Next Article in Journal
Microstructural, Electrochemical and In Vitro Analysis of Mg-0.5Ca-xGd Biodegradable Alloys
Next Article in Special Issue
Estimating Thermal Material Properties Using Solar Loading Lock-in Thermography
Previous Article in Journal
Erratum: Kim, B.-R., et al. A Study on Flow Characteristics and Flow Uniformity for the Efficient Design of a Flow Frame in a Redox Flow Battery. Appl. Sci. 2020, 10, 929
Previous Article in Special Issue
Longitudinal Monostatic Acoustic Effective Bulk Modulus and Effective Density Evaluation of Underground Soil Quality: A Numerical Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of HFM Thermal Contact on the Accuracy of In Situ Measurements of Façades’ U-Value in Operational Stage

Group of Construction Research and Innovation (GRIC), Universitat Politècnica de Catalunya·BarcelonaTech (UPC), C/Colom, 11, Ed. TR5, 08222 Terrassa, Spain
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(3), 979; https://doi.org/10.3390/app11030979
Submission received: 30 November 2020 / Revised: 15 January 2021 / Accepted: 18 January 2021 / Published: 22 January 2021
(This article belongs to the Special Issue Non-destructive Testing in Civil Engineering)

Abstract

:
Accurate information on the actual thermal transmittance of walls is vital to select appropriate energy-saving measures in existing buildings to meet the commitments of the European Green Deal. To obtain accurate results using the heat flow meter (HFM) method, good thermal contact must be made between the heat flow meter plate and the wall surface. This paper aimed to assess the influence of the non-perfect thermal contact of heat flow meter plates on the accuracy of in situ measurement of the façades’ U-value when a film was applied to avoid damage to the wall surface. Given the fact that to avoid harm to the wall surface, the laying of a film is a usual procedure in the installation of equipment during the building’s operational stage. The findings show that deviations between measured U-values when an HFM was installed directly on the wall surface and when an HFM was installed with a PVC film were found to differ significantly from the theoretical effect of including a PVC film during the monitoring process.

1. Introduction

Efficiency in terms of energy and resources is required by the European Green Deal when building and renovating the European building stock [1] for achieving the European Union’s energy efficiency targets [2]. To meet the 2050 long-term climate and energy targets, 97% of the residential building stock, which is responsible for around 36% of CO2 emissions in the European Union [3], needs to be upgraded to become highly energy efficient and obtain the Energy Performance Certificate (EPC) label A [4]. This high percentage is due partly to the average age of European residential building stock: more than 80% of residential buildings are over 25 years old (Table 1) and, therefore, most of them were built without thermal regulation. Construction is considered the sector with the most potential for energy saving. Buildings represent 40% of the energy used in the European Union [5]. Almost 26% of Europe’s total final energy consumption is used in residential buildings [6]. Space heating represents a large amount of energy consumption in the housing sector [7,8] and, therefore, has a significant capacity for energy saving.
According to the IEA EBC Annex 71-project [10], measuring the actual energy performances discloses a notable gap between actual and expected energy performances of buildings. The actual thermal behaviour of building envelopes could contribute to this energy performance gap in buildings. Deviations between the predicted and actual behaviour of envelopes are related to aspects of the design and construction stages [11]. Moreover, assumptions about energy efficiency improvements resulting from building refurbishment are not always met [12,13]. Consequently, accurate in situ measurements of the actual U-value of façades are necessary.
The thermal performance of building façades is a fundamental parameter that should be evaluated to obtain an accurate energy diagnosis of buildings [14,15,16,17,18,19,20,21]. Successful decision-making during energy renovation processes of existing buildings requires precise characterization of thermal properties of building components [22,23]. Classifying the opaque part of the façades is essential to precisely analyse the thermal performance of façades in the housing sector.
In recent years, the thermal behaviour of the materials that compose façades had been studied in depth by several authors such as Laaroussi et al. [24] measuring the thermal properties of brick materials, Kuman and Suman [25] or Björk and Enochsson [26] measuring the properties of thermal insulation materials. However, when analysing the thermal transmittance of walls, assumed U-values have been a meaningful source of error in estimations of energy savings and carbon emissions [27]. Evidence suggests that assumptions concerning heat loss from a dwelling pre-retrofit and post-retrofit are not correct [12]. Therefore, accurate on-site measurements are required to provide information on the actual thermal transmittance of façades. To effectively quantify the actual performance of buildings, optimized on-site measurements combined with dynamic data analysis techniques are needed [28].
Several methods can be used for the in situ measurement of U-value of existing buildings’ façades [29,30]. One of the most common is the heat flow meter method, standardised by ISO 9869-1:2014 [31]. This method obtains the thermal transmittance by measuring the heat flow rate that passes through a wall and the inside and outside environmental temperatures. However, difficulties can arise in on-site measurements of walls U-value in the existing building stock, leading to inaccuracies [32]. These difficulties can be classified into three groups, according to the IEA EBC Annex 58-project [33]: errors related to the measurement accuracy, errors related to the analysis of data and errors related to the boundary conditions of the in situ measurement. Difficulties related to the measurement accuracy include factors like reading and calibration of heat flow meters and temperature sensors and had been deeply analysed by authors such as Ficco et al. [17], Trethowen [34] and Meng et al. [35]. Difficulties related to the analysis of data were recently analysed in depth [32,36,37,38,39]. Difficulties related to the boundary conditions include factors as an imbalance of the heat flow, edge heat loss and accuracy on the position of sensors, which were highly analysed by Peng and Wu [32], Meng et al. [35], Cesaratto et al. [40], Ahmad et al. [41] and Guattari et al. [42]. Delving into boundary monitoring conditions, the factor of contact between the wall surface and the heat flow meter was analysed by simulations but has not been analysed in experimental tests [40], so as for the present study focuses on this aspect.
The IEA EBC Annex 58-project [33] and authors such as Cesaratto et al. [40], Tadeu et al. [43] and Gori and Elwell [44] highlight the importance of ensuring good thermal contact between the heat flow meter plate and the wall to be measured to obtain accurate results. However, conducting HFM in situ measurements during the operational stage of dwellings is challenging due to the need for avoiding damage to wall surfaces. In this sense, the usual practice is placing a PVC film to preserve the wall surface [27,45].
Considering all the aspects aforementioned, the study aimed to analyse the influence of the non-perfect thermal contact of heat flow meter plates on the accuracy of in situ measurement of the façades’ U-value when a film was applied to avoid damage to the wall surface, as a usual equipment installation procedure during the building’s operational stage. This research provides valuable additional evidence on the accuracy of in situ monitoring of the actual U-value of existing buildings’ façades and will therefore assist practitioners in pre-retrofit diagnosis.
The method used to analyse the influence of the non-perfect thermal contact of heat flow meter plates on the accuracy of in situ measurement of the façades’ U-value is explained in the second section following this introduction. The third section presents the results. Finally, the discussion and conclusions are given in the fourth section.

