/(m2·K)[7.2, 8.8]
79Solar absorptance coefficient of external wall surface0.6[0.54, 0.66]
80Longwave emission coefficient of internal wall surface0.9[0.81, 0.99]
Floor
81Conductivity Layer 1 (plaster)0.140 W/(m·K)[0.126, 0.154]
82Density layer 1 (plaster)800 kg/m3[720, 880]
83Specific heat capacity layer 1 (plaster)1500 J/(kg·K)[1350, 1650]
84Thickness layer 1 (plaster)0.005 m[0.0045, 0.0055]
85Conductivity layer 2 (insulation)0.040 W/(m·K)[0.036, 0.044]
86Density layer 2 (insulation)50 kg/m3[45, 55]
87Specific heat capacity layer 2 (insulation)1000 J/(kg·K)[900, 1100]
88Thickness layer 2 (insulation)0.100 m[0.09, 0.11]
89Conductivity layer 3 (concrete)1.200 W/(m·K)[1.08, 1.32]
90Density layer 3 (concrete)2000 kg/m3[1800, 2200]
91Specific heat capacity layer 3 (concrete)1000 J/(kg·K)[900, 1100]
92Thickness layer 3 (concrete)0.200 m[0.18, 0.22]
93Internal convective heat transfer coefficient2.5 W/(m2·K)[2.25, 2.75]
94External convective heat transfer coefficient8 W/(m2·K)[7.2, 8.8]
95Solar absorptance coefficient of external wall surface0.6[0.54, 0.66]
96Longwave emission coefficient of internal wall surface0.9[0.81, 0.99]
Window properties
97Total surface area (window glazing + frame)3 m2[2.70, 3.30]
98U-value (window glazing + frame)1.10 W/(m2·K)[0.99, 1.21]
99g-value (SHGC—Solar Heat Gain Coefficient)0.6[0.54, 0.66]
Operation parameters
100Infiltration rate0.5 1/h[0.45, 0.55]
101Convective internal heat gains80 W[72, 88]
102Radiative internal heat gains120 W[108, 132]
103Albedo (ground reflectance)0.2[0.18, 0.22]
HVAC system
104Heating set-point temperature20 °C[18, 22]
105Cooling set-point temperature28 °C[25.2, 30.8]

References

  1. Fumo, N. A review on the basics of building energy estimation. Renew. Sustain. Energy Rev. 2014, 31, 53–60. [Google Scholar] [CrossRef]
  2. Borgstein, E.; Lamberts, R.; Hensen, J. Evaluating energy performance in non-domestic buildings: A review. Energy Build. 2016, 128, 734–755. [Google Scholar] [CrossRef]
  3. de Wilde, P. The gap between predicted and measured energy performance of buildings: A framework for investigation. Autom. Constr. 2014, 41, 40–49. [Google Scholar] [CrossRef]
  4. Van Dronkelaar, C.; Dowson, M.; Burman, E.; Spataru, C.; Mumovic, D. A Review of the Regulatory Energy Performance Gap and Its Underlying Causes in Non-domestic Buildings. Front. Mech. Eng. 2016, 1, 17. [Google Scholar] [CrossRef]
  5. Bowman, N.T.; Lomas, K.J. Empirical validation of dynamic thermal computer models of buildings. Build. Serv. Eng. Res. Technol. 1985, 6, 153–162. [Google Scholar] [CrossRef]
  6. Judkoff, R. Validation of building energy analysis simulation programs at the solar energy research institute. Energy Build. 1988, 10, 221–239. [Google Scholar] [CrossRef]
  7. Judkoff, R.; Wortman, D.; O’Doherty, B.; Burch, J. Methodology for Validating Building Energy Analysis Simulations; Technical Report; National Renewable Energy Lab (NREL): Golden, CO, USA, 2008. [Google Scholar]
  8. Lauster, M.; Teichmann, J.; Fuchs, M.; Streblow, R.; Mueller, D. Low order thermal network models for dynamic simulations of buildings on city district scale. Build. Environ. 2014, 73, 223–231. [Google Scholar] [CrossRef]
  9. Roberti, F.; Oberegger, U.F.; Gasparella, A. Calibrating historic building energy models to hourly indoor air and surface temperatures: Methodology and case study. Energy Build. 2015, 108, 236–243. [Google Scholar] [CrossRef]
  10. Ferroukhi, M.Y.; Abahri, K.; Belarbi, R.; Limam, K.; Nouviaire, A. Experimental validation of coupled heat, air and moisture transfer modeling in multilayer building components. Heat Mass Transf. 2016, 52, 2257–2269. [Google Scholar] [CrossRef]
