A comparative study on the applicability of six radiant floor, wall, and ceiling heating systems based on thermal performance analysis

https://doi.org/10.1016/j.jobe.2020.102133Get rights and content

Highlights

  • Applicability of six floor, wall, and ceiling radiant heating systems was studied.

  • Output, controllability, heat storage, building retrofit and costs were considered.

  • Wall with pipes in plaster performed best, had fast response but limited storage.

  • Classical floor heating performed consistently in all criteria.

  • Ceiling TABS was feasible only when long-term heat storage was needed.

Abstract

Holistic comparisons of radiant heating systems that would help make an informed decision on the selection of the most convenient system for the specific application are lacking. The applicability of six representative radiant floor, wall, and ceiling heating systems was therefore compared in terms of thermal output and surface area required, controllability, short-term and long-term heat storage, suitability for building retrofit, and investments. Temperature and heat flux distribution in the structure, time constant τ63, response time τ90, and the number of operating cycles were computed by a custom-made and verified software tool using the finite volume method. Thermal energy stored was used to determine the ability of energy storage, whereas investment costs indicated affordability. Wall heating with pipes attached to a thermally insulating core had the highest thermal output, was easy to control, suitable for building retrofit, and most affordable while providing limited thermal storage. The performance of the wall system was retained when locating the pipes in plasterboard separated from the core by an air gap. Floor heating performed consistently in all the aspects evaluated. It was demonstrated that inserting a metal fin between pipes and the concrete spread layer improved thermal output, controllability, and storage capacity of the floor system with minor effect on investments. Ceiling with pipes insulated from the core performed well when thermal storage was not required. Ceiling with pipes embedded in the core was only feasible when long-term heat storage was needed.

Introduction

Water-based radiant systems present a potentially viable solution for space heating because they are suitable for integration with renewable energy sources [[1], [2], [3]] and have the ability to create a comfortable thermal environment [[4], [5], [6], [7]]. The applicability of the individual system types depends on their location (floor, wall, or ceiling), the configuration of material layers, and the level of thermal mass. These characteristics are crucial for the selection of the most suitable system for the specific situation such as the construction of new building vs. retrofitting of an existing building, thermal storage vs. fast thermal response, and traditional vs. low-temperature renewable heat source.

Several studies have compared the radiant systems assuming various locations and configurations of the material layers. The first category of studies involves calculations performed for a representative fragment of the heating or cooling element. Oxizidis and Papadopoulos [8] compared the energy performance and thermal comfort created by a floor cooling system, ceiling with pipes embedded in plaster, wall with pipes located in the plaster, and a generic thermally active building system (TABS). Floor cooling consumed the least energy, but it could not provide enough cooling output to attain thermal comfort. Ning et al. [9] calculated the response times of typical radiant systems as defined in ISO 11855 [10]. The systems were classified into three categories according to their thermal response: fast, medium, and slow. The difference in response times of ceiling/floor heating/cooling systems was explained by the different heat transfer coefficients between the radiant surface and indoor air. Krajčík and Šikula [11] compared four wall cooling systems. The system's suitability depended on the requirements such as exploiting thermal storage, avoiding interventions on the interior surfaces, or attaining a rapid thermal response. The cooling output was sensitive to insulation thickness for the systems with pipes located in the thermal core and to pipe spacing for the systems having pipes underneath the surface.