2. Materials and Methods

2.1. Methodology

The method for assessing the impact of the heat flow meter thermal non-perfect contact on the accuracy of in situ measurement of walls U-value consists of three steps (Figure 1):
  • First, two façades with a range of theoretical thermal transmittance values were selected.
  • Second, in situ measurements of walls’ U-value were conducted with two heat flow meters. One heat flow meter was installed by applying a layer of interface material between the heat flow meter and the wall surface, and the other heat flow meter was installed by applying a PVC film between the layer of interface material and the wall surface. During the monitoring process, recommendations on apparatus and environmental conditions were considered [31]. Then, data were analysed using the dynamic method.
  • Third, the variability of results was analysed by comparing the differences between the measured thermal transmittances obtained from the two heat flow meters with the theoretical effect of including a PVC film during the monitoring process, for the two case studies.

2.2. Case Studies

To analyse the implications of using a protective film during the installation process of heat flow meters on the accuracy of in situ measurements of the façades’ thermal transmittance, two brick masonry walls with varying theoretical thermal transmittances were selected as case studies (Case studies 1 and 2).
Case study 1 was built in 1960 and according to Gaspar et al. [46] can be defined as a single-skin wall with no air cavity or insulation. This case study is located between an interior habitable zone and an interior non-habitable zone. Case study 2 was built in 2005, it is a double-skin façade with internal insulation but no air cavities finished with continuous covering [46]. This case study is located between an indoor habitable zone and an outdoor zone.
The theoretical total thermal resistance (RT) of the two multilayer walls can be calculated as follows [47]:
R T ( m 2 · K W ) =   1 U t   = R s i + R 1 + R 2 + + R N + R s e  
where Ut is the theoretical thermal transmittance of the wall, Rsi and Rse are the internal and external superficial resistance values (0.13 m2·K/W and 0.04 m2·K/W, respectively [47]) and R1 + R2 + …. + RN are the design thermal resistance values for layers from 1 to N.
To calculate the design thermal resistance (R) of a uniform layer, the following expression was used:
R   ( m 2 · K W ) =   d   λ
where d is the thickness of the layer and λ is its design thermal conductivity.
Design data to calculate the designed thermal transmittance was obtained from the buildings’ executive projects and reports. This information was corroborated by the facility managers of the buildings. Subsequently, the thickness of the wall was measured on-site. The theoretical U-values of the two case studies were determined according to Equations (1) and (2), following guidelines of ISO 6949:2007 [47] and the Spanish Technical Building Code’s Catalogue of Building Elements [48]. Even though the theoretical U-values were not absolutely certain were taken as reference values. Table 2 provides a summary of the thickness, thermal conductivity and thermal resistance of each layer and the theoretical thermal transmittance for the case studies.

2.3. In Situ Measurement of Façades’ Thermal Transmittance

This section specifies the instrumentation and data collection process and the subsequent analysis of data.

2.3.1. Instrumentation

Proper instrumentation was carefully selected for the in situ measurement of the actual thermal transmittance of the walls. The apparatus consisted of an internal acquisition system, to which two heat flow plates and an internal environmental temperature sensor were connected, and an external environmental temperature sensor with its acquisition system. Table 3 summarises the main specifications and a priori accuracy of the calibrated instrumentation.
The monitoring process was designed following ISO guidelines [31]. Taking into account the considerations of Asdrubali et al. [14], Evangelisti et al. [16], Ahmad et al. [41], Tejedor et al. [49,50], Barreira et al. [51] and Nardi et al. [52], the placement of equipment was examined with an infrared thermographic camera (FLIR E60bx Infrared Camera). Proximity to defects, joints and borders of the wall, direct solar radiation and direct impact of heating or cooling devices were avoided, as recommended by Guattari et al. [42] and Evangelisti et al. [53].
The two heat flow meter plates were placed on the inner side of the wall due to it is the location where the temperature was most stable. Proper thermal contact was ensured between the entire area of one heat flow meter plate and the wall surface by carefully applying a layer of thermal interface material. The other heat flow meter plate was meticulously installed by applying a film between the layer of thermal interface material and the wall (Figure 2), which is usual procedure to protect the wall surface, as described in Section 1.
During the experimental campaign, climatic conditions were observed. The monitoring process took into consideration that the internal environmental temperature always exceeded the external environmental temperature, ensuring a stable heat flow direction. Alternating this flow direction could lead to inaccurate measurements, as described by Tadeu et al. [43], and could greatly influence the minimum test duration and the variability of the results [39]. Data were sampled every 1 s and recorded every 30-min averaged data in both dataloggers.
Data obtained from process monitoring is depicted in Figure 3, where Tin is the inside environmental temperature, Tout is the outside environmental temperature, qNF is the heat flow measured with a heat flow meter plate installed with direct contact on the wall surface and qF is the heat flow measured with a heat flow meter plate installed by applying a PVC film between the layer of thermal interface material and the wall surface. The experimental campaign of Case study 1 was conducted from 10–17 June 2016 (from 12:00 a.m. to 12:00 a.m.) and of Case study 2 from 24–30 October 2016 (from 12:30 p.m. to 12:30 p.m.).