  11. Zhang, R.; Sun, K.; Hong, T.; Yura, Y.; Hinokuma, R. A novel Variable Refrigerant Flow (VRF) heat recovery system model: Development and validation. Energy Build. 2018, 168, 399–412. [Google Scholar] [CrossRef]
  12. Kim, J.; Frank, S.; Im, P.; Braun, J.E.; Goldwasser, D.; Leach, M. Representing Small Commercial Building Faults in EnergyPlus, Part II: Model Validation. Buildings 2019, 9, 239. [Google Scholar] [CrossRef]
  13. Lee, S.H.; Hong, T. Validation of an inverse model of zone air heat balance. Build. Environ. 2019, 161, 106232. [Google Scholar] [CrossRef]
  14. Im, P.; New, J.R.; Joe, J. Empirical validation of building energy modeling using flexible research platform. In Proceedings of the Building Simulation 2019: 16th Conference of IBPSA, Rome, Italy, 2–4 September 2019; pp. 4515–4521. [Google Scholar]
  15. Halimov, A.; Lauster, M.; Müller, D. Development and validation of PCM models integrated into the high order building model of Modelica library—Aixlib. In Proceedings of the Building Simulation 2019: 16th Conference of IBPSA, Rome, Italy, 2–4 September 2019; pp. 4698–4705. [Google Scholar] [CrossRef]
  16. Kalogeras, G.; Rastegarpour, S.; Koulamas, C.; Kalogeras, A.P.; Casillas, J.; Ferrarini, L. Predictive capability testing and sensitivity analysis of a model for building energy efficiency. Build. Simul. 2020, 13, 33–50. [Google Scholar] [CrossRef]
  17. Gutiérrez, G.V.; Ramos Ruiz, G.; Fernández Bandera, C. Empirical and Comparative Validation for a Building Energy Model Calibration Methodology. Sensors 2020, 20, 5003. [Google Scholar] [CrossRef]
  18. Eriksson, M.; Akander, J.; Moshfegh, B. Development and validation of energy signature method—Case study on a multi-family building in Sweden before and after deep renovation. Energy Build. 2020, 210, 109756. [Google Scholar] [CrossRef]
  19. Shi, H.; Chen, Q. Building energy management decision-making in the real world: A comparative study of HVAC cooling strategies. J. Build. Eng. 2021, 33, 101869. [Google Scholar] [CrossRef]
  20. Yi, Z.; Lv, Y.; Xu, D.; Xu, J.; Qian, H.; Zhao, D.; Yang, R. Energy saving analysis of a transparent radiative cooling film for buildings with roof glazing. Energy Built Environ. 2021, 2, 214–222. [Google Scholar] [CrossRef]
  21. Eguía-Oller, P.; Martínez-Mariño, S.; Granada-Álvarez, E.; Febrero-Garrido, L. Empirical validation of a multizone building model coupled with an air flow network under complex realistic situations. Energy Build. 2021, 249, 111197. [Google Scholar] [CrossRef]
  22. Magni, M.; Ochs, F.; Streicher, W. Comprehensive analysis of the influence of different building modelling approaches on the results and computational time using a cross-compared model as a reference. Energy Build. 2022, 259, 111859. [Google Scholar] [CrossRef]
  23. Loutzenhiser, P.G.; Manz, H.; Moosberger, S.; Maxwell, G.M. An empirical validation of window solar gain models and the associated interactions. Int. J. Therm. Sci. 2009, 48, 85–95. [Google Scholar] [CrossRef]
  24. Tabares-Velasco, P.C.; Griffith, B. Diagnostic test cases for verifying surface heat transfer algorithms and boundary conditions in building energy simulation programs. J. Build. Perform. Simul. 2012, 5, 329–346. [Google Scholar] [CrossRef]
  25. Mateus, N.M.; Pinto, A.; da Graça, G.C. Validation of EnergyPlus thermal simulation of a double skin naturally and mechanically ventilated test cell. Energy Build. 2014, 75, 511–522. [Google Scholar] [CrossRef]
  26. Alaidroos, A.; Krarti, M. Experimental validation of a numerical model for ventilated wall cavity with spray evaporative cooling systems for hot and dry climates. Energy Build. 2016, 131, 207–222. [Google Scholar] [CrossRef]
