Вы находитесь на странице: 1из 36

2.

24 Insulation Materials
Brock Conley, Cynthia A Cruickshank, and Christopher Baldwin, Carleton University, Ottawa, ON, Canada
r 2018 Elsevier Inc. All rights reserved.

2.24.1 Introduction 761


2.24.1.1 Purpose and Motivation 761
2.24.2 Background and Fundamentals 762
2.24.2.1 Modes of Heat Transfer 762
2.24.2.1.1 Conduction 762
2.24.2.1.2 Convection 765
2.24.2.1.3 Radiation 765
2.24.2.2 Thermal Bridging 767
2.24.2.3 Building Envelopes in Building and Voluntary Codes 768
2.24.2.4 Thermal Performance of Materials and Assemblies 769
2.24.2.5 Moisture Performance of Materials and Assemblies 770
2.24.2.6 Airtightness 771
2.24.3 Systems and Applications 772
2.24.3.1 Common Building Insulation Materials 772
2.24.3.1.1 Blanket insulation 772
2.24.3.1.2 Foam board based 774
2.24.3.1.3 Spray insulation 774
2.24.3.1.4 Insulated structural materials 774
2.24.3.2 Above-Grade Walls 775
2.24.3.3 Roofs and Attics 775
2.24.3.4 Floors and Below-Grade Walls 775
2.24.3.5 Pipes 776
2.24.3.6 Thermal Storage 776
2.24.4 Analysis and Assessment 777
2.24.4.1 Material Property Testing 777
2.24.4.1.1 Thermal conductivity 777
2.24.4.1.2 Heat capacity 777
2.24.4.1.3 Moisture permeability 778
2.24.4.2 Building Envelope Assembly Testing 778
2.24.4.2.1 Hot-box 778
2.24.4.2.2 Guarded hot-box 779
2.24.4.2.3 Evaluation under environmental conditions 779
2.24.4.2.4 Computer simulation 780
2.24.4.3 Exergy for an Insulated Wall at Steady State 780
2.24.5 Results and Discussion 781
2.24.5.1 Lifecycle Energy of Insulating Materials 781
2.24.6 Case Studies and Examples 785
2.24.6.1 Designing a High Thermal Resistance Wall With Vacuum Insulation Panels – ECHO – Team Ontario’s Entry
to the 2013 Solar Decathlon 785
2.24.6.2 Test Results 786
2.24.6.3 Building Envelope Construction 788
2.24.6.4 Lessons Learned 788
2.24.6.5 Building Enclosure Simulation 788
2.24.7 Future Directions 792
2.24.7.1 Aerogels 792
2.24.7.2 Vacuum Insulation 792
2.24.7.3 Gas-Filled Panels 793
2.24.8 Closing Remarks 793
References 794
Further Reading 794
Relevant Websites 795

760 Comprehensive Energy Systems, Volume 2 doi:10.1016/B978-0-12-809597-3.00252-2


Insulation Materials 761

Nomenclature q Heat flow, W


A Area, m2 q00 Heat flux though a point, W m2
CDD Cooling degree days, days r Radius, m
COP Coefficient of performance for heating or R Thermal resistance of a material or assembly,
cooling equipment m2 K W1
e Emissions, kg T Temperature, 1C or K
E Energy, Wh or J t Time, h
ei Emission intensity, gCO2 kWh1 DT Temperature difference, 1C or K
_
Ex Time rate of exergy, W U Thermal transmittance, W m2 K1
HDD Heating degree days, days W _ Time rate of work, W
k Thermal conductivity, m2 K W1 Dx Thickness of material, m
L Thickness of specimen, m Y Expected lifespan of building, years

Subscripts embodied Encompassing any consumption of resources


CO2 Carbon dioxide outside of operation
cool Cooling equipment heat Heating equipment
d Destruction j Surface or evaluation level
eff Effective value o Reference level

2.24.1 Introduction

2.24.1.1 Purpose and Motivation


Since 1973, the total annual global energy consumption has more than doubled, with annual greenhouse gas emissions increasing
110% in this period [1]. Globally, residential and commercial buildings account for over 20% of the annual global energy
consumption, or 25 TWh annually [2]. This is expected to increase at a rate greater than 1.5% per annum between 2010 and 2040
[2]. When looking at individual developed regions, buildings account for almost 40% of energy consumption in the United States,
29% in Canada, and 38% in Europe [2–4]. When examining annual energy consumption in the residential sector, it can be seen
that within all regions, space conditioning accounts for the greatest percentage of energy consumption ranging between 49 and
64% of annual consumption within buildings. This value is greatest in cold climates (northern countries, such as Canada), where
space heating dominates [2,3].
Based on these statistics, the reduction in required energy for space conditioning within buildings could significantly reduce the
total global energy consumption. Over the past couple of decades, the focus on reducing this energy consumption has been on the
improved performance of mechanical equipment. This focus has seen the average effectiveness of new natural gas furnaces and
boilers increase from below 70% to now approaching 98%, and heat pumps and air conditioning systems obtaining coefficient of
performance (COP) in excess of 5 [5]. These devices are approaching the theoretical limits of their performance and as such, little
room remains for decreasing energy consumption to meet space heating and cooling demands through improved efficiency of
mechanical devices. As a result, building designers and engineers are required to search for new methods to improve the energy
performance of buildings. At the same time, many building owners are looking to reduce their energy footprint and reduce
energy costs.
One of the most efficient and cost-effective methods for reducing energy consumption is improving the building envelope. An
occupied building consumes energy to provide a conditioned space that provides indoor comfort for people to live or work. As the
mechanical equipment heats or cools the space, the building is constantly exchanging energy with the outdoor environment. Space
conditioning loads are predominantly the result of heat loss, or gain through the building envelope, including the exterior walls,
roofs, basements, door, and windows among other penetrations, and the transfer of air between the conditioned and exterior
conditions. As such, a building envelope with a higher thermal resistance and a reduction of air infiltration into the building will
reduce the energy required for space conditioning. Increasing the performance of the building envelope has the potential to have
the greatest impact on reducing the energy consumption on the building. That being said, insulation cannot just be continuously
added due to a limit imposed on the envelope thickness and a diminished return on investment as more insulation is added. These
factors should be weighted and consider during the design of the building envelope.
Building insulation can be found in many different forms, including rigid boards, batt or blankets, or loose foam applied to or
in cavities. In addition to form, they have many different installation methods, including mechanical fastening, loosely filling a
cavity, or spray/blown-in. The different variety of building materials offers different thermal conductivities, and air and moisture
permeance. This chapter introduces the fundamentals required when designing a building envelope including fundamental heat
and mass transfer. A description of currently available insulating materials is presented, including the best applications for
each type with sample assemblies. The evaluation methods for material properties and assembly performance is presented.
762 Insulation Materials

Emerging, high performance insulating materials, which could form the basis of future walls and their potential applications
within building envelopes, are also discussed. Finally, a detailed case study is presented for the design process required when a
highly insulated wall assembly is integrated into the Team Ontario Solar Decathlon house.

2.24.2 Background and Fundamentals

While the envelope can significantly improve the energy efficiency and indoor comfort of a building, there are factors that need to
be considered before integration. Thermal bridges, moisture accumulation, air leakage, and long-term structural problems can
occur if poorly planned or designed. These properties related to overall performance and together contribute to a great or poor
building envelope. During the design of the building envelope, the thermal and moisture performance and the airtightness
requirements have to be balanced and integrated into the remaining parts of the building design, including the architectural
properties and mechanical equipment within the building.

2.24.2.1 Modes of Heat Transfer


The purpose of insulation is to impede the movement of heat in order to maintain the enclosed conditions. The modes of heat
transfer that exist are conduction, convection, and radiation. These modes may act alone but more commonly are acting con-
currently. Conduction is defined as the heat flow from one point to another through a medium (solid, liquid, or gas). Convection
is defined as heat flow by means of a fluid medium flowing past the heat source (medium is in motion, liquid, or gas). Radiation is
heat exchange by electromagnetic waves through a gaseous or vacuum medium between two surfaces. These are the heat transfer
modes and are constantly present in buildings.

2.24.2.1.1 Conduction
Heat conduction is the transfer of thermal energy between “neighbors” due to a temperature gradient. In solids, it is due to the
combination of vibrations of the molecules in a lattice and the energy transport by free electrons. In gases and liquids, conduction
is due to the collisions and diffusion of the molecules during their random motion. Conduction occurs in a building when heat
moves from the warm inside surface of the building enclosure to the cold outside surface. Heat can travel through a single
material, such as concrete, or perhaps multiple materials in contact, like the sheathing faced with batt insulation.
A temperature gradient is formed due to conduction across the building enclosure. The gradient is dependent on the con-
ductivity, k, of the material. A lower conductivity yields a steeper temperature gradient, and a higher conductivity yields a shallower
gradient. Therefore, a higher conductivity material provides heat to a simpler path to flow. The amount of heat transferred through
the material, q, is equal to the temperature difference across the two surfaces, DT, multiplied by the conductivity, k, and surface
area, A, and divided by the material thickness, Dx, as shown in Eq. (1), and illustrated in Fig. 1. In a building, this heat energy lost
through the enclosure needs to be replenished by heat addition to maintain the interior conditions. Another way to denote a
material’s resistance to heat flow is its thermal resistance, R.

q ¼ ðk  A  DTÞ=Dx ð1Þ

In many real systems, the conduction path flows through more than one material, with varying thermal conductivities and
thicknesses. The best method to determine the effective thermal resistance of the insulating system is by creating a resistance
network, illustrated in Fig. 2. The figure shows three materials that vary in thermal conductivity and thickness. Also shown are the
temperature gradients through the assembly. The resistance network uses the thermal resistance value of each material, which is the

Δx

Temperature
A gradient
q

Thot Tcold

Fig. 1 One-dimensional steady state conduction and temperature gradient through a solid material.
Insulation Materials 763

Δx1 Δx2 Δx3

A
Tgradient
q

k1 k2 k3

Thot Tcold

Thot Tcold

R1 R2 R3

Fig. 2 One-dimensional steady state conduction and temperature gradient through multiple layers of varying conductivity.

Δx1 Δx2 Δx3

A k3
q

k1 k2 k4

Thot Tcold

Thot R3 Tcold
R1 R4

R2

Fig. 3 Two-dimensional conduction through multiple layers of varying conductivity.

inverse of the thermal conductivity per unit thickness, and analyzes the system as a steady state heat flow when no heat generation
is present between the wall boundaries. The effective thermal resistance can be calculated using Eq. (2) when the material changes
in the direction parallel to heat flow.
     
Dx Dx Dx
RSIeff ¼ þ þ
k 1 k 2 k 3

RSIeff ¼ R1 þ R2 þ R3
ð2Þ
When materials vary in the plane perpendicular to the heat flow, as in Fig. 3, the effective resistance is the inverse of the effective
thermal resistance and is equal to the sum of the inverse of the thermal resistances, shown in Eq. (3). The resistance network shows
the thermal resistances R2 and R3 in parallel, while others are in series. An example of parallel materials can be a thermal bridge, a
variation in geometry or a change in material.
764 Insulation Materials

Wood stud
Cavity insulation Sheathing
Gypsum board Exterior insulation

A
q

Tinterior Texterior

Fig. 4 Plan view of a typical constructed building envelope in cold dry climates.

