Next Article in Journal
A Comparison of the Quasi-Steady Assumption with Unsteady Effects on Tower Galloping Analysis
Previous Article in Journal
The Stability of Slopes and Building Structures Using an Energy Visualization Procedure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Boundary Conditions on the Estimation of Thermal Properties in Insulated Building Walls

1
LaSIE (UMR CNRS 7356)—La Rochelle University, Avenue Albert Einstein, 17000 La Rochelle, France
2
4evLab, La Rochelle University, CNRS, EDF, Avenue Albert Einstein, 17000 La Rochelle, France
3
TIPEE (Technological Platform), Rue Isabelle Autissier, 17140 Lagord, France
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(12), 3706; https://doi.org/10.3390/buildings14123706
Submission received: 18 October 2024 / Revised: 14 November 2024 / Accepted: 15 November 2024 / Published: 21 November 2024
(This article belongs to the Section Building Energy, Physics, Environment, and Systems)

Abstract

:
The objective of this study is to evaluate the ability of inverse techniques to estimate the resistance and the capacity of a highly insulated multilayer wall under real weather conditions. The wall is equipped with temperature sensors inside and on its inner and outer surfaces, and the boundary conditions have been measured over a 14-day period. Uncertainties on various parameters of the model are evaluated, including internal and external convective heat transfer coefficients (±20% and ±7 W.m-².K−1 respectively), external long-wave heat transfer coefficient (±0.15 W.m−2.K−1) and solar absorption coefficient (±0.06). A sensitivity analysis demonstrated the high correlation with some parameters defining the thermal performance of the walls (thermal resistance or capacity). A solution is proposed to limit the number of identified parameters, while allowing the identification of the thermal resistance and the thermal capacity of the walls. There are two cases: either the weather conditions are accurately measured (temperature, short- and long-wave radiation) and the thermal characteristics can be assessed, or intrusive sensors are installed, and the thermal characteristics can be evaluated more accurately.

1. Introduction

Many factors influence the energy consumption of buildings, but the building envelope plays a key role in the overall performance. Envelope heat losses are affected by conduction through opaque walls, by conduction and radiation through windows, by air leakage [1] and also by deterioration in performance over time [2].
Heat transfer by conduction can be characterized by the physical properties of each wall material. From these physical properties and from geometrical data, it is possible to determine the R-value and C-value of the wall components, which correspond to the thermal resistance and the thermal capacity, respectively [3]. Standard laboratory tests have been developed to evaluate material properties as described in [4,5]. However, these methods cannot be applied to the wall of a constructed building without causing damage.
Alternative methods have been developed to measure the thermal characteristics of the building envelope, either already built or under construction [6], such as the heat flux meter method [7], which consists of calculating the thermal transmittance of walls using two-temperature measurements (inside and outside) and a heat flux meter. However, such a method takes several days, the external surface should not be exposed to solar radiation, and the thermal capacity cannot be estimated. In addition, common uncertainties associated with the heat flow measurement make the identification of high thermal resistances difficult [8]. Teni [9] shows that the methods based on a simple model (where the resistance is expressed by an analytical equation) lead to high uncertainties, except when the boundary conditions are controlled, as in the case of a hot box.
A wide range of other methods are available, mainly using inverse techniques [10,11,12]. They consist of minimizing the gap between a model and the measurements obtained from an experiment in order to infer the unknown parameters of the model. With respect to the opaque walls discussed in this article, inverse techniques can use a wide variety of models such as quadrupole models [13,14], response factor models [15,16], resistance–capacitance models [17,18] or ARX models [7,19,20,21]. Furthermore, inverse techniques can be deterministic approaches [22], Bayesian analyses [23,24,25] or genetic algorithms, for instance [26,27]. These methods can use either temperature measurements (single point measurements with thermocouples or PT100 or mapped measurements with an infrared thermal camera [28]), heat flux measurements or both. The main drawback of these methods is that most of them do not allow the identification of the thermal capacity of the wall. If the method estimates the C-value, heat flux meters are usually required [16,23,25]. Indeed, the method described in this article is designed to be used on highly insulated multilayer walls (R > 4 m2.K.W−1), and heat flux meters do not seem to be suitable for measuring low heat fluxes with sufficient accuracy (<5 W.m²), since in such cases, the signal-to-noise ratio is low. The aim of the method is to avoid the drawback of heat flux meters and addresses an alternative way using appropriate temperature measurements. Moreover, this method is intended to be used in real weather conditions, which requires particular attention to be paid to the measurement of boundary conditions. In contrast to a laboratory, where these conditions are controlled, the heat exchange with the wall is strongly influenced by the wind or the sun, for example.
In the literature, several authors have investigated the influence of boundary conditions on the identification of the thermal characteristics of a building element. Many studies recognize the significant influence of convective heat transfer coefficients on the U-value of walls for infrared thermographic methods (e.g., [29]). For contact-based methods, De Rubeis et al. [30] observed a large deviation of the convective heat transfer coefficient in a guarded hot-box environment compared to default values (−35%). In situ, Evangelisti et al. [31] observed differences in the U-values from 0.5 to 0.9 W.m−2.K−1, depending on the heat transfer coefficient correlations. François et al. [32] also showed the large variations of the internal heat transfer coefficient (between 6 and 12 W.m−2.K−1) under transient conditions. To reduce the influence of uncontrolled boundary conditions on the identification process, a sandwich structure evaluation method was developed by Derbal et al. [33]. An uncertainty below 7% was achieved, but the construction element was limited to single layer materials of PVC, EPS, plaster or concrete. From these different studies, it can be seen that the influence of convective heat transfer is recognized. However, the influence of radiant properties is rarely evaluated for contact-based methods, and these studies are also limited to walls with relatively low thermal resistance (R < 4 m2.K.W−1).
The aim of this paper is therefore to evaluate the performance of inverse techniques for highly insulated multilayer walls under a range of boundary conditions. The calculation of uncertainties on the estimated thermal characteristics is also a key point.
A lightweight, highly insulated wall has been installed in an experimental facility and tests have been carried out. The wall is equipped with temperature sensors on its surfaces and inside. The identification technique (deterministic approach) is applied to a resistance–capacitance model based on temperature measurements, in order to identify the thermal characteristics of walls under real weather conditions.
Estimation of the thermal conductivity and capacity (λ and ρ.c) of the wall materials allows calculation of the R- and C-values, which describe the thermal characteristics of an opaque wall. An inverse technique has been applied to both synthetic (model-based) and measured experimental datasets. The use of synthetic values allows the calculation of uncertainties associated with the boundary conditions.
The first part of this paper presents the case study and the experimental setup. The reference and the resistance–capacitance models are then described and numerically validated. The inverse technique is then introduced. Finally, the effects of two external boundary conditions are discussed in the results section.