2.3.2. Data Analysis

The measured thermal transmittance was determined according to the standardised dynamic method described by ISO 9869-1:2014 [31] and extensively detailed in Gaspar et al. [54]. To apply the dynamic analysis, a spreadsheet was programmed following the recommendations of Gaspar et al. [54]. The best estimate of the thermal transmittance was obtained for each cycle of 24 h. A confidence interval of 95% was adopted in the study to evaluate the quality of the thermal transmittance estimation results [31,55].
The duration of the test was evaluated considering the requirements established in the ISO 9869-1:2014 standard [31,56]. The three requirements are summarised in the following equations (Equations (3)–(5)):
D T   ( d a y s ) 3
| U m i U m i 1 U m i 1 × 100 | 5 %
| U m i = 1 I N T ( 2 × D T 3 ) U m i = D T I N T ( 2 × D T 3 ) + 1 D T U m i = D T I N T ( 2 × D T 3 ) + 1 D T × 100 | 5 %
where DT is the length of the test in days, Um is the measured U-value of the wall, index i indicates the number of the cycle and INT is the integer part.

2.4. Analysis of Variability of Results

The introduction of a PVC film during the monitoring process changes the theoretical U-value of the wall due to its inherent thermal resistance plus the thermal contact resistance between the film and the wall surface. The theoretical maximum thermal resistance associated with the inclusion of a PVC film is estimated according to Equations (1) and (2), as follows:
R f i l m _ m a x ( m 2 · K W ) = 0.0005 +   15 · 10 6 0.12 = 6.25 · 10 4
where the thermal contact resistance is usually between 0.000005 to 0.00005 m2·K/W according to Çengel and Ghajar [57], the thickness of the film is usually between 7 × 10−6 to 15 × 10−6 m, the thermal conductivity of the material is between 0.12 to 0.25 W/m·K and Rfilm_max is the maximum thermal resistance associated to the inclusion of a PVC film, considering the most insulation case of the film inclusion.
The theoretical effect of the inclusion of a PVC film on the thermal transmittance is calculated according to Equation (1) for the two case studies and shown in Table 4. The deviations between the theoretical thermal transmittance (NF) and the thermal transmittance when using a PVC film during the installation of the heat flow meter (F) are practically imperceptible values, 0.14% in Case study 1 and 0.02% in Case study 2.
To assess the use of a protective film during the installation of heat flow meters, two values of the measured thermal transmittance were calculated for each case study. The first value was obtained using a heat flow meter plate installed with a layer of thermal interface material directly on the wall surface (UmNF) and the second value was obtained using a heat flow meter plate installed with a PVC film between the layer of thermal interface material and the wall (UmF). To check the adjustment between both measured U-values, the absolute value of the relative difference between measured U-values were calculated using the following expression:
A b s o l u t e   v a l u e   o f   t h e   r e l a t i v e   d i f f e r e n c e   U m   N F U m   F ( % ) = | ( U m   N F U m   F ) U m   N F × 100 |
where UmNF is the measured U-value of the wall using the dynamic method using an HFM installed without film and UmF is the measured U-value of the wall using the dynamic method using an HFM installed with a PVC film between the layer of thermal interface material and the wall.
The variability of results was analysed by comparing the relative difference between the measured thermal transmittances obtained according to Equation (7), with the theoretical effect of including a PVC during the monitoring process shown in Table 4.

3. Results

The data acquisition process was conducted following the indications in Section 2, with a sampling duration of 168 h in Case study 1 (337 readings) and of 144 h in Case study 2 (289 readings). Two measured thermal transmittances and its confidence interval were calculated, for 24-h test cycles using the dynamic method in the two case studies [54]. One value was calculated using data from the heat flow meter plate installed with a layer of thermal interface material directly on the wall surface and the other value was calculated using data from the heat flow meter installed with a PVC film between the layer of thermal interface material and the wall.
The minimum test duration was checked considering the ISO standard [31]. The first requirement is that the sampling duration must be an integer of 24 h and at least 72 h (Equation (3)). Consequently, the second and third requirements were verified from the third day onwards. The second requirement for test completion is that the value of thermal transmittance obtained at the end of the sampling duration shall not deviate more than 5% from the value obtained 24 h before (Equation (4)). In accordance with this condition, in Case studies 1 and 2 the test could be ended after 72 h, as the requirement was met for all cycles with both HFM installation methods. The last requirement for ending the test is that the U-value obtained by analysing data from the first 2/3 of the sampling duration shall not deviate more than 5% from the value obtained from the data for the last period of the same length (Equation (5)). According to this condition, in Case study 1 the monitoring process could be ended in 96 h using an HFM installed directly on the wall surface and in 120 h when an HFM was used with a PVC film between the layer of thermal interface material and the wall surface. In Case study 2, the test could be stopped at 72 h when the heat flow meter was installed directly on the wall surface and in 120 h when the heat flow meter was installed with a PVC film between the layer of thermal interface material and the wall surface. The minimum test duration when using an HFM installed directly on the wall surface was found to be shorter than when using an HFM installed with a PVC film between the layer of thermal interface material and the wall surface.
The results of the measured thermal transmittance for the two case studies are shown in Table 5 and depicted in Figure 4, where Um-Dyn ± I95% is the measured U-value using the dynamic method, HFMNF indicates the use of a heat flow meter plate installed with a layer of thermal interface material directly on the wall surface and HFMF indicates the use of a heat flow meter installed with a PVC film between the layer of thermal interface material and the wall.
As tests lengthened, the confidence intervals were reduced. These findings are aligned with those analysed in the existing literature, in which the length of the test that was too short led to greater confidence intervals [14,37].
Deviations between the two measured thermal transmittance values were calculated following Equation (7), using an HFM installed with a layer of thermal interface material directly on the wall surface and an HFM installed with a PVC film between the layer of thermal interface material and the wall surface. These relative differences between measured U-values using both heat flow meters for the two case studies are depicted in Figure 5 and summarised in Table 6.

