A thermal bridge, also called a cold bridge, heat bridge, or thermal bypass, is an area or component of an object which has higher thermal conductivity than the surrounding materials,[1] creating a path of least resistance for heat transfer.[2] Thermal bridges result in an overall reduction in thermal resistance of the object. The term is frequently discussed in the context of a building's thermal envelope where thermal bridges result in heat transfer into or out of conditioned space.

Temperature distribution in a thermal bridge
This thermal image shows a thermal bridging of a high-rise building (Aqua in Chicago)

Thermal bridges in buildings may impact the amount of energy required to heat and cool a space, cause condensation (moisture) within the building envelope,[3] and result in thermal discomfort. In colder climates (such as the United Kingdom), thermal heat bridges can result in additional heat losses and require additional energy to mitigate.

There are strategies to reduce or prevent thermal bridging, such as limiting the number of building members that span from unconditioned to conditioned space and applying continuous insulation materials to create thermal breaks.

Concept edit

 
Thermal bridge at junction. Heat moves from the floor structure through the wall because there is no thermal break.

Heat transfer occurs through three mechanisms: convection, radiation, and conduction.[4] A thermal bridge is an example of heat transfer through conduction. The rate of heat transfer depends on the thermal conductivity of the material and the temperature difference experienced on either side of the thermal bridge. When a temperature difference is present, heat flow will follow the path of least resistance through the material with the highest thermal conductivity and lowest thermal resistance; this path is a thermal bridge.[5] Thermal bridging describes a situation in a building where there is a direct connection between the outside and inside through one or more elements that possess a higher thermal conductivity than the rest of the envelope of the building.

Identifying Thermal Bridges edit

Surveying buildings for thermal bridges is performed using passive infrared thermography (IRT) according to the International Organization for Standardization (ISO). Infrared Thermography of buildings can allow thermal signatures that indicate heat leaks. IRT detects thermal abnormalities that are linked to the movement of fluids through building elements, highlighting the variations in the thermal properties of the materials that correspondingly cause a major change in temperature. The drop shadow effect, a situation in which the surrounding environment casts a shadow on the facade of the building, can lead to potential accuracy issues of measurements through inconsistent facade sun exposure. An alternative analysis method, Iterative Filtering (IF), can be used to solve this problem.

In all thermographic building inspections, the thermal image interpretation if performed by a human operator, involving a high level of subjectivity and expertise of the operator. Automated analysis approaches, such as Laser scanning technologies can provide thermal imaging on 3 dimensional CAD model surfaces and metric information to thermographic analyses.[6] Surface temperature data in 3D models can identify and measure thermal irregularities of thermal bridges and insulation leaks. Thermal imaging can also be acquired through the use of unmanned aerial vehicles (UAV), fusing thermal data from multiple cameras and platforms. The UAV uses an infrared camera to generate a thermal field image of recorded temperature values, where every pixel represents radiative energy emitted by the surface of the building.[7]

Thermal Bridging in Construction edit

Frequently, thermal bridging is used in reference to a building’s thermal envelope, which is a layer of the building enclosure system that resists heat flow between the interior conditioned environment and the exterior unconditioned environment. Heat will transfer through a building’s thermal envelope at different rates depending on the materials present throughout the envelope. Heat transfer will be greater at thermal bridge locations than where insulation exists because there is less thermal resistance.[8] In the winter, when exterior temperature is typically lower than interior temperature, heat flows outward and will flow at greater rates through thermal bridges. At a thermal bridge location, the surface temperature on the inside of the building envelope will be lower than the surrounding area. In the summer, when the exterior temperature is typically higher than the interior temperature, heat flows inward, and at greater rates through thermal bridges.[9] This causes winter heat losses and summer heat gains for conditioned spaces in buildings.[10]

Despite insulation requirements specified by various national regulations, thermal bridging in a building's envelope remain a weak spot in the construction industry. Moreover, in many countries building design practices implement partial insulation measurements foreseen by regulations.[11] As a result, thermal losses are greater in practice that is anticipated during the design stage.

An assembly such as an exterior wall or insulated ceiling is generally classified by a U-factor, in W/m2·K, that reflects the overall rate of heat transfer per unit area for all the materials within an assembly, not just the insulation layer. Heat transfer via thermal bridges reduces the overall thermal resistance of an assembly, resulting in an increased U-factor.[12]

Thermal bridges can occur at several locations within a building envelope; most commonly, they occur at junctions between two or more building elements. Common locations include:

  • Floor-to-wall or balcony-to-wall junctions, including slab-on-grade and concrete balconies or outdoor patios that extend the floor slab through the building envelope
  • Roof/Ceiling-to-wall junctions, especially where full ceiling insulation depths may not be achieved
  • Window-to-wall junctions[13]
  • Door-to-wall junctions[13]
  • Wall-to-wall junctions[13]
  • Wood, steel or concrete members, such as studs and joists, incorporated in exterior wall, ceiling, or roof construction[14]
  • Recessed luminaries that penetrate insulated ceilings
  • Windows and doors, especially frames components
  • Areas with gaps in or poorly installed insulation
  • Metal ties in masonry cavity walls[14]

Structural elements remain a weak point in construction, commonly leading to thermal bridges that result in high heat loss and low surface temperatures in a room.