2. Experimental Facility and Boundary Conditions

2.1. Case Study

The lightweight wall that was tested was a wall consisting of three homogeneous layers. The first and the third layers were made of moisture-resistant plywood (pl), and the second layer was made of polyurethane (PUR). Five sensors were installed at the locations Sout, S1-2, S2, S2-3 and Sin, as shown in Figure 1.
The layers of the lightweight wall and their material’s properties are described in Table 1, where e is the thickness. Uncertainties are those reported by the manufacturers or found in the literature. If they are missing, ±10% is used as a default value [34].
It is possible to calculate the theoretical R- and C-values of this wall using Equation (1).
R = i e i λ i C = i e i   ρ i   c i
The following theoretical values have been calculated: R = 5.4 ± 0.5 m2.K.W−1 and C = 53 ± 8 kJ.m−2.K−1.
The shortwave absorptivity αSW is a property derived from a measurement performed on a spectrophotometer with an integrating sphere. The value depends on the reference solar spectrum standard used and the sample analysed. The value found was 0.26 ± 0.06.
The longwave absorptivity αLW is equal to the hemispherical emissivity of the surface εLW, taking into account uniform grey and diffuse surfaces. In [35], it is stated that εLW = 0.9 is «usually appropriate for internal and external surfaces» in buildings. An uncertainty of ±0.02 has been defined according to [36].
Table 1. Material properties of the wall.
Table 1. Material properties of the wall.
LayersProperties
MaterialCodeλρce
PlywoodPl0.21 ***710 ± 10 *1400–1800 ***0.021 **
PolyurethanePUR0.023 *32.5 ± 2.5 *1400 ***0.12 **
Sources: * manufacturers and/or official documents. ** measurement or numerical study output. *** technical document ([37,38]).

2.2. Experimental Setup

To test construction elements, outdoor test cells can be used to account for real climatic environments [39]. In this study, it consisted of a highly insulated cell surrounded by a thermal guard that was kept at a constant temperature. One of its faces was exposed to real weather conditions; it was the façade that was tested. A large amount of sensors were set up on this equipment (full weather station, temperature probes in the walls). A heating system was used to simulate different temperature scenarios, allowing the indoor and outdoor conditions to be decoupled during identification.
The experimental data sets come from the facility where the lightweight wall was installed. This facility is shown in Figure 2.
With respect to indoor conditions, the indoor air temperature (Tair,in) was measured by six four-wire resistance thermometers (Pt100 resistance probes) protected by ventilated radiation shields located at a distance of 0.6 m from the wall at three heights (0.8, 1.6, 2.2 m) in order to check the homogeneity of the air temperature. The surface temperature of the wall (9 sensors inside Sin and 9 sensors outside Sout) and the surface temperature of each surface of the test cell (whose average gives TLW,in) were measured by four-wire resistance thermometers located in the center of each wall.
The wall was monitored by four-wire resistance thermometers installed at three different depths (S1-2, S2, S2-3) as shown in Figure 1.
With respect to outdoor conditions, the radiation was measured by two devices installed vertically near the wall: a pyrgeometer that measured impinging longwave radiation integrated within wavelengths ranging between 4 and 40 μm and a pyranometer which measured global shortwave irradiance integrated within wavelengths ranging between 0.3 and 4 μm. The output of the pyranometer is the solar radiation QSW. The outputs of the pyrgeometer are the temperature of the sensor Tpyrgeo and the net radiation Qnet, i.e., the radiation received by the device, called Q→|, minus the radiation emitted by the device (considered as a black body). Using these two values, it is possible to calculate the brightness temperature Tb, which is “the temperature of a black body that would emit the same amount of radiation as the targeted body in a specified spectral band” [40] as presented in Equation (2).
Q n e t t = Q | t σ   T p y r g e o 4 t = σ T b 4 t σ   T p y r g e o 4 t T b t = Q n e t t σ + T p y r g e o 4 t 1 / 4
The outside air temperature Tair,out was measured using a four-wire resistance thermometer protected by a radiation shield.
The wind speed and direction were measured by the sonic anemometer shown in Figure 2c, which was installed 4 m above the roof of the building (i.e., 15 m above the ground). The wind speed close to the façade vwind (i.e., 2.5 m above the ground) was estimated from the wind speed measured by the anemometer using the conversion method presented in [41] and considering a flat terrain with some isolated obstacles.
Details of the equipment used to measure each physical variable are given in Table 2.

2.3. Experimental Boundary Conditions Measurement

The dataset came from a test called the ROLBS case, which stands for “Randomly Ordered Logarithmic Binary Sequence” [42]. It generated heating sequences of variable time duration in order to decorrelate the indoor and outdoor boundary conditions. The ROLBS case lasted for 14 days during the winter season with a measurement time step of 30 s. A moving average was used to obtain a time step of two minutes and thus reduce the amount of data. The measured data are shown in Figure 3. The whole dataset was divided into 3 parts: initialization, identification and validation.
Two typical weather days were selected from the ROLBS case, called the typical sunny day and the typical cloudy day. As shown in Figure 4, during the typical cloudy day, indoor temperature was almost constant. The outdoor temperature was about 10 °C, and the solar radiation varied due to clouds. The westerly wind speed was high (about 13–14 m.s−1). During the typical sunny day, the indoor temperature was also almost constant. The outdoor temperature was between 7 and 20 °C. The solar radiation curve shows a cloudless day. The wind speed from south-east was low (about 2–3 m.s−1). Both cases lasted 24 h.

2.4. Exact Location of Sensors Inside the Wall

The position of sensors located on surfaces (Sout and Sin) or between layers (S1-2 and S2-3) was known. However, the position of the sensor within the insulating layer (S2) was less certain. Sensor S2 (shown in Figure 1) should have been in the middle of the insulating layer but could have been pushed to one side during assembly. In order to verify this hypothesis, experimental data with steady outdoor and indoor conditions were analysed. Measurements from a 24 h experimental test between 27 and 28 November 2018 were used. The difference between the indoor air temperature and the outdoor air temperature was more than 40 °C. In addition, the standard deviations of the air temperature, surface temperature and internal wall temperature measurements were less than 0.8 °C throughout the period. The temperature profile was theoretically linear. The result is shown in Figure 5. The true position of sensor S2 was not exactly in the center of the insulation layer but was displaced inwards by 4.3 mm.

3. Modelling

Only one-dimensional heat conduction in the wall was considered. Heat transfer through multilayer walls was modelled using the heat equation (Equation (3)).
ρ c T t x λ T x = 0
The studied wall consisted of three layers of different materials, namely, layer 1, layer 2 and layer 3 from outside to inside as shown in Figure 6. The outer surface was at x = 0 and the inner surface was at x = L. This wall has a cross-section of 1 m2. Figure 6 shows the five locations (Sout, S1-2, S2, S2-3 and Sin) where sensors were installed. The initial conditions were the temperature values obtained assuming steady state conditions.