Building envelopes contain multiple different materials to address the structural and thermal requirements, represented in
Fig. 4. To augment the performance of the walls, insulation can be added between the structural members of the envelope thereby
creating a parallel resistance network. The structural members, typically wood or steel, have a much high thermal conductivity
value when compared to the insulation, which is commonly batt or fill insulation. For example, the RSI-value of the batt
insulation is 2.5 m2 K W1 (14 h 1F ft2 BTU1) and a 2  4 wood stud has a RSI-value of three times lower and covers 6% of the
wall area. The effective thermal resistance can be calculated using a modified version of Eq. (3), where only the parallel resistance
network is considered. The effective RSI-value for that section would have an RSI-value of 1.9 m2 K W1. It shows that a small area
of low thermal resistance or high thermal conductivity in a system can have a large impact on the effective performance.
In contrast, there may be a continuous layer of exterior insulation outboard of the sheathing. The continuous layer provides an
RSI-value to the envelope that can be added to the total value, such as a series connection. Continuing the example from before, if
a foam board with an RSI-value of 2.0 m2 K W1 is added continuously along the outside, thermal resistance at the wood stud
improves as much as the thermal resistance between the studs. This would provide an effective RSI-value of 3.9 m2 K W1 for the
assembly, excluding the other material layers that do not provide significant insulation.
In conduction, the addition of insulators to the boundary of the system slows the rate heat flows. For the previous example,
additional insulation to the outside or within the cavity will reduce the amount of heat transferred. The building envelope example
above shows that in order to maintain the interior conditions of the building, additional insulation is added to the outside or within
the cavity. By slowing the conduction through the envelope, the interior conditions require less energy input from mechanical
equipment. Energy savings through proper and sufficient insulating materials within the building envelope can be achieved.
Another example would be the use of a refrigerated container, such as a freezer or shipping container. The refrigerated interior
needs to maintain an air temperature below 01C to avoid contamination, but the surrounding air temperature can differ by
upwards of 251C. If the boundary or refrigerator wall insulation is insufficient, the unit consumes a large amount of energy. When
the insulation is increased, energy use is reduced proportionally. It highlights the importance of proper insulation and its
effectiveness in limiting conduction.
R2 R3
RSIeff ¼ R1 þ þ R4 ð3Þ
R2 þ R3
Insulating materials are not limited to covering flat and square surfaces. They may also be used to cover a cylindrical wall in
some cases or more commonly to insulate cylindrical objects including pipes, tubes, or domestic hot water tanks. These systems
typically contain a fluid that has a significantly different temperature compared to the surroundings. Additionally, cold pipes are
typically insulated to prevent condensation from forming, by reducing the air temperature, and consequently the dew point, at the
surface of the pipe. Radial coordinates are better at analyzing these situations, and cause a variation in calculating the RSI-value. By
beginning with the steady axisymmetric conduction expression for cylindrical coordinates, the RSI-value of the cylindrical wall per
unit length is represented by Eq. (4),
 
1 r2
RSIcylindrical wall ¼ ln ð4Þ
2pk r1
where r2 and r1 are the outer and inner radius in meter. For compounding layers of insulation or material, the principles from the
series layers wall apply. The cylindrical wall diagram and resistance network are shown in Fig. 5.
The performance of the cylindrical insulation has unique effects as it pertains to convection. Unlike square walls, as insulation
is added to the cylindrical walls, the diameter and surface area exposed to the environment increases. Since heat exchange
Insulation Materials 765

Thot
r1
r2
Tcold

Thot Tcold
Ri Rwall Ro

Fig. 5 Temperature profile with resistance diagram of an insulated cylindrical wall.

by convection is a function of the surface area, a situation may occur where insulation is added and heat exchange is increased.
This situation occurs when the thickness of insulation is below the critical radius of insulation. The critical radius is defined as the
point where the energy exchange through conduction and convection is no longer increasing. After initially insulting the
cylindrical wall, there is an increase in the rate of heat transfer because the increase in convection is larger than the reduction in
conduction caused by the insulation. From the series resistance network, the rate of heat transfer is represented in Eq. (5), where
the radius r1 and r2 are the inner and outer radiuses, k is the thermal conductivity of the insulation, h is the convective coefficient,
and Ti and T1 are the temperatures at the interior surface of the wall and the environment, respectively.
Ti  T1
qcylindrical wall ¼   ð5Þ
r2
1
2pk ln r1 þ hð2pr
1

To find the critical radius, the radius where the rate of heat transfer is maximized needs to be evaluated. This is achieved by
differentiating Eq. (5) with respect to the exterior radius and the relationship between thermal conductivity and convection
coefficient, shown in Eq. (6).
k
rcritical ¼ ð6Þ
h
This critical radius is important in many energy systems since reducing the energy exchange is paramount with the addition of
insulation. In these applications, if the improper thickness of insulation is used, this can reduce the performance or efficiency as
a whole.

2.24.2.1.2 Convection
Convection is the heat exchange though the bulk motion of fluid when a temperature difference exists. In buildings, the heat loss
from the interior air to the interior wall surface would occur through convection. Similar to the thermal conductivity property, k,
convection contains a convection coefficient, hc, that represents the amount of heat exchange between the surface or fluid and
surrounding. When conduction and convection are present, the system can be analyzed with a resistance network. In Fig. 6, a
temperature difference between the air temperature, TAir, and the surface temperature, THot, exists. An RSI-value can be created for
the convection component by taking the inverse of the convection coefficient. The resistance network can be analyzed with Ohms
Law to find the overall RSI-value.
There are two distinct types of convective heat: natural and forced convection. Which heat transfer type is applicable is
dependent on the bulk fluid motion. Natural convection is caused by naturally occurring fluid motion because of the difference in
temperature of a surface and the fluid, and results in a lower convective coefficient. An example would be inside a dwelling heated
with a radiator where the airflow is not aided by a fan. Another example is air losing heat to the cooler exterior walls, shown
schematically in Fig. 7. Forced convection has a flow rate that is generated by an external source, such as a pump or fan, in order to
increase the amount of heat transfer. An example of forced convection is high ventilation rates used in air conditioning equipment
present in a dwelling or strong winds that interact with the exterior of the building envelope. Convection is occurring on both sides
of the building envelope and it is an effective mechanism of exchanging energy.
While convective loops and air leakage within the cavities are detrimental to thermal and overall performance, convection is a
useful means to limit any overheating the home may encounter. For a building that is constantly in contact with sunlight, the
exterior wall is exposed to a large amount of solar radiation. The solar radiation causes a significant increase to the temperature
profile within the envelope, potentially reaching 601C. The hot exterior wall rejects its energy to the cooler outdoor air by forced
convection. This is also an example of two heat transfer modes that are occurring simultaneously.

2.24.2.1.3 Radiation
Radiation, previously mentioned as solar radiation, is a heat transfer mechanism through electromagnetic waves. To radiate heat,
there needs to be a temperature difference and a line of sight connection between the surfaces. While radiation is present between
any surfaces that agree with those two requirements, solar radiation is the most significant factor with respect to buildings. Between
the hot sky during sunny days and cold sky during winter nights, a large temperature difference between the building enclosure
766 Insulation Materials

Δx

hc A

Tair

Thot Tcold

Thot
Tair Tcold
Rc R1

Fig. 6 Resistance network incorporating conductive and convective heat transfer.

hc

Tinterior Texterior

Fig. 7 Building envelope with natural convection at the interior surface.

and surrounding is present. The surroundings and building envelope are constantly exchanging energy through radiation.
Radiation is evaluated with the temperature difference of the surface and surroundings to the fourth order, multiplied by the
Stefan–Boltzmann constant, s, and the surface area, shown in Eq. (7).
 
q ¼ s  A Tsurface
4
 Tsurrounding
4
ð7Þ

A building envelope is constantly exchanging energy with the interior and exterior environment, and in some cases is exposed
to all three heat transfer modes, shown in Fig. 8. In this case, heat transfer by conduction occurs through all insulating and
structural materials of the envelope, heat transfer by convection occurs at the interior and exterior surfaces through air movement
and heat transfer by radiation occurs between the outside surface and surroundings. R2–R6 represent the conduction through the
structural and insulating materials. R1 and R7 are the convective effects at the surface of the building envelope interacting with the
interior and exterior air temperatures. Finally, Rrad represents the radiation exchange between the exterior surface and the outdoors.
It was mentioned in the previous section that radiation can overheat the building and temperatures in the envelope. While the
solar radiation is a large heat source, it also provides natural lighting for the building and its occupants. Occupants commonly feel
more comfortable in areas with a larger amount of natural lighting. Windows are the most common penetration in the building
envelope used to provide the interior space with natural sunlight. However, the sunlight provides substantial heat directly to the
Insulation Materials 767

h1 Tsurround

A h2

Tair,int Tair,ext

Tinterior Texterior

Tsurr
Tair,int R3 Rrad

R1 R2 R5 R6 Tair,ext
R4 R7

Fig. 8 Building envelope resistance network including the three heat transfer modes with the environment.

space. Excess glazing can provide the occupant more visual comfort but could raise the potential for overheating and thermal
comfort. These factors need to be addressed during the design of the insulating and envelope system.
The introduction of radiant barriers to the market has provided designers a way to control the radiation. These barriers are
transparent metallic coatings that improve the emissivity of a material from around 90% (typical building materials) to as low as
3%. These materials are adhered to any insulation, roof membrane, or glazing to increase the effective thermal performance.
The three heat transfer modes (conduction, convection, and radiation) are constantly interacting with the building envelope.
Conduction can be reduced by adding insulating materials, either a higher performing material or a greater material thickness.
Convection can be limited through reducing the airflow rates at the interior setting or the addition of baffles or siding at the
exterior. The previous examples were used to highlight how resistance networks can be used to evaluate the effective RSI-value of
insulating systems. While convection and radiation are present during many heat transfer problems, insulating materials are best
used in reducing the amount of conduction the system experiences.

2.24.2.2 Thermal Bridging


A thermal bridge is an undesired path of low thermal resistance, through a material or layer of high thermal resistance. Typically,
these thermal bridges provide an easy path for heat transfer through the enclosure of a building, increasing the amount of heat loss
to the exterior and decreasing the effective thermal resistance of the building envelope. Thermal bridges within buildings are
predominantly the results of two modes. The first is through structural members within the building envelope and the second is
through the joints between individual pieces of insulation.
When designing a building, the first step is to design and develop the structure, ensuring the building is structurally sound and
will stand-up to the loads. These maybe environmental loads (wind, seismic, snow etc.) or the loads placed on the structure
through the building application (live loads). These loads are predominately met using structural members and assemblies
constructed using wood, concrete, or metal. Each of these materials has a thermal conductivity significantly greater than the
insulating materials. When integrated into the insulating layers of a building, the high conductivity materials create a thermal
bridge that reduce the effective thermal resistance of that layer.
In addition to structural members, other common thermal bridges in buildings are the joint in which two insulating materials
are connected or butted up to each other. In the perfect, ideal situation, the joint between the two materials would provide a
continuous, homogenous insulating material, however, in building applications, the perfect joint is rarely possible, and the small
air gap between materials provides an easier path for heat transfer. The thinner the insulating material, the greater the impact of the
thermal bridges between each piece of insulation on the overall performance of the layer. This is particularly important with
emerging insulating materials, such as vacuum insulation panels (VIPs), as each panel is relatively small when compared to a
typical sheet in board insulation, and the difference in thermal resistances between the joint and the center of the insulating panel
is significant.
Thermal bridges are present in many building components and designs, although in almost every occurrence, good design
can reduce or minimize thermal bridges. In residential applications, the most common, and in many cases, most significant
thermal bridge is through the wood structure, whether it be at the studs in the walls, headers, footers, or the space between floors.
768 Insulation Materials