  27. Sandels, C.; Brodén, D.; Widén, J.; Nordström, L.; Andersson, E. Modeling office building consumer load with a combined physical and behavioral approach: Simulation and validation. Appl. Energy 2016, 162, 472–485. [Google Scholar] [CrossRef]
  28. Cattarin, G.; Pagliano, L.; Causone, F.; Kindinis, A. Empirical and comparative validation of an original model to simulate the thermal behaviour of outdoor test cells. Energy Build. 2018, 158, 1711–1723. [Google Scholar] [CrossRef]
  29. Nageler, P.; Schweiger, G.; Pichler, M.; Brandl, D.; Mach, T.; Heimrath, R.; Schranzhofer, H.; Hochenauer, C. Validation of dynamic building energy simulation tools based on a real test-box with thermally activated building systems (TABS). Energy Build. 2018, 168, 42–55. [Google Scholar] [CrossRef]
  30. Barone, G.; Buonomano, A.; Forzano, C.; Palombo, A. Building Energy Performance Analysis: An Experimental Validation of an In-House Dynamic Simulation Tool through a Real Test Room. Energies 2019, 12, 4107. [Google Scholar] [CrossRef]
  31. Nouri, A.; Frisch, J.; van Treeck, C. Statistical methodologies for verification of building energy performance simulation. In Proceedings of the Building Simulation 2021: 17th Conference of IBPSA, Bruges, Belgium, 1–3 September 2021. [Google Scholar] [CrossRef]
  32. Tian, Z.; Love, J.A.; Tian, W. Applying quality control in building energy modelling: Comparative simulation of a high performance building. J. Build. Perform. Simul. 2009, 2, 163–178. [Google Scholar] [CrossRef]
  33. Lucchino, E.C.; Gelesz, A.; Skeie, K.; Gennaro, G.; Reith, A.; Serra, V.; Goia, F. Modelling double skin façades (DSFs) in whole-building energy simulation tools: Validation and inter-software comparison of a mechanically ventilated single-story DSF. Build. Environ. 2021, 199, 107906. [Google Scholar] [CrossRef]
  34. Heiselberg, P.; Brohus, H.; Hesselholt, A.; Rasmussen, H.; Seinre, E.; Thomas, S. Application of sensitivity analysis in design of sustainable buildings. Renew. Energy 2009, 34, 2030–2036. [Google Scholar] [CrossRef]
  35. Spitz, C.; Mora, L.; Wurtz, E.; Jay, A. Practical application of uncertainty analysis and sensitivity analysis on an experimental house. Energy Build. 2012, 55, 459–470. [Google Scholar] [CrossRef]
  36. Strachan, P.; Monari, F.; Kersken, M.; Heusler, I. IEA Annex 58: Full-scale Empirical Validation of Detailed Thermal Simulation Programs. Energy Procedia 2015, 78, 3288–3293. [Google Scholar] [CrossRef]
  37. Cattarin, G.; Pagliano, L.; Causone, F.; Kindinis, A.; Goia, F.; Carlucci, S.; Schlemminger, C. Empirical validation and local sensitivity analysis of a lumped-parameter thermal model of an outdoor test cell. Build. Environ. 2018, 130, 151–161. [Google Scholar] [CrossRef]
  38. Alongi, A.; Angelotti, A.; Mazzarella, L. A numerical model to simulate the dynamic performance of Breathing Walls. J. Build. Perform. Simul. 2021, 14, 155–180. [Google Scholar] [CrossRef]
  39. Raslan, R.; Davies, M. Results variability in accredited building energy performance compliance demonstration software in the UK: An inter-model comparative study. J. Build. Perform. Simul. 2010, 3, 63–85. [Google Scholar] [CrossRef]
  40. Shrestha, S.S.; Maxwell, G. Empirical validation of building energy simulation software: EnergyPlus. In Proceedings of the Building Simulation 2011: 12th Conference of IBPSA, Sydney, Australia, 14–16 November 2011. [Google Scholar]
  41. Tabares-Velasco, P.C.; Christensen, C.; Bianchi, M. Verification and validation of EnergyPlus phase change material model for opaque wall assemblies. Build. Environ. 2012, 54, 186–196. [Google Scholar] [CrossRef]
  42. Antretter, F.; Sauer, F.; Schöpfer, T.; Holm, A. Validation of a hygrothermal whole building simulation software. In Proceedings of the Building Simulation 2011: 12th Conference of IBPSA, Sydney, Australia, 14–16 November 2011. [Google Scholar]