3.1. Boundary Conditions

The influence of external weather conditions is taken into account by two types of boundary conditions (BC): either a Neumann condition (heat fluxes reaching the façade, see Equation (4)) is applied at the location of the sensor Sout (called BC1), or a Dirichlet condition is applied at the location of the sensor S1-2 (called BC2) (see Figure 6). Dirichlet condition was not considered for BC1, as the objective is to use Sout to identify the thermal properties and avoid the use of intrusive sensors. Conversely, a Neumann condition was not considered for BC2 as temperature sensors are more robust and accurate than heat flux sensors. In the following sections, only the models built with BC1 are described in detail. A similar procedure was carried out for BC2.
The boundary conditions were conditions of the third kind, also known as Robin’s conditions [43]. They can be written as:
O u t s i d e :       λ l a y e r 1 T x x = 0 = Q c o n v , o u t t + α L W Q L W , o u t t + α S W Q S W t I n s i d e :       λ l a y e r 3 T x x = L = Q c o n v , i n t + Q L W , i n t
where αSW and αLW are the surface absorptivity for shortwave radiations and for longwave radiations.
The indoor longwave radiative heat flux is linearized. The indoor convective and longwave radiative heat flux Qconv,in+ QLW,in is calculated in Equation (5):
Q c o n v , i n t + Q L W , i n t = h i n T i n t T s , i n t
where h i n t = h c o n v , i n t + h L W , i n t   and h L W , i n t 4 σ ε L W T s , i n t + T L W , i n t 2 3 and T i n t = h c o n v , i n t   T a i r , i n t + h L W , i n t   T L W , i n t h i n t .
Where hconv,in is the indoor convective heat transfer coefficient, hLW,in is the indoor radiative heat transfer coefficient, hin is the indoor convective and radiative heat transfer coefficient, Tin is the indoor operative temperature and TLW,in is the mean radiant temperature.
The heat transfer coefficient hLW,in depends on the surface area of the wall, the surface area of the other walls of the test cell, their view factors and their infrared emissivity. The calculation is described in detail in [44]. In this case, the view factor is equal to 1 and the wall surface is about 10 times smaller than the combined surface of the other walls of the test cell. Thus, the approximation given in Equation (5) can be assumed to hold. The difference between the exact formula and the approximation is 1%.
The outdoor heat flux over time is the sum of the outdoor convective heat flux Qconv,out, the absorbed longwave radiative heat flux αLW QLW,out and the absorbed shortwave radiative heat flux αSWQSW. The outdoor longwave radiative heat flux is linearized. The outdoor heat flux is calculated by Equation (6):
Q c o n v , o u t t + α L W Q L W , o u t t + α S W Q S W t = h o u t T o u t t T s , o u t t
where
Q c o n v , o u t t = h c o n v , o u t t T a i r , o u t t T s , o u t t
and
α L W Q L W , o u t t = h L W , o u t t T b t T s , o u t t
With hconv,out standing for the outdoor convective heat transfer coefficient, hLW,out, the outdoor radiative heat transfer coefficient, hout, the outdoor convective and radiative heat transfer coefficient, Tout, the equivalent outdoor temperature (of the viewed environment) and Ts,out, the outer surface temperature of the wall.
The equivalent outdoor temperature can be expressed by Equation (7)
T o u t t = h c o n v , out t T a i r , out t + h L W , out t T b t + α S W Q S W t h o u t t
where
h o u t t = h c o n v , out t + h L W , out t   and   h L W , out = 4 σ ε L W T b t + T s , o u t t 2 3

3.2. Heat Transfer Coefficients

The models described above are based on the following heat transfer coefficients: hLW,in, hLW,out, hconv,in, hconv,out. hLW,in and hLW,out, which are calculated from Equations (5) and (7). They depend on the surface temperatures (Ts,in or Ts,out), the temperature of their surroundings (TLW,in or Tb) and the hemispherical infrared emissivity of the surface εLW. The uncertainty is estimated to be about ±0.15 W.m−2.K−1 using the propagation of uncertainties.
For convection, hconv,in is assumed to be equal to 2.5 W.m−2.K−1 as defined in [35] for a horizontal heat flux. The same value is given by the correlation of Khalifa and Marshall ([45,46]) considering a room heated by a circulating fan (wall opposite to the fan). As mentioned in [46,47], the uncertainty is about ±20%.
Convective heat transfer on external surfaces is influenced by many variables, such as local wind speed, flow regime and surface roughness. Many empirical correlations exist to estimate hconv,out as a function of wind speed vwind and wind direction dwind [48]. In this article, a simple correlation is chosen according to ISO 6946 [35], as shown in Equation (8). In our case study, it leads to a variation of hconv,out between 5 and 45 W.m−2.K−1.
h c o n v , o u t t = 4 + 4   v w i n d t
In order to estimate the uncertainty of this coefficient, two other correlations were calculated: that of the ASHRAE task group which proposes a non-linear relationship to the local wind speed [49,50] and that of Liu and Harris, which makes a distinction according to the wind incidence (windward-leeward) [51]. These two models were selected, as they are quite widely used and also cover the full range of correlations found in the literature [5]. As shown in Figure 7, the correlation of Liu and Harris gives lower heat transfer coefficients, and the ASHRAE correlation is in line with the simple correlation or slightly higher. A variation of ±7 W.m².K−1 is thus considered for this parameter.

3.3. The Reference Model

The reference model to which the RC model was compared uses COMSOL® 5.5 software [52]. This software is based on the finite element method. The wall is divided into 28 domains. Quadratic interpolation is used for the temperature field. The time step is set at 10 s. The boundary conditions are modelled as described above (Equations (5) and (6)), except for the absorbed longwave radiative heat flux, which is not linearized.

3.4. The Resistance–Capacitance Model

In the resistance–capacitance (RC) model, the walls are considered to be a combination of resistances and capacitances. The nodes between two layers or at the surface (indoor or outdoor) have no capacity [53]. Resistances and capacitances are evenly distributed in each layer.
Considering a simple RC model with one capacity in each layer (Figure 8), also known as the 6R3C model, the following equations are obtained:
d i a g 0 , C 1 , 0 ,   C 2 ,   0 ,   C 3 ,   0 × T s , o u t ˙ T 1 ˙ T 1 2 ˙ T 2 ˙ T 2 3 ˙ T 3 ˙ T s , i n ˙   = h o u t 0 0     0 0 h i n × T o u t T i n + h o u t 2 R 1 2 R 1 0 0 2 R 1 4 R 1 2 R 1 0 2 R 1 2 R 1 2 R 2 2 R 2 2 R 2 4 R 2 2 R 2 2 R 2 2 R 2 2 R 3 2 R 3 0 2 R 3 4 R 3 2 R 3 0 0 2 R 3 h i n 2 R 3 × T s , o u t T 1 T 1 2 T 2 T 2 3 T 3 T s , i n
The system of Equations (9) has been solved using a backward Euler method. However, the 6R3C model is not sufficiently accurate. The number of nodes in each layer is chosen to make the RC model sufficiently accurate. The first and third layers (plywood) are divided into five nodes, and the second layer (PUR) is divided into eleven nodes. The model is called the 24R21C model.

3.5. Model Validation

The model was validated for cases encountered in experimental tests. These were divided into two categories: cloudy days and sunny days, as shown in Figure 4. The condition to validate the model is its ability to accurately reproduce the thermal behaviour of the wall on both typical days. To do this, the results obtained by simulating the 24R21C model were compared with those obtained using the reference model for two types of boundary conditions.
The comparison of the model results was based on the study of the mean absolute deviations (MAD), i.e., the mean of the absolute value of the difference between the reference model and the 24R21C model. The MAD for the cloudy day and for the sunny day are presented in Figure 9. This comparison was made for boundary conditions BC1. The results are similar for boundary conditions BC2 (sensors S2, S2-3 and Sin only).
The 24R21C model has a good accuracy for all sensors and both weather conditions. The largest error is for Sout.