Most common construction practices utilize the space between the structural studs to insulate the house by installing batt
insulation within the cavities. Although this is an effective use of the space and does not require additional wall thickness for
insulation, it also means that the wood studs act as part of the insulating layer. As a result, if a wall constructed with 2  6 lumber
and the studs spaced at 400 mm spacing, a continuous fiberglass batt insulating layer would provide an effective thermal resistance
of 3.5 K m2 W1, while this decreases to 3.25 K m2 W1, or a 7% decrease in effective insulating values. This decrease is much more
substantial if steel structural members are used. This becomes significant, as many new building codes require building envelopes
to meet an effective thermal resistance.
There are a number of methods that can be employed to reduce or eliminate this thermal bridge. The most common when
constructing residential structures is to include a continuous layer of insulation on the exterior of the structural layer, while
maintaining the insulation within the cavities between studs. The most common materials used for this exterior insulation are
board insulations, including expanded and extruded polystyrene (XPS), polyisocyanurate, or high-density rock wool. This provides
a continuous layer of insulation over-top of the structural members, reducing the percent decrease in thermal performance of the
wall. Another method for reducing thermal bridging because of the structural members is to use alternative framing methods and
layouts. Instead of using one thick stud, two smaller studs can be utilized, with each being offset from each other and with a
thermal break between the two layers. This technique is commonly referred to as a double wall construction, and can increase the
effective thermal resistance of the wall when employing similar thicknesses. Although this method provides greater thermal
resistances through reduction in thermal bridges, it is far more complex to construct, and as a result can significantly increase
construction costs, and has seen limited widespread implementation in practice.
Although these methods have the potential to reduce thermal bridges, it does not eliminate the thermal bridge created by the
structural member. To reduce any thermal bridging, one of two strategies must be employed. The first is to place all of the
insulation on the outside of the structural members, removing the highly conductive material from the insulating layer. This
strategy is more commonly utilized in commercial buildings, where a steel or concrete structure is erected, and then all of the
insulation is attached to the outside of the building. The downside to this technique is thicker wall assemblies are required, as the
insulating and structural layers are sequential as opposed to a combined single layer.
The second method that can be employed is using structural materials that can also insulate. Currently, most structural
materials provide little insulating value, as typically insulating materials are lightweight with little to no rigidity and cannot be
used as a structural member. As such, one of the few materials currently available that provide the required building structure,
while also insulating are structurally insulated panels (SIPs), which use an insulating foam core with structural facings (most
commonly orientated strand board) affixed on each side of the foam. These panels are suitable for residential and low-rise
applications, and can provide much higher effective thermal resistances for the same thickness, as a result, of eliminating the
thermal bridge within the building envelope.
In addition to thermal bridging through structural components and gaps in the insulation, many common building com-
ponents also creates thermal bridges within the building envelope. Windows and doors produce some of the most significant
thermal bridges within a building. Most window and doorframes are constructed from wood, aluminum, or polyvinyl chloride
(PVC) plastics. These all create excellent conduits for heat to escape and are why window and doorframes are some of the most
significant penetrations in any building. To reduce the thermal bridge created by these components, new materials are being
integrated into these components. Window frames are being constructed out of insulating materials, including fiberglass and PVC
plastic with embedded foam insulation. Aluminum spacers are being replaced with foam and rubber spacers, increasing the
thermal performance of the window unit. Although significant advancements have been made in residential window units, almost
all commercial and high-rise glazing units continue to use aluminum frames, and is one of the reasons high-rise wall assemblies
have significantly lower effective thermal resistances when compared to low-rise residential units.
Windows and doors are not the only architectural building features where significant thermal bridges exist. For example,
balcony slabs on high-rise buildings are concrete cutouts that extend past the exterior of the building and act as a thermal bridge.
The building itself is insulated, but the balcony slab is uninsulated. During cold windy days, the slab can become very cold,
drawing heat out of the building through the opening required to attach the balcony to the main floor slab. Effectively, the balcony
slabs become heat transfer fins that continuously removes heat from the building. Significant research is being conducted on
different methods of creating a thermal break between the buildings. These include using a strip of dense insulation at the interface
of the building and the balcony. Rebar and structural supports continue to pass through the dense insulation, but the thermal
bridge is significantly reduced when compared to continuous concrete. This cannot only reduce the amount of energy loss through
the thermal bridge, but also increase the occupant comfort by reducing cold spots within the building in close proximity that can
occur near the balcony connection.
When designing an insulating system for a building, for example, it is important to not only consider the material and the
quantity of insulation. The design of the insulating system is of equal importance, as it is critical that a continuous layer of
insulation is present. Insulation is only as good as its weakest point, and even a building envelope that has a very high nominal
insulating value can actually perform very poorly if significant thermal bridging is present.

2.24.2.3 Building Envelopes in Building and Voluntary Codes


When designing a building envelope, it is necessary to examine the local building codes. In recent years, during the development of
new codes, a greater emphasis has been placed on the energy efficiency of buildings through the reduction of energy consumption
Insulation Materials 769

in the building. The changes in code requirements are happening on a global scale, with many governing agencies implementing
some type of energy clause within their building codes. Although each jurisdiction is responsible for implementing their own
codes, many use global codes as the basis for their own codes.
Two such codes are the International Energy Conservation Code (IECC), developed by the International Code Council, and
ASHRAE 90.1 [6,7]. Both of these codes prescribe a minimum effective thermal resistance based on the climate zone of the
building. A climate zone is a region with similar meteorological conditions, and in the case of prescribe insulating levels,
is predominantly based on the heating and or cooling degree days (CDD) for a given location. Based on these zones, the IECC
2012, the minimum required thermal resistance range from 2.8 m2 K W1 (15.8 h 1F ft2 BTU1) for warm temperate climates to
4.9 m2 K W1 (27.7 h 1F ft2 BTU1) for cold, heating dominated climates. ASHRAE defines similar climate zones, and prescribes
between 2.3 m2 K W1 (13 h 1F ft2 BTU1) and 5.0 m2 K W1 (28.4 h 1F ft2 BTU1) for the same conditions.
Individual countries have developed or adopted their own energy codes. A worldwide leader in energy efficiency, Sweden was
one of the first countries that implanted a strict energy code in the early 1970s, and currently require exterior walls to have a
minimum thermal resistance of between 7 m2 K W1 (37.7 h 1F ft2 BTU1) and 8 m2 K W1 (45.3 h 1F ft2 BTU1) depending on
the orientation of the wall [8]. Another northern country with its own energy code is Canada, which requires exterior wall
insulating values of between 2.8 m2 K W1 (15.9 h 1F ft2 BTU1) and 3.1 m2 K W1 (17.6 h 1F ft2 BTU1) [9]. In Canada, each
province is responsible for its own energy code, some of which are more stringent than the Canadian Energy Code. One of these is
Ontario, which has increased the minimum required thermal resistance of exterior walls for new buildings in Ontario, requiring a
minimum effective thermal resistance of 4.2 m2 K W1 (23.8 h 1F ft2 BTU1) and 4.8 m2 K W1 (27.1 h 1F ft2 BTU1) depending
on the other installed energy saving features in the house. Japan has also implemented its own code with a wide range of
prescribed values, as there is a wide range of climatic conditions within Japan. These range from 0.6 m2 K W1 (3.4 h 1F ft2 BTU1)
in the southern areas to 2.6 m2 K W1 (14.7 h 1F ft2 BTU1) in the northern areas for residential buildings. For commercial
buildings, an effective thermal resistance is not prescribed, but rather a maximum energy use intensity of the building. The
building envelope must be designed to meet the required level [8].
The codes prescribe the minimum required performance of building envelopes. However, they do not exclude the use of
additional insulation. Worldwide some national and international voluntary standards exist that can certify a building to a higher
energy performance standard. To meet many of these voluntary standards, the thermal resistance of the building envelope must
exceed, and in some cases greatly exceed the prescribed local codes. Most of these standards are performance-based codes, meaning
a building must meet a certain energy consumption target, but how that target is met is up to the designer. As such, a wide range of
thermal resistances for building envelopes can be seen within a building and achieve the same voluntary code. In addition to the
thermal resistance of the building envelope, other factors must be considered within many of these codes, including the materials
selected (limiting the amount of embodied energy), the airtightness of the envelope, the window area to wall area ratio and the
airtightness of the building. These voluntary standards that may be selected by the designer or owner of the building may include
Passive House [10], Leadership in Energy and Environmental Design (LEED) [11], Green Globes [12], among many others. In
addition to these programs run by different organizations and governments, other initiatives are available including designing net-
zero or net-zero ready buildings [13].
Through both local building codes and voluntary standards, an overall effective thermal resistance of energy performance of the
building is provided, but how that value code is met is up to the designer. The decisions include the type of insulation selected, the
thickness and placement of each layer of insulation and the detailing that is done to prevent air and moisture infiltration. Within
many of these building codes, provisions for air and vapor barriers are included, and it is of vital importance to understand the
interaction between insulating layers and these barriers when designing a complete wall assembly. Although thermal performance
is important, it is not the only factor in designing the envelope. These items and how they impact the wall performance are further
discussed in the coming sections.

2.24.2.4 Thermal Performance of Materials and Assemblies


Typically, to reduce the energy transfer through an enclosure, increasing the effective thermal resistance or effective RSI-value of the
boundary will achieve those goals. In building envelopes or enclosures, a combination of materials with varying thermal con-
ductivity properties can cause problems in evaluating the overall performance. The existence of thermal bridges in building
envelopes create easier paths for heat to escape. Thermal bridges are locations or points where the thermal conductivity is higher,
allowing the heat transfer to become two-dimensional (2D) as opposed to one-dimensional (1D). In building envelopes, these
may be as small as mechanical fasteners used to build and construct the assembly or as large as the wood or metal frame that have
a significantly higher thermal conductivity when compared to the insulation used inside the cavity.
When comparing building envelope designs, there are two separate methods used to evaluate the thermal performance, the
nominal RSI-value, and the effective RSI-value. The nominal RSI-value is the RSI-value of the insulation at the center of the wall.
For example, if a nominal RSI-3.5 m2 K W1 (R-19.8 h 1F ft2 BTU1) wall is required, the equivalent value of insulation is required
in the cavity of the wall. The second method is to consider the effective RSI-value of the assembly. The effective RSI-value
encompasses not only the cavity but also the entire wall assembly components together. For example, the stud and cavity layer
RSI-value is averaged using the framing factor, which is the amount of frame in the wall area, or coverage areas for a single RSI-
value lower than the standard cavity insulation. Then, RSI-value of the remaining components are added to the averaged value to
obtain the effective value. For example, if the insulation in the cavity were RSI-5 m2 K W1 with 23% of the wall area being 38 mm
770 Insulation Materials

Roofs and attics

Building walls

Doors and windows

Basement and foundation

Fig. 9 Illustrated heat losses in a residential building.

by 140 mm (2  6 lumber) studs at 406 mm (16 in.) on-center (OC) framing, the averaged value would be approximately
RSI-3.4 m2 K W1, significantly lower than the nominal value. The building codes have begun to require the effective RSI-value
recently as opposed to the nominal value. The effective value encompasses a value that is more likely to be reached since it
encompasses the thermal bridging that will take place in standard construction.
Heat loss in a building is tied directly to the effective thermal resistance of the envelope. The heat losses or gains are caused
when the outdoor temperature is different from the indoor temperature. In colder climates, an increased amount of heat will be
lost to the outdoor environment and a larger load to the heating systems. However, increasing the effective thermal resistance of
the building envelope will reduce the heat loss. In essence, the heat loss of a building is contributed to by three different
components: heat through enclosure components (such as walls, doors, and windows), heat through ventilation, and heat through
air infiltration of building seals and openings. The summation of these three transfers will be providing a net heat loss or gain
through the boundary. The result will vary based on the indoor and outdoor temperatures, the effective thermal resistance of the
building, and the air ventilation and infiltration rate.
A building will experience heat losses through all the surfaces in contact with the environment, such as the walls, roofs, and
basements, as illustrated in Fig. 9. By increasing the amount or changing the type of insulating material used for the envelope, the
thermal performance can be improved and heat losses to the outside reduced. Improving the insulation of the building envelope
includes benefits of lower energy consumption, reduced capacity requirements for heating and cooling equipment, and improved
indoor comfort. As previously described, the total heat loss of the building is based on the envelope, ventilation, and infiltration.
By reducing the losses through the envelope or infiltration, the required capacity of the heating and cooling equipment is reducing
because the building will need less to offset any losses the building experiences. Through the lower heating, ventilation, and air
conditioning (HVAC) requirements and lower losses, the conditions in the spaces will remain more consistent and reduced airflow
rates provide greater indoor comfort for the occupants.