  43. Zhu, D.; Hong, T.; Yan, D.; Wang, C. A detailed loads comparison of three building energy modeling programs: EnergyPlus, DeST and DOE-2.1E. Build. Simul. 2013, 6, 323–335. [Google Scholar] [CrossRef]
  44. Kim, Y.-J.; Yoon, S.-H.; Park, C.-S. Stochastic comparison between simplified energy calculation and dynamic simulation. Energy Build. 2013, 64, 332–342. [Google Scholar] [CrossRef]
  45. Buonomano, A. Code-to-Code Validation and Application of a Dynamic Simulation Tool for the Building Energy Performance Analysis. Energies 2016, 9, 301. [Google Scholar] [CrossRef]
  46. Moronis, A.; Koulamas, C.; Kalogeras, A. Validation of a monthly quasi-steady-state simulation model for the energy use in buildings. In Proceedings of the 2017 22nd IEEE International Conference on Emerging Technologies and Factory Automation (ETFA), Limassol, Cyprus, 12–15 September 2017; pp. 1–6. [Google Scholar] [CrossRef]
  47. Nouri, A.; Nicolai, A.; Krämer, B.; Hirth, S.; Agudelo, J.; Seifert, C.; Malhotra, A.; Madjidi, M.; Frisch, J.; van Treeck, C. Entwicklung von Qualitätsstandards für die energetische Gebäude- und Anlagensimulation als Planungswerkzeug. In Proceedings of the BauSim Conference 2020: 8th Conference of IBPSA, Graz, Austria, 23–25 September 2020; pp. 79–86. [Google Scholar] [CrossRef]
  48. Zakula, T.; Badun, N.; Ferdelji, N.; Ugrina, I. Framework for the ISO 52016 standard accuracy prediction based on the in-depth sensitivity analysis. Appl. Energy 2021, 298, 117089. [Google Scholar] [CrossRef]
  49. Saltelli, A.; Ratto, M.; Tarantola, S.; Campolongo, F. Update 1 of: Sensitivity Analysis for Chemical Models. Chem. Rev. 2012, 112, PR1–PR21. [Google Scholar] [CrossRef]
  50. Tian, W. A review of sensitivity analysis methods in building energy analysis. Renew. Sustain. Energy Rev. 2013, 20, 411–419. [Google Scholar] [CrossRef]
  51. Yang, J. Convergence and uncertainty analyses in Monte-Carlo based sensitivity analysis. Environ. Model. Softw. 2011, 26, 444–457. [Google Scholar] [CrossRef]
  52. Saltelli, A.; Annoni, P. How to avoid a perfunctory sensitivity analysis. Environ. Model. Softw. 2010, 25, 1508–1517. [Google Scholar] [CrossRef]
  53. Hygh, J.S.; DeCarolis, J.F.; Hill, D.B.; Ranjithan, S.R. Multivariate regression as an energy assessment tool in early building design. Build. Environ. 2012, 57, 165–175. [Google Scholar] [CrossRef]
  54. McLeod, R.S.; Hopfe, C.J.; Kwan, A. An investigation into future performance and overheating risks in Passivhaus dwellings. Build. Environ. 2013, 70, 189–209. [Google Scholar] [CrossRef]
  55. Kristensen, M.H.; Petersen, S. Choosing the appropriate sensitivity analysis method for building energy model-based investigations. Energy Build. 2016, 130, 166–176. [Google Scholar] [CrossRef]
  56. Mechri, H.E.; Capozzoli, A.; Corrado, V. USE of the ANOVA approach for sensitive building energy design. Appl. Energy 2010, 87, 3073–3083. [Google Scholar] [CrossRef]
  57. Pang, Z.; O’Neill, Z.; Li, Y.; Niu, F. The role of sensitivity analysis in the building performance analysis: A critical review. Energy Build. 2020, 209, 109659. [Google Scholar] [CrossRef]
  58. Garcia Sanchez, D.; Lacarrière, B.; Musy, M.; Bourges, B. Application of sensitivity analysis in building energy simulations: Combining first- and second-order elementary effects methods. Energy Build. 2014, 68, 741–750. [Google Scholar] [CrossRef]
  59. Van Gelder, L.; Das, P.; Janssen, H.; Roels, S. Comparative study of metamodelling techniques in building energy simulation: Guidelines for practitioners. Simul. Model. Pract. Theory 2014, 49, 245–257. [Google Scholar] [CrossRef]