4. Discussion and Conclusions

This research assesses the influence of the heat flow meter plate non-perfect thermal contact on the accuracy of the in situ measurement of walls’ thermal transmittance using the standardised heat flow meter method. The assessment considers usual practices in the installation of heat flow meters, consisting of the application of a PVC film to avoid damage to the wall in occupied buildings, in two case studies with different theoretical thermal transmittance values. The experimental campaign was designed to isolate measurement errors related to non-perfect thermal contact from those related to the measurement accuracy and the analysis of data. In situ measurements were conducted in the same conditions and equipment, with two heat flow meters: one was carefully installed with a layer of thermal interface material directly on the wall surface and the other was carefully installed by applying a film between the thermal interface material and the wall. Data were analysed in 24-h periods using the dynamic method. Finally, the variability of results was analysed by comparing the deviation between the measured thermal transmittances obtained from the two heat flow meters with the theoretical effect of including a PVC film during the monitoring process.
Test completion results indicate that the minimum duration of experimental campaigns was influenced by the installation of the heat flow meter plate. Generally, when the heat flow meter was installed by applying a PVC film between the thermal interface material and the wall to protect its surface, the minimum duration of the test was longer than when the heat flow meter was directly installed on the wall surface without a PVC film. In Case study 1, the test lasted 24 h more and in Case study 2, the test lasted 48 h more.
The findings show that the use of a PVC film hinders the installation of sensors, worsening the thermal contact between sensor and wall. Relative differences between the measured U-values were found to be greater than the expected by the theoretical calculation and also by the ISO standard quantification in both case studies. Deviations between the two measured thermal transmittance values, using an HFM installed with a layer of thermal interface material directly on the wall surface and an HFM installed with a PVC film between the layer of thermal interface material and the wall surface, were found to be significantly different than those expected resulting from the calculation of the theoretical effect of including a PVC during the monitoring process summarised in Table 4. In the theoretical estimation, the use of a PVC film between the layer of thermal interface material and the wall surface had an expected deviation of 0.14% in Case study 1 and 0.02% in Case study 2. However, in the experimental tests results from the third cycle onwards showed that for Case study 1, the relative differences were around 24% to 27%, and in Case study 2 around 19% to 21%. On the other hand, random variations caused by a non-perfect thermal contact between the sensors and the surface are quantified by 5% according to ISO 9869-1:2014 [31]. The experimental results also differ significantly from those quantified by ISO standard [31]. These considerable differences between the expected differences and the measured ones might be due to collateral effects when installing an HFM with a PVC film between the layer of thermal interface material and the wall surface that hampers a good thermal contact, as Li et al. [27] suggested. Including a PVC film to protect the wall may complicate the sensor installation process and consequently, the quality of the thermal contact might be hampered. Additionally, the theoretical calculation of including a PVC film on the thermal transmittance does not consider random variations in the thermal contact of the HFM with the surface of the wall tested. Moreover, extending the duration of the test did not seem to reduce the differences between the measured U-values. Therefore, it is highly recommended to avoid the use of a PVC film between the layer of thermal interface material and the wall surface in the in situ measurement of the thermal transmittance of existing buildings’ façades when accurate results are required.
This research could be useful for practitioners when they conduct energy audits. It was found that when HFMs were installed with a PVC film between the layer of thermal interface material and the wall surface, the measured thermal transmittance was around 19 to 27% lower than when HFMs were installed with a layer of thermal interface material directly on the wall surface. Therefore, depending on the installation of heat flow meter plates, technical staff could take into account these possible deviations in the in situ measurement of the façades’ actual thermal transmittance during building’s operational stage and, consequently, propose appropriate energy retrofitting strategies.

Author Contributions

Conceptualization, K.G., M.C. and M.G.; methodology, K.G., M.C. and M.G.; investigation, K.G., M.C. and M.G.; data curation, K.G.; writing—original draft preparation, K.G.; writing—review and editing, K.G., M.C. and M.G.; visualization, K.G.; supervision, M.C. and M.G.; project administration, M.C.; funding acquisition, M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by the Government of Catalonia, Research Grant 2017 —SGR—227.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Nomenclature

dthickness of a layer [m]
HFM NFheat flow meter plate installed without a film between the sensor and the wall surface
HFM Fheat flow meter plate installed with a film between the sensor and the wall surface
Iconfidence interval [%]
q NFheat flux using a heat flow meter plate installed without a film between the sensor and the wall surface [W/m2]
q Fheat flux using a heat flow meter plate installed with a film between the sensor and the wall surface [W/m2]
Rtheoretical thermal resistance of a uniform layer [m2·K/W]
RTtheoretical total thermal resistance of an element [m2·K/W]
Tininternal environmental temperature [°C]
Toutexternal environmental temperature [°C]
Um-Dynmeasured thermal transmittance using the dynamic calculation method [W/m2·K]
Uttheoretical thermal transmittance [W/m2·K]
λdesign thermal conductivity of a material [W/m·K]