4. Identification Technique

4.1. Principle of the Inverse Method

The principle of the inverse method is illustrated in Figure 10. On the one hand, experimental reference data (synthetic or measured) were available ( T s e n s o r ) , and on the other hand, temperatures calculated by the 24R21C model defined above ( T m o d e l ) were generated.
The aim was to find the values of the physical parameters β defined in the 24R21C model, so that the calculated temperatures fitted as close as possible to the experimental values. To achieve this, the inverse method consisted of minimizing the objective function defined in Figure 10. This function was minimized using the Nelder–Mead algorithm [55]. Different initial values were tested to avoid local minima.
The aim was to estimate the physical parameters (λ and ρ.c) of each layer and then calculate the R- and C-values of the walls using Equation (1).

4.2. Calculation of Uncertainties

Two sources of uncertainty were assessed: the uncertainty of the inverse method and model and the uncertainty linked to the boundary conditions. To assess the uncertainty due to the inverse method and model, an identification was performed using the “synthetic data without noise” (synthetized with the measured boundary conditions, as shown in Figure 3). As the theoretical values of R and C are known, the difference between the estimated values and the theoretical values provided the uncertainty of the inverse method.
In order to assess the uncertainty due to the boundary conditions, the identification was carried out with the “synthetic data with noise”; i.e., it included the measurement uncertainties. A white noise was applied to all the boundary conditions (Tout, Tin, Tb, QSW) and to the heat transfer coefficients (hconv,out, hin), given the associated uncertainties mentioned above. εSW and εLW were randomly generated, considering their respective uncertainties. More than 50 tests were performed with a randomly selected noise, with boundary conditions BC1 and BC2.
The total uncertainty is equal to the uncertainty of the inverse method and model plus the uncertainty due to the boundary condition measurements. The uncertainty of the inverse method and model is the difference between the theoretical values and the estimated values in the case of the data “without noise”. The uncertainty due to boundary condition measurements is defined as two times the standard deviation of the data “with noise”, so that 95% of the estimates was included.

5. Results

First, the inverse method described in Section 4.1 was applied to synthetic data with boundary conditions BC1 (see location in Figure 6). The effect of measurement uncertainty was investigated. Then, the method was applied to synthetic data considering boundary conditions BC2 (see location in Figure 6). Finally, the method was applied to measured experimental data.

5.1. Application to Synthetic Data

5.1.1. Influence of the Heat Transfer Coefficients and Absorptivity

As shown in [31,46], heat transfer coefficients vary depending on the chosen correlation. The accuracy of the coefficients is also affected by the measurement uncertainty. It is therefore important to ensure that the uncertainty in the transfer coefficients does not have too large an effect on the output of the models, i.e., the calculated temperatures.
Given the uncertainties defined in Section 2 and Section 3, it is possible to calculate the influence of the variation of each heat transfer coefficient during the typical sunny day and the typical cloudy day on the outputs of the RC model, as shown in Figure 11. In this case, results are obtained with boundary conditions. This influence is measured by calculating the difference between the temperatures calculated by the model when the parameter is equal to its upper limit and to its lower limit. The result for the 24R21C model is shown in Figure 11.
Figure 11 shows that Sout and S1-2 are very sensitive to αSW and hconv,out variations during sunny and cloudy days. The results from sensors Sout and S1-2 were rejected because αSW and hconv,out were not known with high accuracy. Uncertainty may have had a negative impact on the identification. S2 also appears to be sensitive on sunny days.

5.1.2. Sensitivity According to Parameters to Be Identified

With boundary conditions BC1, a sensitivity analysis was carried out to investigate the variation in the response of the sensors to variations in the unknown physical parameters λ and ρ.c. This analysis provides information on the feasibility of estimating all the parameters simultaneously. The reduced sensitivities S* were calculated according to Equation (10) [56]:
S i * t , β = β i T k β i t , β = β i ¯ T k t , β ,   β i + T k t , β ,   β i β i + β i
where β i is the parameter of interest and β i ¯ , β i + and β i are the mean value, upper and lower limits (±10 %), respectively.
The variations of λpl, λPUR, (ρ.c)pl and (ρ.c)PUR were evaluated. The reduced sensitivities versus time are plotted in Figure 12. for the 24R21C model during the cloudy and the sunny days described in Figure 4.
It seems clear that the reduced sensitivities of the four thermal properties are high for sensor S2. The temperature measured at the location of sensor S2-3 is only sensitive to the parameter λPUR. Sunny days seem to be the most convenient for the identification of the parameters, since the wall is more thermally stressed due to the solar radiation. The identification dataset should therefore contain sunny days to identify both R- and C-values.
The sensor Sin cannot be considered as a relevant sensor location, since the sensitivity to the parameters is rather low. Moreover, the measurement of surface temperatures is challenging and has a higher experimental uncertainty than the temperature inside walls. The temperature difference between the wall surface and the air is so small that in most cases it is close to the measurement uncertainty (±0.15 °C). Therefore, only sensors S2 and S2-3 are selected for identification. Figure 12 shows that some parameters may be correlated. This assumption is confirmed by looking at the correlation matrix of the two sensors (using a column-vector of sensitivities [54]) and during the ROLBS case shown in Table 3. All parameters cannot be estimated simultaneously, as (ρ.c)pl and (ρ.c)PUR are strongly correlated.
However, according to theoretical calculations, 96% of the resistance is due to the insulation level of the polyurethane layer (λPUR) and 90% of the capacity is due to the capacity of the plywood layers (ρ.c)pl. The identification will therefore focus on λPUR and (ρ.c)pl. While λpl and (ρ.c)PUR are assumed to be equal to the theoretical value defined in Table 1, as an error in these values would have a limited effect on the estimated R- and C-values.

5.1.3. Identification Results with Synthetic Data and Boundary Conditions BC1

In this part, the Tsensor values in Figure 10 are synthetic data generated by the reference model. Boundary conditions during this test are boundary conditions BC1 (see Section 2.3). Identification was applied to sensors S2 and S2-3. As shown in the previous section, Sout and S1-2 were not used for identification due to the high influence of ill-defined model parameters, and Sin was rejected due to its low sensitivity. The R- and C-values were calculated simultaneously by estimating λPUR and (ρ.c)pl.
The accuracy of the identification technique applied to the 24R21C model was assessed by comparing the differences between the identified values and R and C (see Section 2.1). Figure 13 shows the result of the identification of the thermal characteristics of the wall with and without noise according to the 24R21C model. The results are given as the difference between the estimated value and the theoretical target value. The graph has box plots to be read statistically. The crosses represent the mean value. The box extends from the first to the third quartile. The line in the middle indicates the median. The whiskers are a function of the inter-quartile range and the extreme values. Each black dot represents an estimate.
In terms of identification applied to synthetic data without noise, the C- and R-values are accurately estimated. The difference in estimation of the C-value is +0.75 kJ.m−2.K−1 (+1%) compared to the theoretical value. The R-value is estimated to be +0.12 m2.K.W−1 (+2%) higher than the theoretical value.
When noise is applied, the results are scattered. Half of the estimated resistances are between −4% and +1% of the theoretical value, but more extreme values occur. The error can reach up to −0.9 m2.K.W−1 or −16%. Regarding capacities, the results are also scattered. Half of the estimated capacities are between −5% and +27% of the theoretical value. The error can reach up to 79%.
The total uncertainty can be calculated as shown in Section 4.2. The uncertainties are ±0.6 m2.K.W−1 (10% error) for the R-value and ±26 kJ.m−2.K−1 (48% error) for the C-value.
This indicates that identification from the measurement of the outer boundary conditions is not possible, because the uncertainties in the measurement of the boundary conditions and the heat transfer coefficients are too large in this case study. In the following section, a method is proposed to overcome these uncertainties.