  60. Rivalin, L.; Stabat, P.; Marchio, D.; Caciolo, M.; Hopquin, F. A comparison of methods for uncertainty and sensitivity analysis applied to the energy performance of new commercial buildings. Energy Build. 2018, 166, 489–504. [Google Scholar] [CrossRef]
  61. Wate, P.; Iglesias, M.; Coors, V.; Robinson, D. Framework for emulation and uncertainty quantification of a stochastic building performance simulator. Appl. Energy 2020, 258, 113759. [Google Scholar] [CrossRef]
  62. Hastie, T.; Tibshirani, R.; Friedman, J. The Elements of Statistical Learning: Data Mining, Inference, and Prediction, 2nd ed.; Springer Series in Statistics; Springer Science & Business Media: New York, NY, USA, 2009. [Google Scholar] [CrossRef]
  63. Chou, J.-S.; Bui, D.-K. Modeling heating and cooling loads by artificial intelligence for energy-efficient building design. Energy Build. 2014, 82, 437–446. [Google Scholar] [CrossRef]
  64. Hester, J.; Gregory, J.; Kirchain, R. Sequential early-design guidance for residential single-family buildings using a probabilistic metamodel of energy consumption. Energy Build. 2017, 134, 202–211. [Google Scholar] [CrossRef]
  65. Eisenhower, B.; O’neill, Z.; Narayanan, S.; Fonoberov, V.A.; Mezić, I. A methodology for meta-model based optimization in building energy models. Energy Build. 2012, 47, 292–301. [Google Scholar] [CrossRef]
  66. 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]
  67. Amasyali, K.; El-Gohary, N.M. A review of data-driven building energy consumption prediction studies. Renew. Sustain. Energy Rev. 2018, 81, 1192–1205. [Google Scholar] [CrossRef]
  68. Cheng, M.-Y.; Cao, M.-T. Accurately predicting building energy performance using evolutionary multivariate adaptive regression splines. Appl. Soft Comput. 2014, 22, 178–188. [Google Scholar] [CrossRef]
  69. Østergård, T.; Jensen, R.L.; Maagaard, S.E. A comparison of six metamodeling techniques applied to building performance simulations. Appl. Energy 2018, 211, 89–103. [Google Scholar] [CrossRef]
  70. Calleja Rodríguez, G.; Carrillo Andrés, A.; Domínguez Muñoz, F.; Cejudo López, J.M.; Zhang, Y. Uncertainties and sensitivity analysis in building energy simulation using macroparameters. Energy Build. 2013, 67, 79–87. [Google Scholar] [CrossRef]
  71. Bucking, S.; Zmeureanu, R.; Athienitis, A. A methodology for identifying the influence of design variations on building energy performance. J. Build. Perform. Simul. 2014, 7, 411–426. [Google Scholar] [CrossRef]
  72. Kim, Y.-J. Comparative study of surrogate models for uncertainty quantification of building energy model: Gaussian Process Emulator vs. Polynomial Chaos Expansion. Energy Build. 2016, 133, 46–58. [Google Scholar] [CrossRef]
  73. Morris, M.D. Factorial Sampling Plans for Preliminary Computational Experiments. Technometrics 1991, 33, 161. [Google Scholar] [CrossRef]
  74. Sobol′, I.M. Global sensitivity indices for nonlinear mathematical models and their Monte Carlo estimates. Math. Comput. Simul. 2001, 55, 271–280. [Google Scholar] [CrossRef]
  75. Burhenne, S.; Tsvetkova, O.; Jacob, D.; Henze, G.P.; Wagner, A. Uncertainty quantification for combined building performance and cost-benefit analyses. Build. Environ. 2013, 62, 143–154. [Google Scholar] [CrossRef]
  76. Helton, J.C.; Davis, F.J. Latin hypercube sampling and the propagation of uncertainty in analyses of complex systems. Reliab. Eng. Syst. Saf. 2003, 81, 23–69. [Google Scholar] [CrossRef]
  77. Andersson, C.; Åkesson, J.; Führer, C. PyFMI: A Python Package for Simulation of Coupled Dynamic Models with the Functional Mock-up Interface; Technical Report; Lund University: Lund, Sweden, 2016; Available online: https://lup.lub.lu.se/record/961a50eb-e4a8-43bc-80ac-d467eef26193 (accessed on 22 January 2024).