2.24.2.5 Moisture Performance of Materials and Assemblies


A common concern when designing building envelopes is how they able to wet and dry throughout the year. When using wood-
framed constructions, moisture trapped within the wall cavity could cause health, indoor comfort problems, and component
deterioration. Moisture can be introduced to the building envelope forced through the environment, by rainfall, for example, or by
the humid air diffusing through the envelope toward a dry air environment, which may be interior or exterior. The
inevitable introduction of moisture to the building envelope has spurred the development and analysis into where, when, and
how vapor barriers, retarders, and permeable materials should be installed [14]. The installation of these vapor-specific compo-
nents are dependent and sensitive to many items, such as climatic conditions, cladding types, and structural components, which
offer a wide variety of possibilities. Suggestions and recommendations are based on only using vapor barriers, retarders, or
permeable materials when necessary and not introducing two barriers to the building enclosure to ensure the envelope will have
drying potential inward and outward.
The vapor permeability of building materials uses the unit “perm,” where a perm is 1 ng of water vapor per second, per square
meter, per Pascal of pressure difference across the material (1 perm ¼ ng (s m2 Pa)1). While there is a difference in definitions, a
commonality between the American Society of Heating, Refrigeration, and Air-Conditioning Engineers (ASHRAE), the Canadian
General Standards Board, and the International Building Code are to denote vapor barriers as materials with 0.1 perm or less,
vapor retarders or vapor semi-impermeable materials are 1.0 perm or less, and vapor permeable materials have greater than
10 perms [15]. When materials have 1.0 and 10 perm is where the definition of vapor permeable materials will vary, with some
Insulation Materials 771

denoting the materials as either Class 2 or Class 3 vapor retarders depending on their permeability, while others call materials
vapor semipermeable. The vapor barriers were introduced to retard the migration of moisture through the assembly [15].
In addition to vapor barriers and retarders, air barriers and rain screens exist in building enclosures to limit the movement of air
and moisture. While vapor barriers are not intended to restrict the movement of air and air barriers are not intended to restrict to
movement of moisture, there are materials that do perform both in the building enclosure. Air barriers will restrict the movement
of the convective moisture transport through the building envelope. Air barriers are built to be continuous, durable, and robust
such that they are not compromised during construction and are added to eliminate the flow of air through the assembly. This, in
turn, will improve energy efficiency and indoor comfort levels of the building [16]. Rain screens are installed with the enclosure in
order to control the rain penetration on the exterior wall. They also reduce and minimize the moisture load seen behind the
cladding [17]. This has caused the building designers to integrate overhangs, windowsills, and impermeable surfaces, such as glass
or vinyl, as the exterior cladding of buildings, unlike materials that absorb water like brick, wood, or stucco. Secondly, behind the
cladding, a cavity or secondary boundary with greater water resistance is introduced to ventilate or dissipate the water back to the
exterior [17]. Air barriers and rain screens are not considered vapor barriers or retarders, but are effective and common practices in
reducing the moisture introduced into or transport through the enclosure.
The climate regions are an important factor in deciding the cladding, barriers, and structural components used for the building
enclosure. The hygrothermal loads and effects will be very different between a hot and humid climate and a cold or very cold
climate. For example, in the cold climates wetting from the interior will occur during the heating season by air movement, or vapor
diffusion caused by the higher levels of humidity on the inside compared to the outside. Control measures that can be imple-
mented is to build an airtight enclosures and add vapor retarders or barriers to restrict the air and vapor movement through the
envelope, and design the envelope to promote drying to the exterior [15]. In contrast to cold climates, hot and humid areas are
prone to high levels of interior humidity during cooling periods as well as higher humidity levels on the exterior. The buildings
must facilitate dehumidification of the indoor air, unlike cold climates [15]. Another difference is vapor retarders added toward the
exterior to limit the vapor diffusion from the outside and to promote drying toward the interior.
In new construction, initial moisture problems are nearly nonexistent since the building materials have been pretreated or have
not been exposed to hazardous conditions. Initially, the moisture concerns during new construction are based on the long-term
environmental conditions, properly sealing the building and how to prevent or retard moisture after it is introduced [18]. One
method to alleviate these concerns is to add vapor barriers to the inside of the sheathing, such that moisture cannot permeate
through the conditioned interior space.
In existing buildings, however, moisture problems may have developed since initial construction. These issues may be visible
through water stains, condensation or damaged finishes, or possibly condensation in cavities not visible without deconstructing
layers of the envelope. If the problem sources are not found and resolved during renovations, it is possible that the deterioration
and damage will continue to occur. Apart from existing issues, when updating an existing envelope it is possible that introducing
new vapor barriers or added insulation on the exterior can move the dew point inward, and cause unwanted condensation or
water entrapment. Since the envelope is exposed to the environment prior to renovation, there may be an accumulation or
existence of moisture before the additional insulation or barriers are installed.

2.24.2.6 Airtightness
While heat may be lost through the walls by conduction, convection, and radiation, which is based the thermal properties of the
construction materials, there may also be air leakage through the envelope. The insulating materials in the envelope can be
impermeable to air, however, during construction, a number of joints, seams, and penetrations, including doors and windows,
are introduced to the envelope [19]. Air barrier systems are composed of multiple air impermeable materials to reduce or restrict
the movement of air through the building envelope. The components that comprise the system will be sealed with gaskets,
weather-stripping, or sealant meant to create a continuous seal around the building, however, in reality, many unsealed sections
exist in a building.
The airtightness evaluation will differ for residential and commercial buildings. A blower door test is commonly performed
during an energy audit of a building either to determine retrofit designs or to prove that the construction complies with voluntary
standards. A blower door test uses a large fan unit, mounted on the exterior doorframe, to extract the air from the building or
building zone and creates a lower air pressure compared to the exterior. This forces the higher-pressure outdoor air to leak through
the building cracks and openings and the natural air infiltration rate can be determined by measuring the airflow for an extended
period [20]. At the same time, tracer gasses can be used to find small cracks or leaks that may exist in the home, in a similar fashion
to how infrared images are used to find localized thermal bridges in envelope components. Typically, tracer gasses previously
included common gasses, such as helium, hydrogen, oxygen, and carbon monoxide but recently, professionals have been using
nitrous oxide, sulfur hexafluoride, or various hydrocarbons [20].
In energy retrofits or energy-conscious homes, the air changes per hour (ACH) are targeted as a means to reduce the heat loss
and ventilation requirements, since it prevents the warm or conditioned air inside the dwelling from being lost to the outside.
According to the Canadian Mortgage and Housing Corporation, homes built between the 1950s and 1980s average an airtightness
above 6.0 ACH at 50 Pa [21] and during retrofitted applications, and they require a minimum of 30% improvement on
airtightness during a building envelope retrofit.
772 Insulation Materials

While airtightness can improve the overall thermal efficiency of your building, it may also cause unwanted indoor comfort
effects. Increasing the building tightness can cause occupants to sense a lack of fresh air in the indoor environment due to the
reduced air changes and ventilation. These effects can be mitigated through the addition of heat recovery ventilators (HRVs) or
other equipment to bring fresh air into the building. HRVs provide the home with fresh air from the outside, and precondition the
incoming air with the exhaust air during heating and cooling season. As homes become more airtight, pollutants may not be able
to exit and fresh air may not be able to enter, and HRVs are able to solve both of these problems [22]. Some of the HRV systems are
capable of transferring moisture and moderating the indoor humidity, and thereby improving the resistance against excess
moisture through the building envelope.

2.24.3 Systems and Applications

As the building maintains the interior conditioned space at regulated temperature, it will be constantly losing or gaining energy
from the outside depending on the heating or cooling season. During cooling season, the heat gains are typically found through
the roofs, windows, and wall of the dwelling and during heating season, heat losses are found through the floors and basements in
addition to the roofs, walls, and windows. In an effort to save on the space heating required maintaining the comfort levels,
additional insulation could be added to these sections of the building.

2.24.3.1 Common Building Insulation Materials


Many factors contribute to the selection, which are insulation material to use in the buildings. Buildings are insulated with a
variety of materials that use different types of fastening methods, air and vapor permeability, but above all else their different
thermal insulating properties, specifically RSI-values (R-values) or thermal conductivities.
The conventional building insulations can be found in loose-fill wool, a blanket roll, rigid board, sprayed, or built into
structural materials. The benefits and advantages can include the way they are installed, their RSI-value per unit thickness, where
they can be added, and their effects on thermal bridging, moisture transfer and air leakage. A summary of each insulation material
is included in Table 1. The table lists, for each type of insulation, the range of RSI-values, common areas in the building where they
would be found, methods to install the material, and the advantages of insulating the building with that material.

2.24.3.1.1 Blanket insulation


Blanket insulation is composed of flexible fibers based on fiberglass, mineral wool, plastic, or natural fibers. According to the U.S.
Department of Energy [23], blanket insulation is the most common insulation in buildings and is widely available in the form of
batts or blankets. They are available in widths and thicknesses that are suitable for standard spacing of stud and joists present in
buildings. The insulation is typically installed within unfinished walls, and floor and ceiling joists where they can be fitted inside
very easily since the insulations are made to the standard spacing outlined in building codes. However, if the spacing and
insulation width do not align, the batt or roll can be simply hand-cut or altered on-site without consequence.
Fiberglass blanket insulation is composed of fine glass fibers and provides an RSI-value that will vary based on the density of
the material. For example, the RSI-value of the lower density fiberglass batt is RSI-1.94 versus RSI-2.64 for a high-density fiberglass
batt for a 102 mm (400 ) deep cavity [24]. Mineral wool insulation is composed of either basalt and diabase, commonly named rock
wool, or slag from a blast furnace, commonly referred to as slag wool. The materials that make up mineral wool are typically
recycled materials from industrial processes. Another benefit to mineral wool is that it does not require additional materials or
chemicals to make it fire resistant [25].
Similar to mineral wool, recycled materials are common in manufacturing blanket and wool insulations. Cellulose insulation is
composed of recycled paper products that are fiberized to small pieces and tightly packed into building cavities [25]. The
insulation also does not typically require a moisture barrier in the cavity and during manufacturing chemicals may be added to the
composition to ensure the material is fire or insect retardant. Plastic fiber insulation is mainly composed of recycled plastic
products, as opposed to recycled paper used in cellulose insulation. After being made into insulation batts, the plastic fibers need
to be treated with a fire retardant such that it does not burn quickly, however, the insulation is prone to melting when exposed to
high temperatures or flames [25]. While some insulation may be difficult to handle because of the irritations, they may cause on
your skin, plastic fiber insulations avoid these issues, but they may be difficult to cut with standard tools compared to other batt
insulations.
Finally, these batts can be made of recycled natural materials, such as cotton, sheep wool, and straw among others to create the
insulation fibers [25]. While they may provide an RSI-value nearly equivalent or below that of the previously mentioned fiberglass
batts, they do offer some nonthermal benefits. Batts that use cotton insulation can be installed without any respiratory or skin
personal protective equipment and can be manufactured using the waste from clothing factories and therefore require minimal
energy to produce. Sheep wool insulation provides a comparable RSI-value and has the ability to absorb large amounts of
moisture, however, it may degrade due to the chemicals used to resist fire, insects, and mold. Straw bale insulation offers effective
sound-absorbing properties, but the expected RSI-value per unit thickness is much smaller compared to other available batt
insulations [25].
Table 1 Summary of insulation materials and typical found properties, installation techniques, and advantages

Type RSI-value per meter m K W1 Commonly installed locations in the Installation or fastening method Advantages
(R-value per inch) building
BTU h1 1F1 ft2 in1 [25,28]

Blanket: batts 21–32 (2.9–4.7) • Floors • Standard available sizes fitted for • Easy to install
and rolls • Ceilings between studs, joists, and other • Suitable for standard stud and joists
• Unfinished walls cavities • Cost effective
• Foundation walls
Loose-fill and 21–26 (2.8–3.7) • Existing walls • Blown-in using specialized • Adding insulation to finished, irregular
blown-in • Unfinished walls equipment shapes or around obstructions
• Unfinished attics • Performed by professional • Retrofits or renovations
• Difficult to reach cavities
Rigid foam 25–46 (3.6–6.7) • Floors • Indoor and outdoor applications • Relatively high RSI-value per unit
board • Ceilings • Covered by gypsum board for thickness
• Unfinished walls fire code on interior applications
• Foundation walls • Covered by weatherproof facing
for exterior installation
Spray foam 21–42 (3.0–6.0) • Existing walls • Pressurized spray-foamed • Adding insulation to finished, irregular
• Unfinished walls product shapes, or around obstructions
• Unfinished attics • Performed by professional
• Difficult to reach cavities
Structural insu- 25–34.7 (3.6–5) • Unfinished walls • Installed as a load bearing and • Insulation built into structural
lated panels • Ceilings structural component component
• Roofs • Pieces connect together for rapid • High thermal resistance
construction
• Possible need of crane or
heaving lifting equipment
Insulated con- 3.5 Effective (20 effective) • Foundation walls • Poured during construction • Insulation built into structural
crete forms • Exterior walls • Performed by professionals component

Insulation Materials
• Flooring • Reinforced with steel rods • High thermal resistance

773
774 Insulation Materials

The overall thermal performance of the blanket insulation will vary depending on the individual insulation chosen and
the depth of the cavity that they are fitted. However, on a per unit thickness basis, a standard blanket will provide between RSI
20–26 m2 K W1 per meter (R2.9-3.8 h 1F ft2 BTU1 in.1), while a high performance blanket, involving a higher density material,
may provide an RSI-value 32 m2 K W1 m1 (R-4.7 h 1F ft2 BTU1 in.1) [24].
Another valuable way to install many of the blanket insulations in through loose-fill or blown-in by professionals. Unlike the
batts, loose-fill insulation does not require any handling or cutting of the material, while being of similar composition. The fibers,
and in some cases foams as well, are used to fill wall cavities or enclosed spaces, such as attics and headers. They have the ability to
conform to the space without disturbing the structural elements or finishes in the building. Due to their ease of installation and
ability to conform around spaces, they are used significantly during retrofit or renovation activities.