Masonry Buildings edit

While thermal bridges exist in various types of building enclosures, masonry walls experience significantly increased U-factors caused by thermal bridges. Comparing thermal conductivities between different building materials allows for assessment of performance relative to other design options. Brick materials, which are usually used for facade enclosures, typically have higher thermal conductivities than timber, depending on the brick density and wood type.[15] Concrete, which may be used for floors and edge beams in masonry buildings are common thermal bridges, especially at the corners. Depending on the physical makeup of the concrete, the thermal conductivity can be greater than that of brick materials.[15] In addition to heat transfer, if the indoor environment is not adequately vented, thermal bridging may cause the brick material to absorb rainwater and humidity into the wall, which can result in mold growth and deterioration of building envelope material.

Curtain Wall edit

Similar to masonry walls, curtain walls can experience significantly increased U-factors due to thermal bridging. Curtain wall frames are often constructed with highly conductive aluminum, which has a typical thermal conductivity above 200 W/m·K. In comparison, wood framing members are typically between 0.68 and 1.25 W/m·K.[15] The aluminum frame for most curtain wall constructions extends from the exterior of the building through to the interior, creating thermal bridges.[16]

Impacts of Thermal Bridging edit

Thermal bridging can result in increased energy required to heat or cool a conditioned space due to winter heat loss and summer heat gain. At interior locations near thermal bridges, occupants may experience thermal discomfort due to the difference in temperature.[17] Additionally, when the temperature difference between indoor and outdoor space is large and there is warm and humid air indoors, such as the conditions experienced in the winter, there is a risk of condensation in the building envelope due to the cooler temperature on the interior surface at thermal bridge locations.[17] Condensation can ultimately result in mold growth with consequent poor indoor air quality and insulation degradation, reducing the insulation performance and causing insulation to perform inconsistently throughout the thermal envelope[18]

Design Methods to Reduce Thermal Bridges edit

There are several methods that have been proven to reduce or eliminate thermal bridging depending on the cause, location, and the construction type. The objective of these methods is to either create a thermal break where a building component would span from exterior to interior otherwise, or to reduce the number of building components spanning from exterior to interior. These strategies include:

  • A continuous thermal insulation layer in the thermal envelope, such as with rigid foam board insulation[5]
  • Lapping of insulation where direct continuity is not possible
  • Double and staggered wall assemblies[19]
  • Structural Insulated Panels (SIPs) and Insulating Concrete Forms (ICFs)[19]
  • Reducing framing factor by eliminating unnecessary framing members, such as implemented with advanced framing[19]
  • Raised heel trusses at wall-to-roof junctions to increase insulation depth
  • Quality insulation installation without voids or compressed insulation
  • Installing double or triple pane windows with gas filler and low-emissivity coating[20]
  • Installing windows with thermally broken frames made of low conductivity material[20]

Analysis Methods and Challenges edit

Due to their significant impacts on heat transfer, correctly modeling the impacts of thermal bridges is important to estimate overall energy use. Thermal bridges are characterized by multi-dimensional heat transfer, and therefore they cannot be adequately approximated by steady-state one-dimensional (1D) models of calculation typically used to estimate the thermal performance of buildings in most building energy simulation tools.[21] Steady state heat transfer models are based on simple heat flow where heat is driven by a temperature difference that does not fluctuate over time so that heat flow is always in one direction. This type of 1D model can substantially underestimate heat transfer through the envelope when thermal bridges are present, resulting in lower predicted building energy use.[22]

The currently available solutions are to enable two-dimensional (2D) and three-dimensional (3D) heat transfer capabilities in modeling software or, more commonly, to use a method that translates multi-dimensional heat transfer into an equivalent 1D component to use in building simulation software. This latter method can be accomplished through the equivalent wall method in which a complex dynamic assembly, such as a wall with a thermal bridge, is represented by a 1D multi-layered assembly that has equivalent thermal characteristics.[23]