  78. ASHRAE Guideline 14. Measurement of Energy and Demand Savings; American Society of Heating, Refrigerating and Air-Conditioning Engineers: Atlanta, GA, USA, 2002. [Google Scholar]
  79. FEMP. M&V Guidelines: Measurement and Verification for Performance-Based Contracts, Version 4.0; U.S. Department of Energy Federal Energy Management Program: Washington, DC, USA, 2015. [Google Scholar]
  80. IPMVP®. International Performance Measurement and Verification Protocol: Concepts and Options for Determining Energy and Water Savings; Efficiency Valuation Organization (EVO): Washington, DC, USA, 2002; Volume 1. [Google Scholar]
  81. Reddy, T.; Maor, I. ASHRAE Research Project 1051-RP: Procedures for Reconciling Computer-Calculated Results with Measured Energy Data; American Society of Heating, Refrigerating and Air-Conditioning Engineers: Peachtree Corners, GA, USA, 2006. [Google Scholar]
  82. Campolongo, F.; Cariboni, J.; Saltelli, A. An effective screening design for sensitivity analysis of large models. Environ. Model. Softw. 2007, 22, 1509–1518. [Google Scholar] [CrossRef]
  83. Saltelli, A.; Ratto, M.; Andres, T.; Campolongo, F.; Cariboni, J.; Gatelli, D.; Saisana, M.; Tarantola, S. Global Sensitivity Analysis. The Primer; John Wiley & Sons, Ltd.: London, UK, 2007. [Google Scholar] [CrossRef]
  84. Jin, R.; Chen, W.; Simpson, T. Comparative studies of metamodelling techniques under multiple modelling criteria. Struct. Multidiscip. Optim. 2001, 23, 1–13. [Google Scholar] [CrossRef]
  85. Brieman, L. Random Forests. Mach. Learn. 2001, 45, 5–32. [Google Scholar] [CrossRef]
  86. Vapnik, V. The Nature of Statistical Learning Theory; Springer Science & Business Media: New York, NY, USA, 1995. [Google Scholar]
  87. Boser, B.E.; Guyon, I.M.; Vapnik, V.N. A training algorithm for optimal margin classifiers. In Proceedings of the Fifth Annual Workshop on Computational Learning Theory, Pittsburgh, PA, USA, 27–29 July 1992; Haussler, D., Ed.; pp. 144–152. [Google Scholar] [CrossRef]
  88. Friedman, J.H. Multivariate Adaptive Regression Splines. Ann. Stat. 1991, 19, 1–67. [Google Scholar] [CrossRef]
  89. Modelica®. A Unified Object-Oriented Language for Systems Modeling, Language Specification, Version 3.2.2 [Computer Software]. Available online: https://www.modelica.org/ (accessed on 30 October 2022).
  90. ANSI/ASHRAE/IES Standard 90.1; Energy Standard for Buildings Except Low-Rise Residential Buildings. American Society of Heating, Refrigerating and Air-Conditioning Engineers: Atlanta, GA, USA, 2019.
  91. Ribeiro, M.T.; Singh, S.; Guestrin, C. “Why Should I Trust You?” Explaining the predictions of any classifier. In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, San Francisco, CA, USA, 13–17 August 2016; Association for Computing Machinery: New York, NY, USA, 2016; pp. 1135–1144. [Google Scholar] [CrossRef]
  92. Lundberg, S.M.; Lee, S.I. A unified approach to interpreting model predictions. In Proceedings of the 31st International Conference on Neural Information Processing Systems, Red Hook, NY, USA, 4–9 December 2017; Curran Associates Inc.: Red Hook, NY, USA, 2017; pp. 4768–4777. [Google Scholar]
Figure 1. Schematic of validation approach.
Figure 1. Schematic of validation approach.
Energies 17 00695 g001
Figure 2. Schematic of uncertainty and sensitivity analysis approach.
Figure 2. Schematic of uncertainty and sensitivity analysis approach.
Energies 17 00695 g002
Figure 3. Basis functions (t = 0.5) used by MARS model.
Figure 3. Basis functions (t = 0.5) used by MARS model.
Energies 17 00695 g003
Figure 4. Geometry of the analyzed building.
Figure 4. Geometry of the analyzed building.
Energies 17 00695 g004
Figure 5. (a) Outdoor dry bulb temperature [°C]; (b) global horizontal irradiance [W/m2] used in Dymola.
Figure 5. (a) Outdoor dry bulb temperature [°C]; (b) global horizontal irradiance [W/m2] used in Dymola.
Energies 17 00695 g005
Figure 6. Deterministic simulation results of heating (a) and cooling (b) load [W] calculated with hourly dynamic simulation of reference and synthetic models.
Figure 6. Deterministic simulation results of heating (a) and cooling (b) load [W] calculated with hourly dynamic simulation of reference and synthetic models.
Energies 17 00695 g006
Figure 7. Morris’ method—sensitivity analysis of 105 parameters for annual heating (a); cooling (b); and total (c) consumptions.
Figure 7. Morris’ method—sensitivity analysis of 105 parameters for annual heating (a); cooling (b); and total (c) consumptions.