2.24.3.1.2 Foam board based


Foam board insulation is provided in a rigid panel of insulation and offers an ability to easily fasten extra insulation to many facets
of the building. The product is typically composed of either polystyrene or polyurethane, and is capable of providing an RSI-value
of 27.6–45.0 m2 K W1 m1 (R4-6.5 h 1F ft2 BTU1 in.1) [24]. The insulation can be added all over a building, from the subfloor,
exterior wall, interior wall, foundation wall, attic, and roof.
The types of foam board that exist in expanded polystyrene (EPS), XPS, dense glass, and polyisocyanurate. These rigid boards
are comprised of polystyrene foam or closed-cell plastic beads as the main material component. There are different manufacturing
processes and final products available when polystyrene is used. XPS is manufactured through a continuous extrusion process and
creates the homogenous closed-cell rigid board [25]. EPS is manufactured by using heat and pressure to fuse spherical beads of
polystyrene together [11]. The performance differences between the two products are limited to the moisture permeability and RSI-
values. The condensed-beaded structure of the EPS allows moisture to be entrapped in the material and can degrade the per-
formance and the structural integrity of the foam. This is unlike XPS, which is more comparable to a closed-cell systems since it
does not allow moisture to be absorbed or condense inside the foam. Since there is a large number of companies manufacturing
these products, each product should be evaluated individually, but RSI-values can range from RSI-26.3 to RSI-34.7 m2 K W1 m1
(R-3.8 to R-5.0 h 1F ft2 BTU1 in.1) [11].
Polyisocyanurate insulation is a type of closed-cell plastic that contains low conductivity gases within its cells. This type of
insulation is able to provide a higher thermal performance compared to EPS and XPS but due to the low conductivity gas escaping
the pores over time, it will have a lower RSI-value at the end of its life [25]. However, it remains possible that the end of life
RSI-value remains above the other foam board insulations, around RSI-48 m2 K W1 m1 (R-7 h 1F ft2 BTU1 in.1). Manu-
facturers have used different metallic and plastic foils in order to reduce the degradation that the polyisocyanurate experiences with
promising results to an improved RSI-value of 60 m2 K W1 m1 (R-8.7 h 1F ft2 BTU1 in.1) [25].

2.24.3.1.3 Spray insulation


Spray insulations are liquid foam or fibrous materials and provide a versatile method of insulating difficult shapes, across
obstructions, while providing an above average RSI-value. They are blown-in by specialists or contractors typically in cavities, such
as an attic, wall cavity, or a location with an abundance of obstructions. Since they are liquid foam prior to curing, they typically
create an effective air barrier since they seal all the holes and seams. Sprayed foam insulation can also offer around two times the
RSI-value per meter thickness compared to traditional batt and blanket insulations used for the same applications [24].
The moisture transport through the insulations can differ based on their foam-in-place type. Closed-cell foams offer a high-
density of closed cells filled with a gas that enables the foam to expand and seal all the seams in the cavity. The closed-cell foam also
offers a strong resistance to heat, providing a high RSI-value per meter, and is stronger against moisture and air leakage compared to
the second foam type, open-cell foams. Open-cell foams contain a lower density of cells and are filled with air as oppose to another
expanding agent. The lack of expanding agent causes the foam to have a lower RSI-value per meter, and introduces the risk of the
foam absorbing moisture. Therefore, open-cell foams are not recommended for below ground applications [25].

2.24.3.1.4 Insulated structural materials


Materials that offer structural and insulation properties required for the building enclosures are labeled as insulated structural
materials can be found in the foundation and exterior walls of a building. Homes and buildings are commonly built with concrete
and recently, the introduction of concrete with built-in insulation. Known as concrete block insulation and insulating concrete
forms (ICFs), they offer a significant thermal improvement for basement and foundation walls compared to standard construction.
ICFs have two added layers of foam insulation to provide the benefits of plain concrete as well as improved thermal performance
and insulation values, while maintaining the appearance of a standard building [25]. Polystyrene foam interlocked and tied
together are commonly used for the insulating core material offer an effective RSI-value of 3.5 m2 K W1 (R-20 h 1F ft2 BTU1) and
will also be reinforced with steel rods [26]. While the ICFs are more costly than pouring concrete into standard forms, they do
provide an insulated wall with a vapor barrier in a single construction step and also keeps the foundation from quickly freezing or
thawing, reducing the risks of cracks, or leaks forming [26].
Another type of insulated structural material is structural insulated panels (SIP). The panels are prefabricated and beneficial
when designing rapidly deployable buildings that only require connecting the panels together to build the home enclosure.
Using these, prefabricated materials allow the home to be built much quicker than normal saving builder time and money [27].
The panels are usually built with foam-based insulation inside, like the previously discussed materials, for example, however novel
Insulation Materials 775

materials, like VIPs, have been implemented as well. The SIPs are offered in various sizes and may be so large that they require a
crane in order to assemble the structure. They have been designed with residential buildings and commercial building kits. The
panels must undergo precise and quality manufacturing to maintain their integrity and ensure their performance over the lifetime
[26,27]. They need to have smooth surfaces and edges such that they interconnect and prevent gaps on the construction site.
The connections create an airtight seal along the seams. The thin-profile insulation and airtightness created by the SIPs provide
an excellent enclosure to maintain indoor comfort in an energy efficient manner [26]. The enclosure becomes so airtight when
using SIPs that they require tightly controlled mechanical ventilation, such as HRVs, to ensure sufficient fresh air is being
introduced to the indoor environment. Other areas of concern with the SIPs are the infiltration of animals and insects into the
panels themselves, through small gaps at the connections or penetrating the panels themselves.

2.24.3.2 Above-Grade Walls


When adding insulation to the wall of a building, it can be installed on the interior or exterior of the building frame. In some cases,
installing insulation on the inside or outside are better based on the exterior finish, moisture prevention benefits or other
construction, and renovation-related questions. The walls of the building offer flexibility in the type of insulation used since they
have both cavities and structural components that allow fastening. Inside the cavities, it is possible to add nearly any batt, blanket,
blown-in, sprayed, and rigid foam board insulations.
Building walls may be built using different types of construction, such as solid walls, concrete blocks, and framed walls. The
solid walls are usually built of brick, stone, or concrete and do not contain a cavity, but could still contain a small drainage hole or
plane to removed water from the wall [29]. Concrete block walls will have hollow cores, many thermal bridges and attention
needs to be taken to ensure they are air-sealed between the interior finish and the block wall. Finally, frame walls can be built with
using wood or steel framing pieces and contain cavities that can be insulated. In addition to the cavities, framed walls can be
insulated from the interior or the exterior.
When renovating and reinsulating walls of building, it may be performed through the interior by rebuilding existing wall, to
build an additional interior wall or to renovate the exterior wall [29]. Rebuilding the existing wall will result in completely
removing the wallboard, insulation in the cavity and any barriers prior to adding extra insulation. The cavity can be refilled with
batt insulation or spray-foamed by a professional prior to reapplying the wallboard. Alternatively, the rigid foam board can be
fastened to the exposed frame studs after filling the cavity with insulation to improve the effective RSI-value. Otherwise, building a
framed wall on the interior can be done on any wall construction. However, it will cost valuable interior floor space. The new-
framed wall can be built at any distance from the old wall, as long as the cavities are completely filled and required barriers remain
intact [29]. Proper air sealing along all the edges are required for both types of renovation methods. Unlike the interior, when
adding insulation to the exterior of the walls during renovations, it is suggested to remove the cladding prior to adding insulation
[29]. Although it is possible to apply additional insulation, such as foam board or blanket insulation, to the existing cladding if the
existing moisture conditions allow the enclosure to properly dry after installation. Renovating walls is possible and can be a means
to improve the energy efficiency of the building.
Adding insulation to the building exterior for either new construction or renovation creates a moisture barrier from water
infiltration into the envelope. Installing additional exterior insulation during renovations is less intrusive to the occupants when
compared to adding insulation to the interior [30]. However, in some situation-specific instances where there are existing moisture
concerns or other renovations inside the home, adding insulation to the interior is a better method to increase the thermal
performance of the building.

2.24.3.3 Roofs and Attics


Roofs and attics are typically spaces at the top of the building and are exposed to the same conditions as the exterior walls with
increased radiation or sun exposure. These spaces are typically uninhabited, and therefore, are a convenient space to add insu-
lation, during either initial construction or renovation. Common materials used in the attic are loose-fill or batt insulation
materials since their thickness does not cause many problems. While the recommended amount of insulation varies by climate
and exterior conditions, the U.S. Department of Energy suggest that if the existing insulation in the attic is less than RSI-5.3 (R-30),
the equivalent of 279 mm (1100 ) of fiberglass or 203 mm (800 ) of cellulose approximately, additional insulation could yield energy
efficiency benefits [30]. In order to properly insulate the attic and maintain the benefits of the increased RSI-value, ensure the space
does not contain any unwanted infiltration or leaks between the conditioned interior and the attic.
Commonly used insulation in attics is batt, blanket, or loose-fill, but they are not limited to those types. It is also possible to
use foam board or spray foam and obtain comparable energy efficiency, however, in most instances, it is simpler to handle and
install in what is typically a confined space [31]. It is also important to maintain the ventilation in the attic as it reduces heat build-
up in summer months, prolongs the lifetime of the roof, acts as the second line of defense against moisture penetration and
reduces the chance of ice dams forming [31].

2.24.3.4 Floors and Below-Grade Walls


Floors may act as a barrier between conditioned and unconditioned spaces in commercial and residential buildings. A
common example of this situation in a residential building is an unconditioned garage below a conditioned space inside the
776 Insulation Materials

dwelling. During the winter months, the temperature inside the garage will approach exterior conditions and without
the insulation or proper air sealing, there will be heat loss as well as potential pollutants that may diffuse into the building
interior.
Basements are another source of heat loss during the wintertime due to the cooling of the surrounding ground. Approximately
20% of a standard home’s total heat loss can be attributed to the basement consisting of a largely uninsulated surface area below
grade [30]. Some insulation may only be performed during the initial construction of the building through the use of ICFs,
however, insulation can be still be added through renovation. The foundation walls that make up the basement will be made up of
either poured concrete, concrete block or brick none of which offer much in the way of insulating the building. The below-grade
rooms can be made more comfortable and the risk of moisture problems can be limited through the use of insulation on the
basement or foundation walls.
The basement has the ability to be insulated on the interior or exterior, each offering advantages and disadvantages. In most
situations, insulating on the interior is the most practical and economical approach and can be achieved by adding rigid foam
board or batt insulation with a wood-framed wall and gypsum board. This represents the easiest way to incorporate insulation in a
renovation project, however, must not be performed when moisture issues exist. Adding the insulation to the foundation wall will
cause the moisture to condense on the surface and rot the insulated timber wall [32]. In order to insulate the exterior of the
foundation, an excavation must occur around the building prior to waterproofing and insulating the foundation. The benefits
include moisture problems in the foundation are more visually apparent from the outside after excavating around the foundation;
there are no disruptions inside the building and freeze-thaw stresses on the foundation are eliminated after the insulation is
completed. However, some issues, in addition to the potentially high costs of excavating around the building, are the potential
difficulties associated with excavating during certain periods of the year [32].

2.24.3.5 Pipes
Pipes are used to transport fluids, such as refrigerant, steam, or conditioned water at various temperatures and to inject or reject
energy to a system. In some cases, such as district heating or building loops, the supply fluid may travel long distances before it
reaches its application.
Due to heat loss from the pipe to its environment, the supply temperature will decrease as the length of the pipe increases.
Therefore, to meet a target temperature downstream, it is necessary to increase the level of insulation on the pipe or increase the
supply temperature. The latter can be achieved using a boiler.
Pipe insulation also reduces condensation that may form along the outer surface of a cold pipe by reducing the heat transfer
between the surface of the pipe and its surroundings. For example, for indoor conditions of 211C and 40% relative humidity,
condensation will form on the pipe when the surface temperature reaches around 71C. The heat exchanged at the pipe surface due
to condensation will be transferred to the pipe fluid and increase its temperature.
Pipes are common place in buildings to transport energy to meet the heating and cooling demands of the building. They may
also be used in district heating and cooling systems, where the length of pipes is usually larger. Insulating these pipes will reduce
the load on the heating or cooling system, as well as mitigate the unwanted energy exchange during transmission.