The second category of studies involves experiments and computer simulations on a whole-room level. Mustakallio et al. [12] experimentally compared a chilled beam, chilled beam combined with a radiant panel, chilled ceiling combined with mixing ventilation, and four localized cooling panels combined with mixing ventilation. The differences in the thermal environment created by the systems were small. Le Dréau and Heiselberg [13] performed computer simulations of an active chilled beam, radiant floor, radiant ceiling, and radiant wall in an occupied room. Using radiant floor resulted in the lowest cooling demand but created least uniform thermal conditions. The most homogeneous thermal environment was attained by the radiant ceiling. Bojić et al. [14] numerically investigated the performance of ceiling, wall, floor, and floor-ceiling heating systems. The floor-ceiling system performed best in terms of energy and exergy saving, exergy destruction, CO2 emissions, and operation costs whereas a single ceiling system was the least preferable option. Karabay et al. [15] recommended using wall over floor heating because it provided more favourable thermal conditions with lower water temperature, thus reducing energy consumption. On the other hand, the computer simulations by Myhren and Holmberg [16] showed that the vertical temperature gradient occuring in the centre of the room was smaller for floor heating than for wall heating.

Thermal performance of radiant systems represented by, e.g., heating capacity, thermal resistance, and thermal response, is more relevant for design, testing and control of radiant systems compared to geometry and structure [9]. The thermal response is a decisive factor to determine the control strategy that is appropriate for the specific application. It is especially important for the design and operation of radiant systems with larger amounts of thermal mass. The related literature describes several approaches to assess the dynamic thermal performance of radiant systems. The literature survey has shown that response time τ63, also referred to as the time constant, is frequently used as an indicator of thermal response. Time constant represents the time to reach 63% of the final value of the surface temperature, thermal output, or room temperature [[17], [18], [19]]. Alternatively, other percentages of the final value, e.g. 80% [20], 95% [9], may be used for the calculation. Depending on the radiant system used, the response time ranges from several minutes for suspended ceiling panels up to several tens of hours for TABS [9]. A system with such a long response time requires using precise control strategy to provide comfortable conditions [[21], [22], [23]].

As reported by Ning et al. [9], a single number like τ63 or τ95 may not be fully representative of the thermal response of radiant systems. In such a case, calculating several response times, e.g., τ25, τ50, τ63, or τ80 may be needed [9,24]. Other indicators of thermal response found in the related literature include the peak value of surface temperature [25], thermal admittance and transmittance [26,27], visual comparison of the step-up and decay curves of surface temperature [20,[28], [29], [30]], the required number of operation cycles to maintain thermal output between defined boundaries [11], and the heat storage efficiency that takes into account the evolution of surface temperature until reaching steady-state [31].

The literature review has shown that the existing studies usually compare the systems only from thermal comfort or energy efficiency point of view. Most of the research focuses on systems with massive thermal layers, whereas less attention is given to lightweight systems and their comparison with the massive systems. Moreover, the potential use of radiant systems for building retrofit has rarely been considered. It follows that holistic comparisons of radiant heating systems that would help make an informed decision on the selection of the most convenient system are lacking. The present study aims to facilitate the selection of the most suitable radiant heating system for both newly constructed and renovated buildings. The emphasize is on the thermal performance of the systems while also considering their price and suitability for building retrofit.

To accomplish this, numerical investigations of six representative radiant floor, wall, and ceiling heating systems (Fig. 1) have been performed in terms of thermal output, area of the heating surface required, controllability, short-term and long-term energy storage, suitability for building retrofit, and investment costs. Thermal fields and heat flux distribution in the structure were computed to evaluate the thermal output. Time constant (τ63), response time (τ90), and the number of operating cycles were used to assess the thermal response and controllability. Thermal energy stored in the structure and investment costs were also determined to help evaluate the systems in a broader context.

Section snippets

Radiant heating systems investigated

Fig. 1 shows the composition and configuration of the heating systems studied. The systems were designed to cover the design heat load which consisted of the design heat loss (744 W) and the heat-up capacity (10% of the design heat loss). The design and construction of the systems were supposed to closely reflect reality. Therefore, the differences between the systems in this respect were considered. The pipe spacing and diameter varied depending on the heating system used, which resulted in

Methodology

The heating system was located in the living room of a residential building. The dimensions of the room were 6 m x 4 m which can be considered usual in the region of Central Europe. The room design heat load was determined to be 818 W following EN 12831-1 [32]. This heat loss corresponds to an external temperature of −12 °C and a room operative temperature of 20 °C. These temperatures are representative of the design conditions in the humid continental climate typical of, e.g., Central and

Results

Both the steady-state and dynamic thermal performance of the systems was evaluated. The steady-state performance was characterized by the temperature and heat flux distribution and thermal output. The dynamic thermal performance was analysed using time constant τ63 and response time τ95 as well as the number of operation cycles within 24 h. Thermal energy stored over 24 h and system costs were also elaborated to allow a complex comparison of the heating systems.