References

  1. Martynov, A.; Sushama, L.; Laprise, R. Simulation of temperate freezing lakes by one-dimensional lake models: Performance assessment for interactive coupling with regional climate models. Boreal Environ. Res. 2010, 15, 143–164. [Google Scholar] [CrossRef]
  2. Pedregosa, F.; Varoquaux, G.; Gramfort, A.; Michel, V.; Thirion, B.; Grisel, O.; Vanderplas, J. Scikit-learn: Machine Learning in Python. J. Mach. Learn. Res. 2012, 12, 2825–2830. [Google Scholar] [CrossRef]
  3. European Union. Directive (EU) 2018/844 of the European Parliament and of the Council of 30 May 2018 Amending Directive 2010/31/EU on the Energy Performance of Buildings and Directive 2012/27/EU on Energy Efficiency. 2018. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/PDF/?uri=CELEX:32018L0844&from=EN (accessed on 24 October 2019).
  4. Buildings Performance Institute Europe (BPIE). 97% of Buildings in the EU Need To Be Upgraded. 2017. Available online: http://bpie.eu/publication/97-of-buildings-in-the-eu-need-to-be-upgraded/ (accessed on 3 September 2020).
  5. European Commission. Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee and the Committee of the Regions. A Renovation Wave for Europe—Greening Our Buildings, Creating Jobs, Improving Lives. COM(2020) 662 Final. Available online: https://eur-lex.europa.eu/resource.html?uri=cellar:0638aa1d-0f02-11eb-bc07-01aa75ed71a1.0003.02/DOC_1&format=PDF (accessed on 21 October 2020).
  6. European Union. EU Energy in Figures; Publications Office of the European Union: Luxembourg, 2020. [Google Scholar]
  7. Government of Spain. 2017–2020 National Energy Efficiency Action Plan; Government of Spain: Madrid, Spain, 2017; pp. 1–216. [Google Scholar]
  8. Gangolells, M.; Casals, M.; Forcada, N.; Macarulla, M.; Cuerva, E. Energy mapping of existing building stock in Spain. J. Clean. Prod. 2016, 112, 3895–3904. [Google Scholar] [CrossRef]
  9. Buildings Performance Institute Europe (BPIE). Europe’s Buildings Under the Microscope; A Country-by-Country Review of the Energy Performance of Buildings; Buildings Performance Institute Europe (BPIE): Brussels, Belgium, 2011. [Google Scholar]
  10. Roels, S. IEA EBC Annex 71: Building Energy Performance Assessment Based on In-Situ Measurements; EBC: Birmingham, UK, 2017. [Google Scholar]
  11. Kampelis, N.; Gobakis, K.; Vagias, V.; Kolokotsa, D.; Standardi, L.; Isidori, D.; Cristalli, C.; Montagino, F.M.; Paredes, F.; Muratore, P.; et al. Evaluation of the performance gap in industrial, residential & tertiary near-Zero energy buildings. Energy Build. 2017, 148, 58–73. [Google Scholar] [CrossRef] [Green Version]
  12. Farmer, D.; Gorse, C.; Swan, W.; Fitton, R.; Brooke-Peat, M.; Miles-Shenton, D.; Johnston, D. Measuring thermal performance in steady-state conditions at each stage of a full fabric retrofit to a solid wall dwelling. Energy Build. 2017, 156, 404–414. [Google Scholar] [CrossRef]
  13. Evangelisti, L.; Guattari, C.; Gori, P.; Asdrubali, F. Assessment of equivalent thermal properties of multilayer building walls coupling simulations and experimental measurements. Build. Environ. 2018, 127, 77–85. [Google Scholar] [CrossRef]
  14. Asdrubali, F.; D’Alessandro, F.; Baldinelli, G.; Bianchi, F. Evaluating in situ thermal transmittance of green buildings masonries—A case study. Case Stud. Constr. Mater. 2014, 1, 53–59. [Google Scholar] [CrossRef]
  15. Desogus, G.; Mura, S.; Ricciu, R. Comparing different approaches to in situ measurement of building components thermal resistance. Energy Build. 2011, 43, 2613–2620. [Google Scholar] [CrossRef]
  16. Evangelisti, L.; Guattari, C.; Gori, P.; Vollaro, R.D.L. In Situ Thermal Transmittance Measurements for Investigating Differences between Wall Models and Actual Building Performance. Sustainability 2015, 7, 10388–10398. [Google Scholar] [CrossRef] [Green Version]
  17. Ficco, G.; Iannetta, F.; Ianniello, E.; Alfano, F.R.D.; Dell’Isola, M. U-value in situ measurement for energy diagnosis of existing buildings. Energy Build. 2015, 104, 108–121. [Google Scholar] [CrossRef]
  18. Hughes, M.; Palmer, J.; Cheng, V.; Shipworth, D. Sensitivity and uncertainty analysis of England’s housing energy model. Build. Res. Inf. 2013, 41, 156–167. [Google Scholar] [CrossRef]
  19. Sunikka-Blank, M.; Galvin, R. Introducing the prebound effect: The gap between performance and actual energy consumption. Build. Res. Inf. 2012, 40, 260–273. [Google Scholar] [CrossRef]
  20. Symonds, P.; Taylor, J.; Mavrogianni, A.; Davies, M.; Shrubsole, C.; Hamilton, I.; Chalabi, Z. Overheating in English dwellings: Comparing modelled and monitored large-scale datasets. Build. Res. Inf. 2016, 45, 195–208. [Google Scholar] [CrossRef]
  21. Zheng, K.; Cho, Y.K.; Wang, C.; Li, H. Noninvasive Residential Building Envelope R-Value Measurement Method Based on Interfacial Thermal Resistance. J. Arch. Eng. 2016, 22, A4015002. [Google Scholar] [CrossRef]