2.24.3.6 Thermal Storage


Thermal storage is a means to store energy that is not needed during the time period that it is produced. For example, solar energy
can be stored during the day when it is available but used later in the evening when it is needed by occupants. Common thermal
storage mediums include liquid, solid, air, or a mixture. An application of thermal storage can involve the use of a heat pump and
cold and hot thermal storage, on the source and load side, respectively, to reduce peak demand on the electrical grid during the
day. In this case, the thermal storages would be charged overnight and used during the day when electricity is most expensive (e.g.,
if electricity is priced using time-of-use billing).
Insulating these systems is as important as efficiently harvesting the energy. Throughout the day or season, while the thermal
storage is not either in use or charging, it will be exchanging energy with the environment. If the storage does not have sufficient
continuous insulation or if a thermal bridge exists, the thermal energy could be lost faster than it is gained. Diurnal storage tanks
used for domestic hot water systems, for example, contain water or a liquid solution. Since these storage tanks are charged and
discharged daily, the effect of the overall heat transfer coefficient will be less due to the short time period in comparison to larger
(more surface area) tanks and longer periods.
Heat transfer characteristics are dependent on the surface and environment temperatures as well as the surface area in contact
with the environment. The storage temperature will vary in temperature depending on the type of storage that exists, its appli-
cation, temperature set-point and its location. Storages, for example, can be installed in a conditioned space (e.g., inside a
building), a nonconditioned space (e.g., exposed to the outdoor weather) or buried in the ground. The importance of insulating
storage is obvious when considering a seasonal thermal storage system, which has an annual charging and discharging cycle and is
required to remain at the set-point longer than a diurnal thermal storage intended for daily consumption. If the rate of heat
transfers are the same between the daily and seasonal storages, the seasonal storage would exchange more total energy over time
that the daily storage.
Insulation Materials 777

2.24.4 Analysis and Assessment

The main purpose of building envelopes and insulating materials are to maintain the interior conditions related by their heat and
mass transfer properties. Before the building materials can be implemented into building assemblies, it is important to understand
their material properties, including the thermal conductivity, moisture permeability, air penetration, and heat capacity. These
properties are important to determine the best use of each material, as well as to determine the interaction between materials. In
addition, material properties are vitally important in the development of heat and mass transfer computer models of proposed
wall assemblies.
Once the material properties for each component of a wall assembly are determined, the complete wall assembly must be
assessed as a unit. This is important to better understand the interaction between materials and how they operate as an assembly. It
is important to measure the total effective thermal resistance of the assembly, as this will take into account any thermal weak-
nesses, including thermal bridges (usually the result of structural members) and penetrations. The thermal assessment of insu-
lating materials and their components in the envelope are RSI-value for insulation or U-value, typically found in for windows.
These properties can be found experimentally by using many different types of apparatuses, evaluating the materials at steady state
or environmental conditions.
This section will outline the different testing methods to determine both the fundamental properties of different insulations
and structural materials, as well as measuring the heat and mass transfer within complete assemblies.

2.24.4.1 Material Property Testing


In order to understand how heat and mass transfer through the building envelope, it is necessary to understand the heat and mass
transfer characteristics of each material. The main fundamental properties of the materials required are the thermal conductivity,
heat capacitance, and moisture permeability. Other properties could be required for specific applications, including absorptivity
and transitivity, however, are not usually applicable to opaque wall sections.

2.24.4.1.1 Thermal conductivity


Thermal conductivity (k) measures the materials resistance to heat transfer through it and is inversely proportional to the thermal
resistance of the material. A common apparatus used to determine the thermal conductivity is a guarded hot plate, shown
schematically in Fig. 10. A guarded hot plate is composed of at least two plates, one with a cold temperature and one with a hot
temperature. Thermal conductivity is typically measured using a hot surface and cold surface at constant temperatures, and the
material sample between the two surfaces. The total heat transfer is measured, either using a heat flux meter or by measuring the
total heat input into the hot plate. Once the test reaches steady state conditions, which is noted as when the heat flux through the
test material stops changing and remains constant for a prolonged period of time, readings can begin.
Measurements over an extended period of time should be taken and averaged, reducing the error caused by minor fluctuations in
the measurements. An uncertainty analysis should be conducted to determine the accuracy of the measurements. American Society
for Testing Materials (ASTM) Standard C177-13 provides a detailed design guideline as to how to construct a guarded hot plate [33].
By knowing the temperature of each plate and the total heat flux, the thermal conductivity can be calculated using Eq. (8):

ðqLÞ
k¼ ð8Þ
ADT

where q is the total heat transfer through the specimen, L is the thickness of the specimen (i.e., the distance between the hot and cold
plates), A is the area in which the heat flux is being measured, and DT is the temperature difference between the hot and cold plates.

2.24.4.1.2 Heat capacity


When designing a building envelope, the heat capacity of a material relates the amount of energy that is available to be stored as
the wall heats up and cools down. The heat capacity also allows the amount of energy required for the wall to reach the steady state
to be determined. This property is typically measured one of two ways. The first is with a sample at a constant temperature; it is
placed against a hot surface, while the remaining sides are well insulated. Using temperature sensors on each surface, the amount
of energy it takes for the sample to reach the temperature of the plate can be measured. When the heat capacity is coupled with the
known rise in temperature of the sample, the heat capacitance can be determined.

Cold plate
Specimen Metering plate
Guard plate Guard plate
Specimen Edge insulation
Cold plate

Fig. 10 Schematic of a general guarded hot plate used to measure the thermal conductivity.
778 Insulation Materials

The second method uses a calorimeter. In this method, a sample is heated up to a uniform temperature. It is then placed in a
sealed, well-insulated container with a fluid (typically water) of a known volume, temperature, and heat capacitance. The material
is left within the fluid until the sample and fluid reach a constant temperature. Using the final temperature of the sample and fluid,
the heat capacitance of the sample can be calculated.

2.24.4.1.3 Moisture permeability


Moisture permeability is the material’s resistance to the water vapor diffusion through a unit of surface area. This can be
experimentally determined using one of two methods, as described in ASTM standard E96M-15 [34]. The two methods are the
desiccant method and the water method. In both methods, samples can be up to 32 mm (1.25 in.) in thickness. In the desiccant
method, a desiccant material is first placed within a watertight container. The sample is then placed across the top of the container
and sealed to the container, which is typically done using wax or a similar product. The sealed unit is then weighed using an
analytical balance and then placed into a test chamber at a constant temperature and high relative humidity. Over time, the unit is
periodically weighed, to determine the amount of moisture that has passed through the sample and accumulated in the desiccant
material. Using the weight of the moisture that has passed through the sample and the time that has passed, the moisture
permeability can be calculated, and is typically quantified as g h1 m2. The wetted method uses a similar process; however, water is
placed in the container instead of a desiccant. The container is then sealed and instead of placing it into a chamber with high
humidity, it is placed into a very dry chamber. Over time, the container is weighed to determine the amount of water that has
exited the container and using these data, the moisture transmission can be determined.

2.24.4.2 Building Envelope Assembly Testing


After measuring the thermal or moisture performance of individual building envelope components, the enclosure assemblies
should be evaluated to measure how the components perform together. This usually entails building a representative sample of the
wall assembly and testing it within a laboratory or test facility or installing instrumentation within an existing wall. Better results
are typically obtained in a laboratory or test facility, and as such is the preferred method for new wall assemblies, however,
evaluating existing walls may be necessary, for example, historical or irreproducible assemblies. The different test methods are
outlined in detailed as follows.

2.24.4.2.1 Hot-box
The main function of a hot-box apparatus is to introduce a temperature difference on either side of an assembly to measure the
heat flow and effective thermal resistance. The hot-box will regulate the air conditions on either side of the wall to replicate indoor
and outdoor conditions. The apparatus tightly controls the temperature, limits the air velocity and sometimes the relative
humidity on the indoor and outdoor side in order to reach steady state conditions and obtain a consistent heat flow through the
assembly.
In Fig. 11, a general schematic of a hot-box apparatus is shown, where the indoor and exterior chambers are labeled with the
assembly specimen located between the chambers. The interior heating unit energy consumption will be measured and will be
representative of the energy lost through the assembly. However, the hot-box will lose energy to the surrounding environment
when there is an air temperature difference between the environment and interior chamber. Due to the losses that may accrue over
a test period, the hot-box must be calibrated and using an assembly or material with a known RSI-value to quantify the heat
through the chamber walls over the test period [35].

Specimen

Baffle Baffle

Fans

Heating unit Cooling unit

Interior chamber

Exterior chamber

Fans Surround panel

Fig. 11 Schematic of a generic hot-box.


Insulation Materials 779

2.24.4.2.2 Guarded hot-box


A guarded hot-box is similar in functionality to the hot-box. However, an additional chamber is built to eliminate the heat losses
through the interior chamber walls. A third section, called the guard chamber shown in Fig. 12, is added and surrounds the interior
hot chamber. The guarded air temperature is tightly controlled to mimic the air temperature in the interior chamber. This helps to
control the heat through the chamber walls. This heat loss can be neglected if the chamber walls are sufficiently insulated and the
temperature difference between the chambers remain minimal, typically 0.11C. The heat flux of the metering chamber walls is
measured during the evaluation period, but the purpose of the guarded chamber is to have an insignificant amount of heat
through the metering chamber walls. By using the guarded hot-box apparatus, the heat flow is forced from the interior surface to
the exterior surface of the specimen and the surrounding environment does not affect the metering chamber and heat flow if the
apparatus is properly designed.
In both the guarded hot-box and hot-box apparatuses, the effective insulating value of an assembly or individual material can
be determined from steady state conditions. ASTM has developed a standard for operation, conditions, and requirements for these
apparatuses [35]. In order to ascertain that the wall specimen is experiencing steady state conditions, multiple test conditions need
to be met for five consecutive test periods. These conditions include:

• the average specimen surface temperature in the metering chamber does not vary by more than 70.251C;
• the average specimen surface temperature in the climate chamber does not vary by more than 70.251C;
• the average temperature within the air curtain in the metering chamber does not vary by greater than 70.251C; and
• the average energy input to the metering chamber does not vary by more than 71%.

Afterward, the effective RSI-value, R, of the assembly can be calculated using Eq. (9),
DT ðtAÞ
R¼ ð9Þ
E
where DT is the temperature difference between the interior and exterior surface of the assembly in 1C, E in the heat input to the
metering chamber in Wh, t is the test period in hours, and A is the metering area or the area of the wall assembly specimen in
contact with the metering chamber in m2.
In addition to the effective RSI-value of an assembly, the steady state conditions, and thermal bridge effects can be determined
by the hot-box and guarded hot-box apparatuses. In a laboratory setting, by using embedded instrumentation, such as temperature
and heat flux sensors, temperature profiles and gradients within the assembly can be experimentally determined. The gradients for
conventional building insulation materials will be through the thickness of the wall, however, using the novel materials men-
tioned later in the chapter, there can be significant temperature gradients at the interface of certain layers. The embedded sensors
are also able to evaluate the RSI-value or thermal resistance at a point in the assembly (e.g., thermal bridge, center of stud, etc.).
By measuring the heat flux and temperature difference across an interface or material, the RSI-value, R, can be calculated using
Eq. (10), where q00 is the measured heat flux in W m2 and DT is the temperature difference in 1C.
R ¼ DT=q00 ð10Þ

2.24.4.2.3 Evaluation under environmental conditions


Unlike the hot-box and guarded hot-box apparatuses, in this test method the exterior of the building envelope is evaluated, while
being exposed to the outdoor environmental conditions. This is known as in situ testing. The objective of in situ testing is to
expose the wall sections to the full seasonal cycle using the unregulated outdoor conditions to evaluate the heat and moisture
transfer through the assembly. The exposure to all weather and environmental conditions cause water infiltration and changes in

Interior chamber Specimen

Baffle Baffle

Heating unit (metered) Fans

Heating unit Cooling unit

Guarded chamber

Exterior chamber

Fans Surround panel

Fig. 12 Schematic of a generic guarded hot-box.