5.1.4. Application to Synthetic Data with Boundary Conditions BC2

A similar methodology to that presented in Section 5.1.3 was used to assess the robustness of the identification with boundary conditions BC2 (shown in Figure 6). As the outer boundary conditions is placed at the interface between the plywood and the polyurethane, the sensor S2 is no longer sensitive to the thermal capacity of the plywood (Figure 14). However, sensor S2-3 is sensitive enough to the thermal capacity when active heating is applied with a relatively rapid change in the indoor temperature (see the dashed line in Figure 14). This sensitivity is small but greater than the uncertainty of temperature sensors (over 0.15 °C). The limited influence of the indoor heat transfer coefficient was also observed (as in Figure 11). Therefore, the identification procedure was applied to sensors S2 and S2-3. The R- and C-values were calculated simultaneously by estimating λPUR and (ρ.c)pl.
The accuracy of the identification technique was assessed by comparing the differences between the estimated values of R and C (see Section 2.1). Figure 15 shows the result of the identification of the thermal characteristics of the wall with and without noise using boundary conditions BC2.
When the identification is applied to synthetic data without noise, the R-value is slightly underestimated by 2% and the C-value is also slightly underestimated by 1%. When noise is applied, the difference between the identified resistance and the theoretical value is about −0.15 m2.K.W−1 or −2.5% for all estimates. For the capacities, the difference is also rather small (always less than 1 kJ.m−2.K−1 or 1% error).
The total uncertainty can be calculated as shown in Section 4.2. The uncertainties for 24R21C model are ±0.15 m2.K.W−1 (3% error) for R-values and ±1.5 kJ.m−2.K−1 (3% error) for C-values. The use of boundary conditions BC2 greatly improves the accuracy of the R- and C-value estimation.

5.2. Application to Measured Experimental Data

In order to validate the previous observations, the identification methods were tested on measured experimental data. The wall properties were estimated from the measured experimental data. The identification results are shown in Figure 16. The grey area represents the range of theoretical values with the uncertainties defined in Section 2.1. The light pink rectangle represents the range of estimated R- and C-values, with the boundary condition BC1, and the deep pink rectangle represents the range of estimated R- and C-values, with the boundary condition BC2.
It can be seen that the results with BC1 encompass extended ranges, as the uncertainties are large in this case. Moreover, the R-value is significantly overestimated.
The area corresponding to boundary conditions BC2 is smaller and closer to the theoretical area. The resistance can be estimated between 5.05 and 5.35 m2.K.W−1. The estimated capacities are slightly lower than the theoretical capacity, between 41.5 and 44.5 kJ.m−2.K−1. The difference between theory and estimate may be due to the uncertainty of the model, the duration of the wall loading, the reduced sensitivity to the plywood thermal capacity or the theoretical values estimated from tabulated data.
To verify that the RC model correctly reproduces the thermal behaviour of the wall, the validation part of the ROLBS case was used (see data in Figure 3). Figure 17 shows the temperatures measured by sensors S2 and S2-3 and the temperature estimated by the model and their residuals. The residuals are small, indicating that the model correctly reproduces the thermal behaviour of the wall.

6. Discussions

Table 4 summarizes the results. The main differences between the identification results lie in the choice of boundary conditions. It is shown that when boundary conditions BC1 are chosen, the uncertainties are so large that the results are unusable. When boundary conditions BC2 are considered, the RC model estimates the resistance and capacity with acceptable uncertainties.
A procedure is provided as shown in Figure 18. Two cases are highlighted regarding the boundary conditions BC1 or BC2. If BC2 is chosen, special attention must be paid to the location of the sensors in order to minimize their sensitivity to the boundary conditions, while setting the experimental design. Once the sensors have been installed and their uncertainty calculated, the parameters to be estimated can be derived from a second sensitivity study. The uncertainty of identification associated with these boundary conditions can then be calculated. If the accuracy is satisfactory, the experiment is performed with dynamic boundary conditions.
If BC1 is chosen, attention must be paid to the accuracy of the sensors measuring the boundary conditions, as this has been shown in this article to be a limitation of this method. The calculation of the uncertainties of the parameter estimation allows the method to be validated or not.

7. Conclusions

The objective of this article has been to evaluate the one-dimensional heat transfer in a multilayer wall under real weather conditions using temperature sensors.
A sensitivity analysis was performed to reject the sensors that were too sensitive to parameters defined with low accuracy, such as the internal and external radiative and convective heat transfer coefficients and the external solar absorptivity. Differences up to 8 °C could be observed, depending on the assumptions. This step determined that two of the five sensors were useful.
To increase the confidence in estimated parameters, the identification was limited to the parameters contributing the most to the global thermal properties of the wall assembly. Therefore, it was found that estimating the thermal conductivity of the insulation layer and the specific heat capacity of the wooden layer was enough for assessing accurately the R- and C-values of the wall.
The parameter estimation method was carried out with two types of boundary conditions: either the outside boundary conditions (air temperature, brightness temperature and solar radiation) or the temperature measured a few centimeters below the outer surface of the wall (intrusive sensor). The uncertainty on the estimated parameters obtained with the outside boundary conditions was very large (over 1 m2.K.W−1 for R) compared to the one obtained by using an intrusive sensor (0.3 m2.K.W−1). Using an intrusive sensor, with adapted sensitivity to the applied thermal loads, considerably reduced the uncertainty of the estimation. Considering the results, the estimated resistance fell within the range of variation defined from the theoretical values. The estimated capacity was slightly lower than the theoretical value.
Finally, this article provided guidelines for measuring thermal properties of homogenous, highly insulated multilayer walls with uncertainties. It requires a rough knowledge of the wall structure. A procedure is presented in Figure 18. It helps to confirm the relevance of carrying out experimental tests according to the location of sensors and the expected uncertainty of the measurements. This methodology could also be applied to highly insulated heavyweight walls to see if the same conclusions can be drawn about the ability of the model to estimate thermal characteristics.