  22. Lucchi, E. Thermal transmittance of historical brick masonries: A comparison among standard data, analytical calculation procedures, and in situ heat flow meter measurements. Energy Build. 2017, 134, 171–184. [Google Scholar] [CrossRef]
  23. Sassine, E. A practical method for in-situ thermal characterization of walls. Case Stud. Therm. Eng. 2016, 8, 84–93. [Google Scholar] [CrossRef] [Green Version]
  24. Laaroussi, N.; Lauriat, G.; Garoum, M.; Cherki, A.; Jannot, Y. Measurement of thermal properties of brick materials based on clay mixtures. Constr. Build. Mater. 2014, 70, 351–361. [Google Scholar] [CrossRef]
  25. Kumar, A.; Suman, B. Experimental evaluation of insulation materials for walls and roofs and their impact on indoor thermal comfort under composite climate. Build. Environ. 2013, 59, 635–643. [Google Scholar] [CrossRef]
  26. Björk, F.; Enochsson, T. Properties of thermal insulation materials during extreme environment changes. Constr. Build. Mater. 2009, 23, 2189–2195. [Google Scholar] [CrossRef]
  27. Li, F.G.N.; Smith, A.; Biddulph, P.; Hamilton, I.; Lowe, R.J.; Mavrogianni, A.; Oikonomou, E.; Raslan, R.; Stamp, S.; Stone, A.; et al. Solid-wallU-values: Heat flux measurements compared with standard assumptions. Build. Res. Inf. 2015, 43, 238–252. [Google Scholar] [CrossRef] [Green Version]
  28. Roels, S. IEA EBC Annex 71: Building Energy Performance Assessment Based on Optimized In-Situ Measurements; EBC: Birmingham, UK, 2018. [Google Scholar]
  29. Bienvenido-Huertas, D.; Moyano, J.; Marín, D.; Fresco-Contreras, R.; Marín-García, D. Review of in situ methods for assessing the thermal transmittance of walls. Renew. Sustain. Energy Rev. 2019, 102, 356–371. [Google Scholar] [CrossRef]
  30. Soares, N.; Martins, C.; Gonçalves, M.; Santos, P.; Da Silva, L.S.; Costa, J.J. Laboratory and in-situ non-destructive methods to evaluate the thermal transmittance and behavior of walls, windows, and construction elements with innovative materials: A review. Energy Build. 2019, 182, 88–110. [Google Scholar] [CrossRef]
  31. International Organization for Standardization (ISO). Thermal Insulation—Building Elements—In-Situ Measurement of Thermal Resistance and Thermal Transmittance—Part 1: Heat Flow Meter Method; ISO Standard: Geneva, Switzerland, 2014. [Google Scholar]
  32. Peng, C.; Wu, Z. In situ measuring and evaluating the thermal resistance of building construction. Energy Build. 2008, 40, 2076–2082. [Google Scholar] [CrossRef]
  33. Janssens, A. IEA EBC Annex 58: Reliable Building Energy Performance Characterisation Based on Full Scale Dynamic Measurements; KU Leuven: Leuven, Belgium, 2016. [Google Scholar]
  34. Trethowen, H. Measurement errors with surface-mounted heat flux sensors. Build. Environ. 1986, 21, 41–56. [Google Scholar] [CrossRef]
  35. Meng, X.; Yan, B.; Gao, Y.; Wang, J.; Zhang, W.; Long, E. Factors affecting the in situ measurement accuracy of the wall heat transfer coefficient using the heat flow meter method. Energy Build. 2015, 86, 754–765. [Google Scholar] [CrossRef]
  36. Albatici, R.; Tonelli, A.M.; Chiogna, M. A comprehensive experimental approach for the validation of quantitative infrared thermography in the evaluation of building thermal transmittance. Appl. Energy 2015, 141, 218–228. [Google Scholar] [CrossRef]
  37. Nardi, I.; Ambrosini, D.; De Rubeis, T.; Sfarra, S.; Perilli, S.; Pasqualoni, G. A comparison between thermographic and flow-meter methods for the evaluation of thermal transmittance of different wall constructions. J. Phys. Conf. Ser. 2015, 655, 012007. [Google Scholar] [CrossRef] [Green Version]
  38. Deconinck, A.-H.; Roels, S. Comparison of characterisation methods determining the thermal resistance of building components from onsite measurements. Energy Build. 2016, 130, 309–320. [Google Scholar] [CrossRef] [Green Version]
  39. Atsonios, I.A.; Mandilaras, I.; Kontogeorgos, D.A.; Founti, M.A. A comparative assessment of the standardized methods for the in–situ measurement of the thermal resistance of building walls. Energy Build. 2017, 154, 198–206. [Google Scholar] [CrossRef]
  40. Cesaratto, P.G.; De Carli, M.; Marinetti, S. Effect of different parameters on the in situ thermal conductance evaluation. Energy Build. 2011, 43, 1792–1801. [Google Scholar] [CrossRef]
  41. Ahmad, A.; Maslehuddin, M.; Al-Hadhrami, L.M. In situ measurement of thermal transmittance and thermal resistance of hollow reinforced precast concrete walls. Energy Build. 2014, 84, 132–141. [Google Scholar] [CrossRef]
  42. Guattari, C.; Evangelisti, L.; Gori, P.; Asdrubali, F. Influence of internal heat sources on thermal resistance evaluation through the heat flow meter method. Energy Build. 2017, 135, 187–200. [Google Scholar] [CrossRef]
  43. Tadeu, A.; Simões, N.; Simões, I.; Pedro, F.; Škerget, L. In-Situ Thermal Resistance Evaluation of Walls Using an Iterative Dynamic Model. Numer. Heat Transf. Part A Appl. 2014, 67, 33–51. [Google Scholar] [CrossRef]