780 Insulation Materials

temperature and humidity that may not be accurately replicated in a laboratory setting. The in situ specimens are typically installed
for a test period of at least a year, in order to monitor the wetting and drying of the building envelope and associated insulating
materials.
During the testing period, the temperature and heat flux is measured at the interface of every layer for each cross-section to
determine the nominal thermal resistance at that specific location. After collecting sufficient data determine by regression formulas
outlined through ASTM or ISO testing standards, the weighted average effective thermal resistance of the building envelope can be
calculated based on the coverage area of each cross-section [36]. Multiple specimens can be installed in the facilities and be
independently tested and evaluated accurately since, unlike the hot-boxes, the heat addition inside the facility does not need to
be monitored or is not a variable in the thermal evaluation.
In addition to testing for moisture transfer within the wall assemblies, in situ testing can also be used to measure the thermal
performance of the wall assembly. To acquire the necessary information, a temperature sensor is installed on the interior and
exterior surfaces along with heat flux sensors. Measurements are taken at regular periods, for an extended period of time, and the
average temperature difference between the interior and exterior surfaces as well as the average heat flux are used to calculate
the thermal resistance at the point of measurement. The difference between this method and using a guarded hot-box is that only
the thermal resistance at the point of measurement is calculated, as opposed to the overall effective thermal resistance of the entire
assembly. As such, the thermal resistance at each unique cross-section must be measured. Using the calculated value of each
cross-section, and taking an area weighted average of the complete wall assembly, the overall effective thermal resistance can be
calculated.
For most in situ building envelope apparatuses, there are multiple openings to install specimens since their test period are at
least year. These opening may be oriented in different directions (e.g., North, South, East, and West, etc.) or perhaps all along the
same face of a building in order to properly compare and evaluate the envelope assemblies. Since the apparatus will be set to a
single locale for long testing periods, the calibration of computer models and simulations for an envelope design in one locale can
aid in evaluating the same design in a different climate by using readily available weather data.

2.24.4.2.4 Computer simulation


Heat and moisture transfer software have been developed to study building envelope assemblies at steady state conditions
or transient yearlong conditions. Certain programs are adept at performing hygrothermal simulations of the moisture transport
and drying potential through the building envelope over a yearlong period, which when performed in situ would require a
year worth of acquisition. Other programs offering the ability to simulate 2D, or three-dimensional (3D) heat transfer conditions
are able to provide the effective RSI-value of the assembly and the isothermal profiles through the cross-section in order to
compare different designs prior to the laboratory or in situ testing, which require more resources in terms of time, equipment, and
materials.
Two-dimensional heat transfer programs will simulate the heat transfer through the many building envelope components,
typically using a finite element heat transfer analysis. The finite element method continuously solves the heat transfer conduction
equation through each element and evaluates the total heat flow through the 2D cross-section. The program is capable of
determining the effective RSI-value of the assemblies as well as the temperature gradients along the interfaces, which identify
thermal bridge effects, and potential moisture, and condensation concerns.
The hygrothermal simulations are performed at dynamic conditions, while coupling heat and moisture transfer through
the building components. The simulation utilizes the exterior climate conditions from the chosen locale as the exterior
boundary conditions and illustrates the moisture movement that is relevant in multilayered components, such as roofs, walls,
or balconies, for example. After the simulation is completed, the interior building conditions and comfort levels can be deter-
mined, as well as whether the individual building envelope component interfaces can withstand the moisture damage accrued
over time.

2.24.4.3 Exergy for an Insulated Wall at Steady State


Exergy (Ex) is a concept based on the Second Law of Thermodynamics and defined as the maximum available energy able
to be extracted from a system before it reaches equilibrium with the environment (also known as the reference or dead state).
It is the attempt to quantify the quality or work potential of energy. In contrast to energy, exergy can be destroyed through
irreversible processes within the system. For a closed system at steady state, the exergy rate balance equation is expressed
through Eq. (11),
dEx X 
To _
¼ 1 QW _  Ex
_ d ¼0 ð11Þ
dt Tj
 
where 1  TToj Q _ is the time rate of exergy transfer accompanying heat transfer in watts, To is the reference temperature in Kelvin, Tj
is the surface temperature in Kelvin, W _ is the time rate of exergy accompanying work in watts, and Ex _ d is the time rate of exergy
destruction in watts.
Let us consider the concept of exergy for an insulated wall at steady state where one surface of the wall is at 400K and the other
surface is at 293K. An exergy analysis can be performed to determine the rate of exergy destruction through the wall. Assuming a
reference temperature of 273K and a rate of exergy transfer accompanying heat transfer of 0.3 kW m2 through the wall, the rate of
Insulation Materials 781

exergy destruction, in kW m2 of wall surface, where no work is produced is:


X 
To _
_ d =A ¼
Ex 1 Q=A
Tj
     
_ d¼ 273K kW 273K kW
Ex 1 0:3  1 0:3
400K m2 293K m2

_ d =A ¼ 0:075 kW
Ex
m2
In this example, the exergy transferred into the wall is either destroyed within the wall (spontaneous heat transfer) or
transferred out of the wall, where it is lost to the surroundings. Through this exercise, it becomes apparent that to minimize the
exergy destroyed, the temperature gradient between both surfaces needs to minimized so that the heat flow is small. For example,
if the heat flow were to be reduced to 0.15 kW m2 by adding thicker or higher insulative material, Ex_ d =A would be 0.037 kW m2
or effectively half. This analysis is relevant for many applications, including thermal energy storage systems where potential for
work from the thermal storage can be significantly augmented by improving the insulation of the boundary.

2.24.5 Results and Discussion

While adding insulation is a simple and effective solution to conserving energy and improving energy efficiency, there is a point at
which additional insulation will result in diminishing energy savings and no longer outweigh the monetary investment. This
section analyses the embodied energy of common insulation materials for building envelopes in Canada.

2.24.5.1 Lifecycle Energy of Insulating Materials


When selecting the type and quantity of insulation to install within a building envelope, it is important to consider all energy
aspects of the materials. In general practice today, a significant emphasis is placed on the amount of energy the building consumes
as it is occupied. The amount of insulation is optimized to reduce the amount of heat transfer into or out of the building
depending on the climate, while also commonly factoring in the cost of the material and the installation costs. What is often
overlooked during the design phase is the lifecycle energy of the material. Commonly referred to as the cradle-to-grave energy, this
value takes into account the amount of energy required for the production and installation of the insulation material, the amount
of energy consumed, or saved during the life of the product and the amount of energy required to dispose of the material at the
end of its useful life. A second value of importance is the lifecycle CO2 or CO2 equivalent, which measures the amount of CO2 or
CO2 equivalent that is released into the atmosphere during the production, installation, and life of the unit.
For both lifecycle calculations, the embodied energy or embodied CO2 of the material must be first calculated. The embodied
values are the amount of energy or CO2 that must be considered before factoring in the energy and CO2 released during the
operation of the building. What is included in calculating the embodied values varies depending on the source and the material
being examined, but for most sources and materials, the following items are included when calculating the amount of energy
consumed:

• the energy required to extract the raw materials required to manufacture the product;
• the energy required to manufacture the finished product;
• the energy required to transport the materials (raw material to factory to building site);
• the energy required to install the material within the building; and
• the energy required to properly dispose of the material at the end of its useful life.

In addition to these values, some studies also include the following items:

• the energy used to maintain the product through its life; and
• the energy used when repairing or replacing the product during the lifespan of the total building.

Although there is a variety of embodied energy and CO2 equivalent values available, a comprehensive database of building
materials has been assembled by Hammond and Jones [37]. Based on this database, the average embodied energy for the different
insulations examined was 46.84 MJ kg1. Table 2 below includes the embodied energy and embodied CO2 equivalent for some of
the most common insulating materials.
When looking at the building as a whole, the most common metric is to examine the lifecycle energy of the building. This takes
into account the total-embodied energy within the building, and the amount of energy consumed to operate the building.
Typically, only the energy to condition and operate the building (and not the total energy consumed within the building) are
considered in the lifecycle energy and CO2 values. For example, the amount of energy required for space heating in the building
would be included in the lifecycle energy, along with the amount of CO2 released to provide that heating, while the electricity to
run the computers in the building would not be included in the energy and CO2 lifecycle calculations. To calculate the lifecycle
energy, Elife, and CO2 values, eCO2 ;life , the embodied, heating, and cooling values must be calculated and summed. To do this,
782 Insulation Materials

Table 2 Embodied energy and carbon for common insulating


materials

Material Embodied energy Embodied carbon


(MJ kg1) (kg CO2eq kg1)

Cellulose 0.94–3.3 0
Fiberglass 28 1.35
Mineral wool 16.6 1.28
Rock wool 16.8 1.12
Expanded polystyrene 88.6 3.29
Polyurethane spray foam 101.5 4.26

Source: International Energy Agency. Key world energy statistics. Paris:


International Energy Agency; 2016.

Table 3 Lifecycle energy for 1 m2 wall section using expanded polystyrene for 10-, 25-, and 50-year periods

Thickness (cm) Annual space conditioning load (MJ) Embodied energy (MJ) Lifecycle energy (MJ)

10-Year life 25-Year life 50-Year life

5 124.4 111 1355 3222 6332


10 62.2 222 844 1777 3332
15 41.5 332 747 1369 2406
20 31.1 443 754 1221 1998
25 24.9 554 803 1176 1798
30 20.7 665 872 1183 1701
35 17.8 775 953 1220 1664
40 15.6 886 1042 1275 1664
45 13.8 997 1135 1342 1688
50 12.4 1108 1232 1419 1730
55 11.3 1218 1331 1501 1784
60 10.4 1329 1433 1588 1847

Eq. (12) must be utilized.


X 24  3:6
Elife ¼ mi Eembodied;i þ ðHDD  COPheat þ CDD  COPcool Þ  Ueff Atotal   Ylife ð12Þ
1000

where mi is the mass of material i in kg, Eembodied,i is the embodied energy of material i in MJ kg1, HDD is the heating degree days,
CDD is cooling degree days, COPheat is the coefficient of performance of the heating system, COPcool is the coefficient of
performance of the cooling system, Ueff is the effective heat transfer coefficient of the building in W m2 K1, Atotal is the overall
area of the building envelope and Ylife is the expected lifespan of the building in years. This produces a final lifecycle energy
value in MJ.
Using this equation, the optimal level of insulation can be found. As additional insulation is added to the exterior of the
building, its reduction in total energy consumption decreases. As such, as additional insulation is added, the total-embodied
energy increases linearly; however, the resulting reduction in lifecycle energy is not linear and decreases as the thickness of the
insulation increases. Consequently, there is an optimal point that can be found to have the lowest total lifecycle energy.
When optimizing the thickness of the insulation installed within the building envelope, the location where the building will be
located and its expected life are both critical in determining the optimal insulation level from a lifecycle energy perspective.
Depending on the amount of insulation installed, the expected life of the building and the climate in which it is located,
the embodied energy in the insulating materials can actually be greater than the operating energy required to heat and cool the
building on an annual basis.
Let us consider a building located in a climate with 2000 combined HDD and CDD. To simplify the calculation, a 1 m2 section
of wall will be examined, the heating and cooling system has a COP of 1, and a single insulating material of increasing thickness
will be assumed. Table 3 and Figs. 13 and 14 show the resulting total lifecycle energy for this small section using EPS and fiberglass
insulation and assuming a 10-, 25-, and 50-year lifespan.
From this example, we can observe that for buildings with a shorter expected life, the embodied energy becomes the dominant
factor in the lifecycle energy calculation much quicker. For a 10-year life expectancy, the optimal insulation level is only 15 cm
(5.9 in.). However, as the expected life increases, the operating energy required for space heating and/or cooling the building
Insulation Materials 783

Embodied energy compared to operating energy


3500

3000 Embodied energy

2500
Energy (MJ)

2000

Space conditioning
1500 load − 25 years

1000

500
Space conditioning
load − 50 years
0
0 10 20 30 40 50 60 70
Thickness (cm)
Fig. 13 Comparison of embodied energy and operating energy for 25- and 50-year building life.