Sensitivity of thermal output and thermal response to system design

The parametric simulations have confirmed the sensitivity of thermal output to pipe spacing, hi, water temperature, and room temperature. The thermal output is also sensitive to the distance of pipes from the surface (Table 3). The effect of concrete and insulation thickness has not been explicitly considered in the present study, but certain guidance is provided in the existing studies. The studies have shown that insulation thickness has a substantial effect on the thermal output for systems

Conclusion

At present, holistic comparisons of various types of radiant systems that would guide the selection of the most suitable heating system for the specific situation are lacking. The results elaborated in this study should therefore facilitate the process of selecting the most convenient radiant heating system for both newly constructed and renovated buildings. To accomplish this, a comparative study of six representative radiant wall, floor and ceiling systems has been conducted in terms of their

CRediT authorship contribution statement

Jakub Oravec: Conceptualization, Methodology, Validation, Formal analysis, Investigation, Resources, Data curation, Writing - original draft, Writing - review & editing, Visualization. Ondřej Šikula: Conceptualization, Methodology, Software, Validation, Formal analysis, Resources, Writing - original draft, Writing - review & editing, Supervision, Project administration, Funding acquisition. Michal Krajčík: Conceptualization, Methodology, Validation, Formal analysis, Resources, Data curation,

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This research was supported by the Czech Science Foundation project GA 20-00630S, Brno University of Technology project FAST-K-21-6877, the Slovak Research and Development Agency under contract No. APVV-16-0126, and the Ministry of Education, Science, Research and Sport grant VEGA 1/0847/18.

References (43)

  • H. Karabay et al.

    A numerical investigation of fluid flow and heat transfer inside a room for floor heating and wall heating systems

    Energy Build.

    (2013)
  • J.A. Myhren et al.

    Flow patterns and thermal comfort in a room with panel, floor and wall heating

    Energy Build.

    (2008)
  • G. Yu et al.

    The experimental research on the heating and cooling performance of light floor radiant panels

    Procedia Eng.

    (2015)
  • K. Zhao et al.

    Dynamic performance of water-based radiant floors during start-up and high-intensity solar radiation

    Sol. Energy

    (2014)
  • M. Schmelas et al.

    Adaptive predictive control of thermo-active building systems (TABS) based on a multiple regression algorithm: first practical test

    Energy Build.

    (2016)
  • R. Hu et al.

    Operation dynamics of building with radiant cooling system based on Beijing weather

    Energy Build.

    (2017)
  • B.A. Price et al.

    Thermal response of composite building envelopes accounting for thermal radiation

    Energy Convers. Manag.

    (1995)
  • Z. Tian et al.

    Establishment and experimental validation of a dynamic heat transfer model for concrete radiant cooling slab based on reaction coefficient method

    Energy Build.

    (2014)
  • H. Sun et al.

    Research on a radiant heating terminal integrated with a thermoelectric unit and flat heat pipe

    Energy Build.

    (2018)
  • L. Shen et al.

    The optimization design and parametric study of thermoelectric radiant cooling and heating panel

    Appl. Therm. Eng.

    (2017)
  • M. Krajčík et al.

    Heat storage efficiency and effective thermal output: indicators of thermal response and output of radiant heating and cooling systems

    Energy Build.

    (2020)
  • Cited by (48)

    View all citing articles on Scopus
    View full text