  44. Gori, V.; Elwell, C.A. Estimation of thermophysical properties from in-situ measurements in all seasons: Quantifying and reducing errors using dynamic grey-box methods. Energy Build. 2018, 167, 290–300. [Google Scholar] [CrossRef]
  45. Biddulph, P.; Gori, V.; Elwell, C.A.; Scott, C.; Rye, C.; Lowe, R.J.; Oreszczyn, T.; 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] [Green Version]
  46. Gaspar, K.; Casals, M.; Gangolells, M. Classifying System for Façades and Anomalies. J. Perform. Constr. Facil. 2016, 30, 04014187. [Google Scholar] [CrossRef] [Green Version]
  47. International Organization for Standardization. Building Components and Building Elements—Thermal Resistance and Thermal Transmittance—Calculation Method; ISO: Geneva, Switzerland, 2007. [Google Scholar]
  48. Spanish Ministry of Housing; Instituto de Ciencias de la Construcción Eduardo Torroja (IETcc). Constructive Elements Catalogue of Technical Building Code. 2010. Available online: https://www.codigotecnico.org/pdf/Programas/CEC/CAT-EC-v06.3_marzo_10.pdf (accessed on 21 April 2020).
  49. Tejedor, B.; Casals, M.; Macarulla, M.; Giretti, A. U-value time series analyses: Evaluating the feasibility of in-situ short-lasting IRT tests for heavy multi-leaf walls. Build. Environ. 2019, 159, 106123. [Google Scholar] [CrossRef]
  50. Tejedor, B.; Casals, M.; Gangolells, M. Assessing the influence of operating conditions and thermophysical properties on the accuracy of in-situ measured U -values using quantitative internal infrared thermography. Energy Build. 2018, 171, 64–75. [Google Scholar] [CrossRef]
  51. Barreira, E.; Almeida, R.M.; Delgado, J.M.P.Q. Infrared thermography for assessing moisture related phenomena in building components. Constr. Build. Mater. 2016, 110, 251–269. [Google Scholar] [CrossRef]
  52. Nardi, I.; Lucchi, E.; De Rubeis, T.; Ambrosini, D. Quantification of heat energy losses through the building envelope: A state-of-the-art analysis with critical and comprehensive review on infrared thermography. Build. Environ. 2018, 146, 190–205. [Google Scholar] [CrossRef] [Green Version]
  53. Evangelisti, L.; Guattari, C.; Asdrubali, F. Influence of heating systems on thermal transmittance evaluations: Simulations, experimental measurements and data post-processing. Energy Build. 2018, 168, 180–190. [Google Scholar] [CrossRef]
  54. Gaspar, K.; Casals, M.; Gangolells, M. A comparison of standardized calculation methods for in situ measurements of façades U-value. Energy Build. 2016, 130, 592–599. [Google Scholar] [CrossRef]
  55. Roulet, C.; Gass, J.; Markus, I. In-Situ U-Value Measurement: Reliable Results in Shorter Time by Dynamic Interpretation of Measured Data. 1985. Available online: https://web.ornl.gov/sci/buildings/conf-archive/1985 B3 papers/057.pdf (accessed on 23 April 2020).
  56. Gaspar, K.; Casals, M.; Gangolells, M. Review of criteria for determining HFM minimum test duration. Energy Build. 2018, 176, 360–370. [Google Scholar] [CrossRef]
  57. Çengel, Y.A.; Ghajar, A.J. Heat and Mass Transfer: Fundamentals and Applications, 5th ed.; McGraw-Hill Education: New York, NY, USA, 2015. [Google Scholar]
Figure 1. Research methodology.
Figure 1. Research methodology.
Applsci 11 00979 g001
Figure 2. View of two heat flow meter plates, without and with a film between the layer of thermal interface material and the wall surface.
Figure 2. View of two heat flow meter plates, without and with a film between the layer of thermal interface material and the wall surface.
Applsci 11 00979 g002
Figure 3. Indoor temperature (Tin), outdoor temperature (Tout), heat flow measured with an heat flow meter (HFM) installed with a layer of thermal interface material directly on the wall surface (qNF) and heat flow measured with an HFM installed with a PVC film between the layer of thermal interface material and the wall surface (qF) for (a) Case study 1 and (b) Case study 2.
Figure 3. Indoor temperature (Tin), outdoor temperature (Tout), heat flow measured with an heat flow meter (HFM) installed with a layer of thermal interface material directly on the wall surface (qNF) and heat flow measured with an HFM installed with a PVC film between the layer of thermal interface material and the wall surface (qF) for (a) Case study 1 and (b) Case study 2.
Applsci 11 00979 g003
Figure 4. Theoretical and measured thermal transmittance values, (1) using an HFM installed with a layer of thermal interface material directly on the wall surface (NF) and, (2) an HFM installed with a PVC film between the layer of thermal interface material and the wall surface (F) for (a) Case study 1 and (b) Case study 2.
Figure 4. Theoretical and measured thermal transmittance values, (1) using an HFM installed with a layer of thermal interface material directly on the wall surface (NF) and, (2) an HFM installed with a PVC film between the layer of thermal interface material and the wall surface (F) for (a) Case study 1 and (b) Case study 2.