Author Contributions

Conceptualization, M.R., J.L.D., P.S. and M.D.; methodology, M.R., J.L.D., P.S. and M.D.; software, M.R.; validation, M.R., J.L.D., P.S. and M.D.; formal analysis, M.R.; data curation, M.R.; writing—original draft preparation, M.R.; writing—review and editing, M.R., J.L.D., P.S. and M.D.; supervision, J.L.D. and P.S.; project administration, funding acquisition, P.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Union and the French region Nouvelle-Aquitaine with the industrial partners TIPEE Platform and EDF (Research Project “CITEE—Innovative components for building envelopes”). The authors are grateful for their technical and financial support.

Data Availability Statement

The code and data can be found at: https://gitlab.univ-lr.fr/jledreau/Inverse_Wall1D (accessed on 14 November 2024).

Acknowledgments

A special thanks goes to Michel Burlot, David Gaillard and Laurent Servant for their involvement in the experimental work and to Denis Maillet and Mathilde Colmet-Daage for their relevant advice.

Conflicts of Interest

There are no conflicts of interest.

Nomenclature

cSpecific heat capacity, J.kg−1.K−1QHeat flux density, W.m−2
CTotal thermal mass, J.m−2.K−1RThermal resistance, m2.K.W−1
dWind direction, °SSensor
eThickness of a layer, mtTime, s
hHeat transfer coefficient, W.m−2.K−1TTemperature, °C or K
LThickness of wall, mvVelocity, m.s−1
MADMean absolute deviation, °CxSpace variable, m
Greek symbols
αAbsorptivity, -λThermal conductivity, W.m−1.K−1
βList of parameters to be estimatedρDensity, kg.m−3
εEmissivity, -σStefan-Boltzmann constant, W.m−2.K−4
Index and Exponent
→|Received by devicePURPolyurethane
bBrightnesspyrgeoPyrgeometer
convConvectionoutOutdoor
inIndoorSSurface
LWLongwave radiationSWShortwave radiation
plPlywood