See also edit

References edit

  1. ^ Binggeli, C. (2010). Building Systems for Interior Designers. Hoboken, NJ: John Wiley & Sons.
  2. ^ Gorse, Christopher A., and David Johnston (2012). "Thermal bridge", in Oxford Dictionary of Construction, Surveying, and Civil Engineering. 3rd ed. Oxford: Oxford UP, 2012 pp. 440-441. Print.
  3. ^ Arena, Lois (July 2016). "Construction Guidelines for High R-Value Walls without Exterior Rigid Insulation" (PDF). NREL.gov. Golden, CO: National Renewable Energy Laboratory (NREL).
  4. ^ Kaviany, Massoud (2011). Essentials of Heat Transfer: Principles, Materials, and Applications. New York, NY: Cambridge University Press. ISBN 978-1107012400.
  5. ^ a b "Definition and effects of thermal bridges [ ]". passipedia.org. Retrieved 2017-11-05.
  6. ^ Previtali, Mattia; Barazzetti, Luigi; Roncoroni, Fabio (24–27 June 2013). "Spatial Data Management for Energy Efficient Envelope Retrofitting". Computational Science and Its Applications – ICCSA 2013. Lecture Notes in Computer Science. Vol. 7971. pp. 608–621. doi:10.1007/978-3-642-39637-3_48. ISBN 978-3-642-39636-6.
  7. ^ Garrido, I.; Lagüela, S.; Arias, P.; Balado, J. (1 January 2018). "Thermal-based analysis for the automatic detection and characterization of thermal bridges in buildings". Energy and Buildings. 158: 1358–1367. doi:10.1016/j.enbuild.2017.11.031. hdl:11093/1459.
  8. ^ "RR-0901: Thermal Metrics for High-Performance Walls—The Limitations of R-Value". Building Science Corporation. Retrieved 2017-11-19.
  9. ^ Grondzik, Walter; Kwok, Alison (2014). Mechanical and Electrical Equipment for Buildings. John Wiley & Sons. ISBN 978-0470195659.
  10. ^ Larbi, A. Ben (2005). "Statistical modelling of heat transfer for thermal bridges of buildings". Energy and Buildings. 37 (9): 945–951. doi:10.1016/j.enbuild.2004.12.013.
  11. ^ THEODOSIOU, T. G, and A. M PAPADOPOULOS. 2008. “The Impact of Thermal Bridges on the Energy Demand of Buildings with Double Brick Wall Constructions.” Energy and Buildings, no. 11: 2083.
  12. ^ Kossecka, E.; Kosny, J. (2016-09-16). "Equivalent Wall as a Dynamic Model of a Complex Thermal Structure". Journal of Thermal Insulation and Building Envelopes. 20 (3): 249–268. doi:10.1177/109719639702000306. S2CID 108777777.
  13. ^ a b c Christian, Jeffery; Kosny, Jan (December 1995). "Toward a National Opaque Wall Rating Label". Proceedings Thermal Performance of the Exterior Envelopes VI, ASHRAE.
  14. ^ a b Allen, E. and J. Lano, Fundamentals of Building Construction: materials and methods. Hoboken, NJ: John Wiley & Sons. 2009.
  15. ^ a b c American Society of Heating, Refrigeration and Air-Conditioning Engineers, Inc. (ASHRAE) (2017). 2017 ASHRAE Handbook: Fundamentals. Atlanta, GA: ASHRAE. ISBN 978-1939200570.{{cite book}}: CS1 maint: multiple names: authors list (link)
  16. ^ Totten, Paul E.; O’Brien, Sean M. (2008). "The Effects of Thermal Bridging at Interface Conditions". Building Enclosure Science & Technology.
  17. ^ a b Ge, Hua; McClung, Victoria Ruth; Zhang, Shenshu (2013). "Impact of balcony thermal bridges on the overall thermal performance of multi-unit residential buildings: A case study". Energy and Buildings. 60: 163–173. doi:10.1016/j.enbuild.2013.01.004.
  18. ^ Matilainen, Miimu; Jarek, Kurnitski (2002). "Moisture conditions in highly insulated outdoor ventilated crawl spaces in cold climates". Energy and Buildings. 35 (2): 175–187. doi:10.1016/S0378-7788(02)00029-4.
  19. ^ a b c California Energy Commission (CEC) (2015). Residential Compliance Manual for the 2016 Building Energy Efficiency Standards. California Energy Commission.
  20. ^ a b Gustavsen, Arild; Grynning, Steinar; Arasteh, Dariush; Jelle, Bjørn Petter; Goudey, Howdy (2011). "Key elements of and material performance targets for highly insulating window frames". Energy and Buildings. 43 (10): 2583–2594. doi:10.1016/j.enbuild.2011.05.010. OSTI 1051278. S2CID 72987269.
  21. ^ Martin, K.; Erkoreka, A.; Flores, I.; Odriozola, M.; Sala, J.M. (2011). "Problems in the calculation of thermal bridges in dynamic conditions". Energy and Buildings. 43 (2–3): 529–535. doi:10.1016/j.enbuild.2010.10.018.
  22. ^ Mao, Guofeng; Johanneson, Gudni (1997). "Dynamic Calculation of Thermal Bridges". Energy and Buildings. 26 (3): 233–240. doi:10.1016/s0378-7788(97)00005-4.
  23. ^ Kossecka, E.; Kosny, J. (January 1997). "Equivalent Wall as a Dynamic Model of a Complex Thermal Structure". J. Therm. Insul. Build. Envelopes. 20 (3): 249–268. doi:10.1177/109719639702000306. S2CID 108777777.

External links edit