Applsci 11 00979 g004
Figure 5. Deviation between the measured thermal transmittances for the two case studies, using an HFM installed with a layer of thermal interface material directly on the wall surface and an HFM installed with a PVC film between the layer of thermal interface material and the wall surface.
Figure 5. Deviation between the measured thermal transmittances for the two case studies, using an HFM installed with a layer of thermal interface material directly on the wall surface and an HFM installed with a PVC film between the layer of thermal interface material and the wall surface.
Applsci 11 00979 g005
Table 1. Distribution of residential floor space by year of construction in the European Union 27, Switzerland and Norway [9].
Table 1. Distribution of residential floor space by year of construction in the European Union 27, Switzerland and Norway [9].
RegionFloor Space DistributionAverage Age of Residential Floor Space
Pre 19601961–19901991–2010
North and West50%42%39%19%
Central and East14%35%48%17%
South36%37%49%14%
Table 2. Composition of the case studies.
Table 2. Composition of the case studies.
Case StudyNo. LayerMaterial Layer (Inside-Outside)Thickness (m)Thermal Conductivity (W/m·K)Thermal Resistance (m2·K/W)Total Thickness (m)Theoretical U-Value (W/m2·K)
Case study 1 0.130.122.20
1Mortar plaster0.010.5700.018
2Hollow brick wall0.10 0.160
3Mortar plaster0.010.5700.018
0.13
Case study 2 0.130.340.36
1Mortar plaster0.021.3000.015
2Hollow brick wall0.10 0.160
3Polyurethane insulation0.060.0282.143
4Perforated brick wall0.14 0.210
5Single-layer mortar plaster0.020.3400.059
0.04
Table 3. Main specifications of the instrumentation.
Table 3. Main specifications of the instrumentation.
Type of EquipmentModel and ManufacturerRangeA Priori Accuracy
Heat flow meter platesHFP01, Hukseflux±2000 W/m2±5%
Inside environmental temperature sensorT107, Campbell Scientific, Inc.−35° to +50 °C±0.5 °C
Inside acquisition systemCR850, Campbell Scientific, Inc.Input ±5 Vdc±0.06% of reading
Outside environmental temperature sensor and its acquisition system175T1, Instrumentos Testo, SA−35° to +50 °C±0.5 °C
Table 4. Theoretical impact of the inclusion of a PVC film on the thermal transmittance in the two case studies.
Table 4. Theoretical impact of the inclusion of a PVC film on the thermal transmittance in the two case studies.
Case StudyTheoretical U-Value NF (W/m2·K)Theoretical U-Value F
(W/m2·K)
Deviation between U-Values(%)
Case study 12.202.190.14
Case study 20.360.360.02
Table 5. Measured thermal transmittance values using the dynamic method in 24-h cycles, using an HFM installed with a layer of thermal interface material directly on the wall surface (HFMNF) and an HFM installed with a PVC film between the layer of thermal interface material and the wall surface (HFMF).
Table 5. Measured thermal transmittance values using the dynamic method in 24-h cycles, using an HFM installed with a layer of thermal interface material directly on the wall surface (HFMNF) and an HFM installed with a PVC film between the layer of thermal interface material and the wall surface (HFMF).
Duration of the Test (h)Case Study 1Case Study 2
Um-Dyn ± I95% (W/m2·K)Um-Dyn ± I95% (W/m2·K)
HFMNFHFMFHFMNFHFMF
24 h2.19 ± 0.122.03 ± 0.100.34 ± 0.020.27 ± 0.02
48 h2.21 ± 0.101.62 ± 0.050.35 ± 0.010.28 ± 0.01
72 h2.25 ± 0.091.64 ± 0.050.35 ± 0.010.27 ± 0.01
96 h2.24 ± 0.061.69 ± 0.030.34 ± 0.010.27 ± 0.01
120 h2.24 ± 0.051.70 ± 0.030.35 ± 0.010.28 ± 0.01
144 h2.23 ± 0.041.70 ± 0.020.35 ± 0.010.28 ± 0.01
168 h2.24 ± 0.041.70 ± 0.02--
Table 6. Influence of using a PVC film between the layer of thermal interface material and the wall surface on the in situ measurement of the thermal transmittance in the two case studies.
Table 6. Influence of using a PVC film between the layer of thermal interface material and the wall surface on the in situ measurement of the thermal transmittance in the two case studies.
Duration of the Test (h) | ( U q N F U q F ) U q N F |   ( % )
Case Study 1Case Study 2
24 h7.2720.17
48 h26.8120.01
72 h27.2720.70
96 h24.3321.74
120 h23.9420.13
144 h23.9119.09
168 h23.96-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gaspar, K.; Casals, M.; Gangolells, M. Influence of HFM Thermal Contact on the Accuracy of In Situ Measurements of Façades’ U-Value in Operational Stage. Appl. Sci. 2021, 11, 979. https://doi.org/10.3390/app11030979

AMA Style

Gaspar K, Casals M, Gangolells M. Influence of HFM Thermal Contact on the Accuracy of In Situ Measurements of Façades’ U-Value in Operational Stage. Applied Sciences. 2021; 11(3):979. https://doi.org/10.3390/app11030979

Chicago/Turabian Style

Gaspar, Katia, Miquel Casals, and Marta Gangolells. 2021. "Influence of HFM Thermal Contact on the Accuracy of In Situ Measurements of Façades’ U-Value in Operational Stage" Applied Sciences 11, no. 3: 979. https://doi.org/10.3390/app11030979

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