Energies 17 00695 g007
Figure 8. Ranking parameters by Morris’ method, considering an uncertainty of ±10% for all parameters. Note that in this study, thermal bridges are not considered.
Figure 8. Ranking parameters by Morris’ method, considering an uncertainty of ±10% for all parameters. Note that in this study, thermal bridges are not considered.
Energies 17 00695 g008
Figure 9. CVRMSE, GOF, and R2 of investigated meta-models.
Figure 9. CVRMSE, GOF, and R2 of investigated meta-models.
Energies 17 00695 g009
Figure 10. Scatter plot of MARS (R2 = 0.99).
Figure 10. Scatter plot of MARS (R2 = 0.99).
Energies 17 00695 g010
Figure 11. Ranking parameters by Sobol’ method (total sensitivity index).
Figure 11. Ranking parameters by Sobol’ method (total sensitivity index).
Energies 17 00695 g011
Table 1. Validation methods, sensitivity analysis, and statistical performance metrics found in the literature for the assessment of BEPS, where ‘X’ indicates methods/metrics involved in the reviewed literature.
Table 1. Validation methods, sensitivity analysis, and statistical performance metrics found in the literature for the assessment of BEPS, where ‘X’ indicates methods/metrics involved in the reviewed literature.
ReferenceYearValidation MethodsSensitivity AnalysisMetrics
EmpiricalComparativeAnalyticalMBENMBERMSECVRMSER2MAEOther
Judkoff et al. [7]2008XXXX
Loutzenhiser et al. [23]2009XX X XX
Tian et al. [32]2009 X X X
Raslan et al. [39]2010 X
Shrestha et al. [40]2011X
Tabares-Velasco et al. [41]2011 XX
Antretter et al. [42]2011XX
Tabares-Velasco et al. [24]2012XXX X
Spitz et al. [35]2012X X
Zhu et al. [43]2013 X
Kim et al. [44]2013 X X X
Lauster et al. [8]2014X X X X
Mateus et al. [25]2014X XX X X
Strachan et al. [36]2015X X
Roberti et al. [9]2015X X X X
Alaidroos and Krarti [26]2016X X XX
Sandels et al. [27]2016X X X
Ferroukhi et al. [10]2016X X X
Buonomano [45]2016 X
Moronis et al. [46]2017 X X
Cattarin et al. [28]2018XX X X
Cattarin et al. [37]2018X X
Zhang et al. [11]2018X X X
Nageler et al. [29]2018XX X X X
Kim et al. [12]2019X X X
Lee and Hong [13]2019X X X
Barone et al. [30]2019XX XX
Im et al. [14]2019X X X
Halimov et al. [15]2019XXX X X
Kalogeras et al. [16]2020 X X X X X
Gutiérrez et al. [17]2020XX X XXXX
Eriksson et al. [18]2020X X X XX
Nouri et al. [47]2020 X
Shi and Chen [19]2021XX X X XX
Yi et al. [20]2021X X X
Alongi et al. [38]2021X XX
Catto Lucchino et al. [33]2021XX XXXX
Nouri et al. [31]2021 X XXXXX
Zakula et al. [48]2021 X X XX X
Eguía-Oller et al. [21]2021X XXX X
Magni et al. [22]2022 X XXX X XX
Table 2. Statistical performance metrics.
Table 2. Statistical performance metrics.
MetricFormulaEquation
MBE k = 1 n ( r p ) n (3)
NMBE k = 1 n r p n × r ¯ × 100 (4)
RMSE k = 1 n ( r p ) 2 n (5)
CVRMSE k = 1 n ( r p ) 2 n r ¯ × 100 (6)
GOF 2 2 × C V R M S E 2 + N M B E 2 (7)
R2 1 k = 1 n ( r p ) 2 k = 1 n ( r r ¯ ) 2 (8)
Table 3. Material properties of the building envelope.
Table 3. Material properties of the building envelope.
Layer
Nr.
Components
(Outside to Inside)
Thickness
d [m]
Thermal Conductivity
λ W m · K
Density
ρ k g m 3
Specific Heat Capacity
C J k g · K
1Plaster0.0050.148001500
2Thermal insulation0.1000.04501000
3Concrete0.2001.2020001000
Table 4. Performance evaluation for cooling load.
Table 4. Performance evaluation for cooling load.
MetricsComparison of Reference and Synthetic Models
MBE [W]−32.28
NMBE [%]−15.18
RMSE [W]43.10
CVRMSE [%]20.27
GOF [%]17.91
R20.97
Table 5. Selected input parameters and their μ* by Morris’ method.