References

  1. Meiss, A.; Feijó-Muñoz, J. The energy impact of infiltration: A study on buildings located in north central Spain. Energy Effic. 2015, 8, 51–64. [Google Scholar] [CrossRef]
  2. Eleftheriadis, G.; Hamdy, M. Impact of building envelope and mechanical component degradation on the whole building performance: A review paper. Energy Procedia 2017, 132, 321–326. [Google Scholar] [CrossRef]
  3. Asan, H.; Sancaktar, Y. Effects of Wall’s thermophysical properties on time lag and decrement factor. Energy Build. 1998, 28, 159–166. [Google Scholar] [CrossRef]
  4. NF EN 12939; Thermal Performance of Building Materials and products—Determination of Thermal Resistance by Means of Guarded Hot Plate and Heat Flow Meter Methods—Thick Products of High and Medium Thermal Resistance. British Standards Institution: London, UK, 2000.
  5. NF EN 12667; Thermal Performance of Building Materials and Products—Determination of Thermal Resistance by Means of Guarded Hot Plate and Heat Flow Meter Methods—Products of High and Medium Thermal Resistance. British Standards Institution: London, UK, 2001.
  6. Bienvenido-Huertas, D.; Moyano, J.; Marín, D.; Fresco-Contreras, R. Review of in situ methods for assessing the thermal transmittance of walls. Renew. Sustain. Energy Rev. 2019, 102, 356–371. [Google Scholar] [CrossRef]
  7. ISO 9869-1; Thermal Insulation—Building Elements—In-Situ Measurement of Thermal Resistance and Thermal Transmittance—Part 1: Heat Flow Meter Method. International Organization for Standardization: Geneva, Switzerland, 2014.
  8. Peng, C.; Wu, Z. In situ measuring and evaluating the thermal resistance of building construction. Energy Build. 2008, 40, 2076–2082. [Google Scholar] [CrossRef]
  9. Teni, M.; Krstić, H.; Kosiński, P. Review and comparison of current experimental approaches for in-situ measurements of building walls thermal transmittance. Energy Build. 2019, 203, 109417. [Google Scholar] [CrossRef]
  10. Beck, J.V.; Blackwell, B.; Clair, C.R.S. Inverse Heat Conduction: Ill-Posed Problems; Wiley: Hoboken, NJ, USA, 1985. [Google Scholar]
  11. Beck, J.V.; Arnold, K.J. Parameter Estimation in Engineering and Science; Wiley: Hoboken, NJ, USA, 1977. [Google Scholar]
  12. Orlande, H.R.; Fudym, O.; Maillet, D.; Cotta, R.M. Thermal Measurements and Inverse Techniques; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar] [CrossRef]
  13. Defer, D.; Shen, J.; Lassue, S.; Duthoit, B. Non-destructive testing of a building wall by studying natural thermal signals. Energy Build. 2002, 34, 63–69. [Google Scholar] [CrossRef]
  14. Sassine, E. A practical method for in-situ thermal characterization of walls. Case Stud. Therm. Eng. 2016, 8, 84–93. [Google Scholar] [CrossRef]
  15. Rasooli, A.; Itard, L.; Ferreira, C.I. A response factor-based method for the rapid in-situ determination of wall’s thermal resistance in existing buildings. Energy Build. 2016, 119, 51–61. [Google Scholar] [CrossRef]
  16. Rasooli, A.; Itard, L. In-situ rapid determination of walls’ thermal conductivity, volumetric heat capacity, and thermal resistance, using response factors. Appl. Energy 2019, 253, 113539. [Google Scholar] [CrossRef]
  17. Gutschker, O. Parameter identification with the software package LORD. Build. Environ. 2008, 43, 163–169. [Google Scholar] [CrossRef]
  18. Baker, P.; van Dijk, H. PASLINK and dynamic outdoor testing of building components. Build. Environ. 2008, 43, 143–151. [Google Scholar] [CrossRef]
  19. Loussouarn, T.; Maillet, D.; Remy, B.; Schick, V.; Dan, D. Indirect measurement of temperature inside a furnace, ARX model identification. J. Phys. Conf. Ser. 2018, 1047, 012006. [Google Scholar] [CrossRef]
  20. Jiménez, M.; Heras, M. Application of multi-output ARX models for estimation of the U and g values of building components in outdoor testing. Sol. Energy 2005, 79, 302–310. [Google Scholar] [CrossRef]
  21. Naveros, I.; Ghiaus, C.; Ruíz, D.; Castaño, S. Physical parameters identification of walls using ARX models obtained by deduction. Energy Build. 2015, 108, 317–329. [Google Scholar] [CrossRef]
  22. Limam, K.; Bouache, T.; Ginestet, S.; Popescu, L.O. Numerical and experimental identification of simplified building walls using the reflective Newton method. J. Build. Phys. 2018, 41, 321–338. [Google Scholar] [CrossRef]
  23. Naveros, I.; Bacher, P.; Ruiz, D.; Jiménez, M.; Madsen, H. Setting up and validating a complex model for a simple homogeneous wall. Energy Build. 2014, 70, 303–317. [Google Scholar] [CrossRef]
  24. Berger, J.; Orlande, H.R.; Mendes, N.; Guernouti, S. Bayesian inference for estimating thermal properties of a historic building wall. Build. Environ. 2016, 106, 327–339. [Google Scholar] [CrossRef]
  25. Biddulph, P.; Gori, V.; Elwell, C.A.; Scott, C.; Rye, C.; Lowe, R.; Oreszczyn, T. Inferring the thermal resistance and effective thermal mass of a wall using frequent temperature and heat flux measurements. Energy Build. 2014, 78, 10–16. [Google Scholar] [CrossRef]
  26. Rouchier, S.; Woloszyn, M.; Kedowide, Y.; Béjat, T. Identification of the hygrothermal properties of a building envelope material by the covariance matrix adaptation evolution strategy. J. Build. Perform. Simul. 2015, 9, 101–114. [Google Scholar] [CrossRef]
  27. Wang, S.; Xu, X. Parameter estimation of internal thermal mass of building dynamic models using genetic algorithm. Energy Convers. Manag. 2006, 47, 1927–1941. [Google Scholar] [CrossRef]
  28. Chaffar, K.; Chauchois, A.; Defer, D.; Zalewski, L. Thermal characterization of homogeneous walls using inverse method. Energy Build. 2014, 78, 248–255. [Google Scholar] [CrossRef]
  29. Dall‘O’, G.; Sarto, L.; Panza, A. Infrared Screening of Residential Buildings for Energy Audit Purposes: Results of a Field Test. Energies 2013, 6, 3859–3878. [Google Scholar] [CrossRef]
  30. de Rubeis, T.; Evangelisti, L.; Guattari, C.; De Berardinis, P.; Asdrubali, F.; Ambrosini, D. On the Influence of Environmental Boundary Conditions on Surface Thermal Resistance of Walls: Experimental Evaluation through a Guarded Hot Box. Case Stud. Therm. Eng. 2022, 34, 101915. [Google Scholar] [CrossRef]
  31. Evangelisti, L.; Guattari, C.; Gori, P.; Vollaro, R.d.L.; Asdrubali, F. Experimental Investigation of the Influence of Convective and Radiative Heat Transfers on Thermal Transmittance Measurements. Int. Commun. Heat Mass Transf. 2016, 78, 214–223. [Google Scholar] [CrossRef]
  32. François, A.; Ibos, L.; Feuillet, V.; Meulemans, J. Novel in Situ Measurement Methods of the Total Heat Transfer Coefficient on Building Walls. Energy Build. 2020, 219, 110004. [Google Scholar] [CrossRef]
  33. Derbal, R.; Defer, D.; Chauchois, A.; Antczak, E. A simple method for building materials thermophysical properties estimation. Constr. Build. Mater. 2014, 63, 197–205. [Google Scholar] [CrossRef]
  34. 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]
  35. EN ISO 6946; Building Components and Building Elements—Thermal Resistance and Thermal Transmittance—Calculation Method. International Organization for Standardization: Geneva, Switzerland, 2017.
  36. Macdonald, I.A. Quantifying the Effects of Uncertainty in Building Simulation. Ph.D. Thesis, University of Strathclyde, Glasgow, UK, 2002. [Google Scholar]
  37. CSTB. Règles Th-bat—Fascicule Matériaux (French Building Regulation). France, 2017. Available online: https://rt-re-batiment.developpement-durable.gouv.fr/IMG/pdf/2-fascicule_materiaux.pdf (accessed on 14 November 2024).
  38. Ross, R.J. Wood Handbook: Wood as an Engineering Material; U.S. Dept. of Agriculture, Forest Service, Forest Products Laboratory: Madison, WI, USA, 2010; Volume 190. [Google Scholar]
  39. Cattarin, G.; Causone, F.; Kindinis, A.; Pagliano, L. Outdoor test cells for building envelope experimental characterization—A literature review. Renew. Sustain. Energy Rev. 2016, 54, 606–625. [Google Scholar] [CrossRef]
  40. Spampinato, L.; Calvari, S.; Oppenheimer, C.; Boschi, E. Volcano surveillance using infrared cameras. Earth-Sci. Rev. 2011, 106, 63–91. [Google Scholar] [CrossRef]
  41. Orme, M. Applicable Models for Air Infiltration and Ventilation Calculations; AIVC: Vancouver, BC, Canada, 1999. [Google Scholar]
  42. Madsen, H.; Bacher, P.; Bauwens, G.; Deconinck, A.H.; Reynders, G.; Roels, S.; Himpe, E.; Lethé, G. Thermal Performance Characterization using Time Series Data. In IEA EBC Annex 58 Guidelines; International Energy Agency: Paris, France, 2015. [Google Scholar]
  43. Maillet, D.; André, S.; Batsale, J.-C.; Degiovanni, A.; Moyne, C. Thermal Quadrupoles: Solving the Heat Equation Through Integral Transforms; Wiley: Hoboken, NJ, USA, 2000. [Google Scholar]