Table 5. Selected input parameters and their μ* by Morris’ method.
RankParameterUnitIndexμ*
1Cooling set-point [°C]105841.97
2g-value window-99300.56
3Window aream297220.35
4Heating set-point [°C]104129.51
5Internal heat gain (radiative)W102112.71
6Albedo-10380.10
7Internal heat gain (convective)W10177.70
8Solar absorptance ceiling-7958.59
9External convective coefficient wall westW/(m2·K)3057.46
10Solar absorptance west-3156.35
11External convective coefficient ceilingW/(m2·K)7854.48
12External convective coefficient wall eastW/(m2·K)6247.37
13Solar absorptance east-6346.00
14Infiltration rate1/h (ACH)10040.04
15Thickness floor layer 2m8827.09
16U-value windowW/(m2·K)9825.39
17Thickness ceiling layer 2m7222.43
18Conductivity ceiling layer 2W/(m·K)6920.78
19Conductivity floor layer 2W/(m·K)8518.80
20Conductivity wall east layer 2W/(m·K)5316.28
21External convective coefficient floorW/(m2·K)9414.77
22Internal convective coefficient wall eastW/(m2·K)6114.35
22External convective coefficient wall southW/(m2·K)1414.35
24Internal convective coefficient wall westW/(m2·K)2914.25
25Solar absorptance wall south-1514.04
26Thickness wall west layer 2M2413.58
27Conductivity wall west layer 2W/(m·K)2113.05
28Thickness wall north layer 2M4011.65
29Solar absorptance floor-9511.50
30External convective coefficient wall northW/(m2·K)4611.21
31Conductivity wall north layer 2W/(m·K)379.94
32Solar absorptance wall north-479.72
33Internal convective coefficient ceilingW/(m2·K)779.03
34Internal convective coefficient floorW/(m2·K)937.48
35Thickness wall east layer 2m566.27
36Internal convective coefficient wall northW/(m2·K)455.54
37Specific heat wall west layer 3J/(kg·K)274.12
38Specific heat wall east layer 3J/(kg·K)594.03
39Density wall east layer 3kg/m3583.72
40Density wall west layer 3kg/m3263.71
41Thickness wall south layer 2m83.52
42Conductivity wall south layer 2W/(m·K)53.30
43Internal convective coefficient wall southW/(m2·K)132.74
44Thickness ceiling layer 3m762.55
Table 6. Performance evaluation of meta-models.
Table 6. Performance evaluation of meta-models.
Meta-ModelMBE
[kWh]
NMBE
[%]
RMSE
[kWh]
CVRMSE
[%]
GOF
[%]
R2
MARS−0.73−0.0433.601.741.230.99
PR−13.32−0.6952.482.711.980.97
RF−7.94−0.41104.325.393.820.88
SVR (kernel: RBF)−9.28−0.48115.565.974.240.85
Table 7. Input parameter rankings and their total sensitivity index by Sobol’ method.
Table 7. Input parameter rankings and their total sensitivity index by Sobol’ method.
RankParameterUnitST
1Cooling set-point[°C]0.7656
2g-value window-0.0960
3Window aream20.0768
4Heating set-point[°C]0.0140
5Internal heat gain (radiative)W0.0113
6Albedo-0.0067
7Internal heat gain (convective)W0.0063
8Solar absorptance ceiling-0.0040
9External convective coeff. ceilingW/(m2 K)0.0034
10Solar absorptance west-0.0030
11External convective coeff. eastW/(m2 K)0.0025
12Solar absorptance east-0.0021
13External convective coeff. westW/(m2 K)0.0021
14Infiltration rate1/h (ACH)0.0020
15U-value windowW/(m2 K)0.0016
16Solar absorptance south-0.0012
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

Nouri, A.; van Treeck, C.; Frisch, J. Sensitivity Assessment of Building Energy Performance Simulations Using MARS Meta-Modeling in Combination with Sobol’ Method. Energies 2024, 17, 695. https://doi.org/10.3390/en17030695

AMA Style

Nouri A, van Treeck C, Frisch J. Sensitivity Assessment of Building Energy Performance Simulations Using MARS Meta-Modeling in Combination with Sobol’ Method. Energies. 2024; 17(3):695. https://doi.org/10.3390/en17030695

Chicago/Turabian Style

Nouri, Amin, Christoph van Treeck, and Jérôme Frisch. 2024. "Sensitivity Assessment of Building Energy Performance Simulations Using MARS Meta-Modeling in Combination with Sobol’ Method" Energies 17, no. 3: 695. https://doi.org/10.3390/en17030695

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