  44. Brun, A. Amélioration du confort d’été dans des bâtiments à ossature par ventilation de l’enveloppe et stockage thermique. Ph.D. Thesis, Université de Grenoble, Saint-Martin-d’Hères, France, 2011. [Google Scholar]
  45. Khalifa, A.; Marshall, R. Validation of heat transfer coefficients on interior building surfaces using a real-sized indoor test cell. Int. J. Heat Mass Transf. 1990, 33, 2219–2236. [Google Scholar] [CrossRef]
  46. Peeters, L.; Beausoleil-Morrison, I.; Novoselac, A. Internal convective heat transfer modeling: Critical review and discussion of experimentally derived correlations. Energy Build. 2011, 43, 2227–2239. [Google Scholar] [CrossRef]
  47. Awbi, H.; Hatton, A. Natural convection from heated room surfaces. Energy Build. 1999, 30, 233–244. [Google Scholar] [CrossRef]
  48. Defraeye, T.; Blocken, B.; Carmeliet, J. Convective heat transfer coefficients for exterior building surfaces: Existing correlations and CFD modelling. Energy Convers. Manag. 2011, 52, 512–522. [Google Scholar] [CrossRef]
  49. ASHRAE Task Group. Procedure for Determining Heating and Cooling Loads for Computerizing Energy Calculations: Algorithms for Building Heat Transfer Subroutines; American Society of Heating, Refrigerating and Air-Conditioning Engineers: Peachtree Corners, GA, USA, 1976. [Google Scholar]
  50. Sharples, S. Full-scale measurements of convective energy losses from exterior building surfaces. Build. Environ. 1984, 19, 31–39. [Google Scholar] [CrossRef]
  51. Liu, Y.; Harris, D. Full-scale measurements of convective coefficient on external surface of a low-rise building in sheltered conditions. Build. Environ. 2007, 42, 2718–2736. [Google Scholar] [CrossRef]
  52. COMSOL. COMSOL Multiphysics 5.5 Reference Manual. 2019. Available online: https://doc.comsol.com/5.5/doc/com.comsol.help.comsol/COMSOL_ReferenceManual.pdf (accessed on 14 November 2024).
  53. Tsilingiris, P. Parametric space distribution effects of wall heat capacity and thermal resistance on the dynamic thermal behavior of walls and structures. Energy Build. 2006, 38, 1200–1211. [Google Scholar] [CrossRef]
  54. Petit, D.; Maillet, D. Techniques inverses et estimation de paramètres. Partie 1. Tech. Ing. Phys. Stat. Mathématique 2008, 42619210, 1–18. [Google Scholar]
  55. Nelder, J.A.; Mead, R. A Simplex Method for Function Minimization. Comput. J. 1965, 7, 308–313. [Google Scholar] [CrossRef]
  56. Remy, B.; Degiovanni, A. Parameters estimation and measurement of thermophysical properties of liquids. Int. J. Heat Mass Transf. 2005, 48, 4103–4120. [Google Scholar] [CrossRef]
Figure 1. Locations of sensors in the wall.
Figure 1. Locations of sensors in the wall.
Buildings 14 03706 g001
Figure 2. Photos of the experimental setup: (a) outdoor air temperature sensor; (b) radiometers; (c) sonic anemometer; (d) outdoor surface temperature sensors; (e) indoor air temperature sensors with ventilated shield and indoor surface temperature sensors.
Figure 2. Photos of the experimental setup: (a) outdoor air temperature sensor; (b) radiometers; (c) sonic anemometer; (d) outdoor surface temperature sensors; (e) indoor air temperature sensors with ventilated shield and indoor surface temperature sensors.
Buildings 14 03706 g002
Figure 3. Boundary conditions for the ROLBS case: temperature and solar radiation (a), the related wind rose (b) and wind speed and direction against time (c).
Figure 3. Boundary conditions for the ROLBS case: temperature and solar radiation (a), the related wind rose (b) and wind speed and direction against time (c).
Buildings 14 03706 g003aBuildings 14 03706 g003b
Figure 4. Boundary conditions for the typical cloudy day and the typical sunny day.
Figure 4. Boundary conditions for the typical cloudy day and the typical sunny day.
Buildings 14 03706 g004
Figure 5. Estimation of location of sensor S2 based on the temperature distribution under steady-state conditions.
Figure 5. Estimation of location of sensor S2 based on the temperature distribution under steady-state conditions.
Buildings 14 03706 g005
Figure 6. Location of boundary conditions and sensors.
Figure 6. Location of boundary conditions and sensors.
Buildings 14 03706 g006
Figure 7. Outdoor convective heat transfer coefficients for different correlations in the ROLBS case.
Figure 7. Outdoor convective heat transfer coefficients for different correlations in the ROLBS case.
Buildings 14 03706 g007
Figure 8. RC model structure for the 6R3C model.
Figure 8. RC model structure for the 6R3C model.
Buildings 14 03706 g008
Figure 9. Validation of the 24R21C model.
Figure 9. Validation of the 24R21C model.
Buildings 14 03706 g009
Figure 10. Principle of inverse method, adapted from [54].
Figure 10. Principle of inverse method, adapted from [54].
Buildings 14 03706 g010
Figure 11. Study of the impact of the variation of heat transfer coefficients during the cloudy day and the sunny day.
Figure 11. Study of the impact of the variation of heat transfer coefficients during the cloudy day and the sunny day.
Buildings 14 03706 g011
Figure 12. Sensitivity analysis of the parameters estimated during both typical days (BC1).
Figure 12. Sensitivity analysis of the parameters estimated during both typical days (BC1).
Buildings 14 03706 g012
Figure 13. Estimation of R- and C-values—data with and without noise—BC1.
Figure 13. Estimation of R- and C-values—data with and without noise—BC1.
Buildings 14 03706 g013
Figure 14. Sensitivity analysis of the parameters to be estimated with free floating conditions (sunny day) and with active heating (1 March 2019) with BC2.
Figure 14. Sensitivity analysis of the parameters to be estimated with free floating conditions (sunny day) and with active heating (1 March 2019) with BC2.
Buildings 14 03706 g014
Figure 15. Estimation of R- and C-values—data with and without noise—BC2.
Figure 15. Estimation of R- and C-values—data with and without noise—BC2.
Buildings 14 03706 g015
Figure 16. Estimation of R- and C-values from measured experimental data and using outside and inside boundary conditions.
Figure 16. Estimation of R- and C-values from measured experimental data and using outside and inside boundary conditions.
Buildings 14 03706 g016
Figure 17. Temperature measured in the wall and estimated by the model and the residuals.
Figure 17. Temperature measured in the wall and estimated by the model and the residuals.
Buildings 14 03706 g017
Figure 18. Procedure for experimental design.
Figure 18. Procedure for experimental design.
Buildings 14 03706 g018
Table 2. Details on devices used for measurements.
Table 2. Details on devices used for measurements.
Physical VariableSensorAcquisitionAccuracy
Indoor air temperaturePT100 1/5 B DIN, 4 wires (TCSA)KEYSIGHT 34980A±0.15 °C *
Surface temperaturePT100 1/3 B DIN, 4 wires (TCSA)
Interface temperature
Outdoor air temperaturePT100 (VAISALA HMP1550)COMPACT DAQ NI 9188
Shortwave radiationSMP21 (KIPP AND ZONEN)RS 485—Modbus±3 W.m−2 **
Brightness temperatureSGR4 (KIPP AND ZONEN)±0.7 °C **
Wind velocityWINCAP Ultrasonic WMT701 (VAISALA)±0.1 m.s−1 **
Wind direction±2° **
* Estimated from calibration and technical documents. ** From technical documents.
Table 3. Correlation matrix.
Table 3. Correlation matrix.
λplλPUR(ρ.c)pl(ρ.c)PUR
λpl1−0.240.24−0.03
λPUR 1−0.050.16
(ρ.c)pl 1−0.95
(ρ.c)PUR 1
Table 4. Results of estimation.
Table 4. Results of estimation.
ParameterRCCOMSOL
ModellingSpaceNumber21 nodes28 elements
TimeBoundary
conditions
Buildings 14 03706 i001Buildings 14 03706 i002
Time step120 s10 s
Identification
results
BC1
Buildings 14 03706 i003
R±10%
C±47%
BC2
Buildings 14 03706 i004
R±3%
C±3%
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

Rendu, M.; Le Dréau, J.; Salagnac, P.; Doya, M. Influence of Boundary Conditions on the Estimation of Thermal Properties in Insulated Building Walls. Buildings 2024, 14, 3706. https://doi.org/10.3390/buildings14123706

AMA Style

Rendu M, Le Dréau J, Salagnac P, Doya M. Influence of Boundary Conditions on the Estimation of Thermal Properties in Insulated Building Walls. Buildings. 2024; 14(12):3706. https://doi.org/10.3390/buildings14123706

Chicago/Turabian Style

Rendu, Manon, Jérôme Le Dréau, Patrick Salagnac, and Maxime Doya. 2024. "Influence of Boundary Conditions on the Estimation of Thermal Properties in Insulated Building Walls" Buildings 14, no. 12: 3706. https://doi.org/10.3390/buildings14123706

APA Style

Rendu, M., Le Dréau, J., Salagnac, P., & Doya, M. (2024). Influence of Boundary Conditions on the Estimation of Thermal Properties in Insulated Building Walls. Buildings, 14(12), 3706. https://doi.org/10.3390/buildings14123706

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