Comparison of lifecycle energy


7000

6000

5000
Total lifecycle energy (MJ)

4000

50 years
25 years
3000
10 years

2000

1000

0
0 10 20 30 40 50 60 70
Thickness (cm)
Fig. 14 Lifecycle energy for a building exposed to 2000 degree days and an expected life of 10, 25, and 50 years.

becomes more dominant, with a building expected to last 50 years having an optimal insulation level for lifecycle energy of 35 cm
(13.8 in.). As such, it is imperative that when designing a structure, the intended use of the building needs to be understood and
that its expected life and the location where it is being installed will determine the impact of the embodied energy that will have
on the optimization of the building envelope. To determine the impact on the optimal insulation levels, the number of degree
days that the building is exposed to in the example was varied from the base 2000 degree days up to 6000 degree days. The impact
of this change is shown in Fig. 15.
784 Insulation Materials

Comparison of lifecycle energy


20,000

18,000

16,000
2000 Degree
days
14,000
Total lifecycle energy (MJ)

12,000

10,000 4000 Degree


days
8000

6000
6000 Degree
4000 days

2000

0
0 10 20 30 40 50 60 70
Thickness (cm)
Fig. 15 Impact of degree days on the optimal level of insulation for the building envelope.

From this example, we can observe that as the number of degree days increases, the amount of insulation required to reach the
optimal level for lifecycle energy increases. This is because when considering a building with a very high number of heating or
CDD, it is more optimal from a pure energy perspective to continue to increase the insulation due to the disproportionately large
amount of energy loss through the building envelope in comparison to the total-embodied energy within the insulating materials.
For a 50-year period, if the building is exposed to 4000 degree days, an optimal insulation level of 55 cm (21.7 in.) is required,
while at 6000 degree days, the optimal level is greater than 60 cm (23.6 in.). As such, buildings typically located in northern,
heating dominated climates become an optimization exercise not in total lifecycle energy, but in cost of adding additional material
in comparison to the energy being saved.
In this example, a heating and cooling COP of 1 was used to illustrate the impact of location and expected life of the building
on lifecycle energy calculations. In reality, the COP value can vary from as low as 0.9 using an on-site natural gas boiler to 4 when
using a heat pump for heating and cooling. If the 25-year case is reexamined, varying the COP of the heating and cooling system, a
new set of results are plotted in Fig. 16.
From this example, we can observe that the higher the efficiency of the heating and cooling system, the greater the influence
that the embodied energy within the insulating materials has on the lifecycle energy of the building. When a heating and cooling
system provides a COP on an annual basis of 4, only 10 cm (4 in.) of insulation are required, compared to the 35 cm (13.8 in.)
required when a COP of 1 is present for the space conditioning systems.
While this previous example focused on examining the total lifecycle energy of a building, it is also important to look at the
total lifecycle CO2 emissions of a building as well. The influence that the embedded CO2 has on the overall lifecycle CO2 is
dependent not only on the heating and cooling load of the building, but also the fuel source being used to meet these loads. If
natural gas is being used as the primary heating source in a heating dominated location, or the electrical supply within the
jurisdiction where the building is located is predominately generated through the burning of fossil fuels (natural gas, coal, diesel,
etc.), the amount of CO2 released for space heating and cooling will be many magnitudes greater than the amount of CO2
embodied within the material used to increase the thickness of the wall insulation. As such, if the design objective is to minimize
the total lifecycle CO2, as much insulation should be used as is physically possible, while also considering cost and lifecycle energy
optimization. If the building’s heating and cooling loads are met using electricity, and the building is located in a location that has
a very high renewable energy penetration (most energy is generated using wind, solar, and hydro), at this point the embodied CO2
has a much larger impact on the total lifecycle CO2. To calculate the lifecycle CO2 of a building, Eq. (13) is used.

X
eCO2 ;life ¼ mi Eembodied CO2 ;i þ ðHDD  COPheat þ CDD  COPcool Þ  Ueff Atotal  24  ei  Ylife ð13Þ

where Eembodied CO2 ;i is the embodied CO2 of material i in kgCO2 kg1, ei is the carbon emission intensity of the heating and cooling
fuel in g CO2 kWh1. This produces a final lifecycle CO2 emissions, eCO2 ;life , value in kg.
Insulation Materials 785

Comparison of varying coefficient of performances (COPs) on lifecycle energy consumption


3500

3000

2500
Lifecycle energy (MJ)

2000
COP1
COP2
1500 COP3
COP4

1000

500

0
0 10 20 30 40 50 60 70
Thickness (cm)
Fig. 16 Impact of heating and cooling coefficient of performance (COP) on optimal insulation levels.

Using the equations for lifecycle energy and CO2 emissions provides an additional metric to examine the insulation being
installed within a building. Depending on the overall objective of the building design, the insulation level can be optimized for
either lifecycle energy or lifecycle CO2 emissions. These values can be used within the design constraints of the building and a cost
optimization exercise can be used to develop an optimal building envelope for any given building, based on design goals,
materials, expected life, and building location.

2.24.6 Case Studies and Examples

This section contains a thorough case study about the design, testing, and development of a high RSI-value building envelope for
the Team Ontario’s entry into the 2013 solar decathlon to show the importance of insulation in residential buildings and its
impact on energy consumption.

2.24.6.1 Designing a High Thermal Resistance Wall With Vacuum Insulation Panels – ECHO – Team Ontario’s Entry to the
2013 Solar Decathlon
The building envelope is critical in the design of high performance, energy efficient housing. As codes and voluntarily performance
standards (LEED, Passive House, Net-Zero) [10,11], the ability to design and construct a high thermally insulating wall is pivotal
to the success of these walls. Traditionally, the high R-values required for these projects have been obtained by simply increasing
the thickness of the wall assembly. This has led to double wall construction, with or without exterior board insulation and
interior insulation between wall constructions, leading to walls that commonly have a total thickness in excess of 300–400 mm
(12–16 in.) and insulating values of approximately RSI-7 m2 K W1 (R-40 h 1F ft2 BTU1). Although this provides a significant
energy improvement over a code built wall, these highly insulated walls are in excess of twice as thick as a traditional wall. In the
current construction and real estate environment, where house are built to the allowable extent of the desired piece of land, this
increase in wall thickness leads to a subsequent decrease in usable floor space. In a typical, single detached two-storey home, this
lost floor space could be the equivalent of adding an extra bedroom.
To help combat and start to reverse this trend of simply building thicker walls to increase thermal resistance, new, highly
insulated building materials need to be integrated into the residential sector. As such, one of the design goals for Team Ontario as
they designed and subsequently built their competition home for the 2013 U.S. Department of Energy Solar Decathlon was to
develop a building envelope assembly that achieves a minimum total effective thermal resistance if RSI-8 m2 K W1 (R-45 h 1F ft2
BTU1), while having a total thickness equal to or less than 300 mm (12 in.). To achieve this effective thermal resistance goal, VIPs
were selected due to their very high thermal resistance in comparison to their thin profile.
786 Insulation Materials

This project had a unique design statement, in which the house was being designed and built in the cold climate of Eastern
Ontario, but also needed to be shipped 10,000 km (6000 mi) round trip to the competition site in Southern California.
Additionally, the house had to be built in a modular fashion and be able to withstand the lifting and assembly process in both
Canada before the competition and at the competition site. As such, a higher level of structural rigidity was required when
compared to a site built, stick-framed house. This required that 38 mm by 152 mm (1.5 in. by 5.5 in.) lumber be used as the main
structural members, as opposed to the preferred 38 mm by 89 mm (2  4) lumber, which would have allowed a thinner wall
profile. Using this as the starting point, a wall was designed that incorporated VIPs to the outside of the structural members,
allowing for an almost continuous layer of insulation and significantly reducing the thermal bridging within the assembly.
Although VIPs have significant insulating benefits, they posed a substantial challenge in terms of how to successfully integrate
the panels into a residential wall assembly. These panels are very fragile, and as a result, must be protected from threats, such as
fasteners, improper handling, and other potential methods of puncture. As a result, the panels were placed in the center of the
assembly, maximizing the distance between the interior and exterior surfaces, reducing the probability of accidently puncture
caused by the homeowner making modifications to the house. A comprehensive installation plan was also developed, ensuring the
panels could be installed without incident, including material handling and storage practices, as well as the installation process.
Another design consideration was the impact of VIPs on the moisture transfer that occurs within a wall assembly. VIPs can
significantly influence the moisture transfer within a wall assembly, as a VIP layer can cause a second vapor barrier and move the
dew point of the wall. As the VIPs account for approximately 50% of the insulation, it was decided that the panels would be placed
on the outside of the main structural members, ensuring the dew point remains outboard of the structural elements throughout
the year, removing the potential for bulk moisture accumulation. This significantly reduced the possibility of rot or mold growth.
However, it did not eliminate it, as there was still the potential for moisture to diffuse into the wall, and become entrapped
between the VIPs and the traditional vapor barrier. To counter the possible moisture issues, no additional vapor barrier (typically
polyethylene sheets) was installed on the inside of the wall, and instead, 50 mm (2 in.) of spray foam was applied within the stud
cavities, allowing moisture a path out of the wall through the studs. As this is not traditionally the way walls dry, which typically
sees moisture migrating to the exterior, two – 100 mm (4 in.) bands of moisture permeable EPS were installed within the wall. In
addition to serving as drying paths, these bands also provided a location for mechanical and electrical penetrations, as well as a
plane to fasten the exterior supports to the main structural components.
Throughout the process, thermal modeling within THERM [37] was conducted to determine the overall effective thermal
resistance of the wall. As different iterations were proposed, quick simulations were conducted to determine their effective thermal
resistances, which were then inputted into a complete energy model in EnergyPlus [38] to determine the annual energy perfor-
mance. Through these iterations, a final wall section was proposed (Fig. 17).
During the initial design phase, a number of parameters were unknown and needed to be assumed. The biggest of these was the
thermal resistance of the VIPs being used. A wide variety of quoted values were available within literature and product specifi-
cations, however, it was unknown the true value of the panels being proposed. As a result, a thermal conductivity of 0.002 W m1
K1 was assumed in the initial modeling, and an overall effective thermal resistance of the wall assembly of 14.7 m2 K W1 (R-83 h
1F ft2 BTU1) was obtained. To remove many of the assumptions, experimental validation of the thermal models was done, first
through in situ testing, followed by laboratory testing in a guarded hot-box.

2.24.6.2 Test Results


In situ testing was first conducted on the proposed wall design, with a 2.4 m by 2.4 m (8 ft by 8 ft) exact replica built,
instrumented, and installed within an outdoor test facility at Carleton University in Ottawa, Canada. Measurements were taken for

Exterior
Graphite EPS

Vacuum insulation panels Rainscreen airspace


2×3 Strapping Wood siding

Plywood sheathing 2×6 Framing


Medium density spray foam Ultility airspace

Gypsum wall board


Interior
Fig. 17 Cross-section of the final wall design, including 20 mm (13/16 in.) of vacuum insulation panels (VIPs) placed between the interior
structural elements and an exterior 2×3 wall assembly for which allowed for cladding to be attached and provided a rain screen.
Insulation Materials 787

4 weeks in January and February, and measurements were taken at 1 min intervals. Each unique cross-section was instrumented to
measure the temperature profile through the wall assembly, as shown in Fig. 18 for the test period, and to determine the average
thermal resistance of the cross-section. Using the in situ data, and taking a weighted average of each cross-section, an overall
effective thermal resistance of the wall assembly was determined to be 10.4 m2 K W1 (R-59 h 1F ft2 BTU1).
This experimentally obtained thermal resistance was significantly lower than the originally predicted value, so the thermal
resistance of each layer was determined using the measured temperature profile and heat flux through the wall assemblies. These
were compared to the values used within the THERM modeling, and the actual thermal resistance of the VIPs for the 20 mm
(13/16 in.) layer was experimentally found to be 5 m2 K W1 (R-28 h 1F ft2 BTU1), compared to the initially assumed value of
8.3 m2 K W1 (R-47 h 1F ft2 BTU1) for the VIP layer. When this experimentally determined thermal resistance was used in an
updated model, the predicted effective thermal resistance was found to be 8.9 m2 K W1 (R-50 h 1F ft2 BTU1).
To validate the in situ test and THERM modeling results, another replica wall was constructed and installed within a guarded
hot-box at Carleton University, as shown in Fig. 19, allowing the wall to be experimentally evaluated at steady state conditions.

Team Ontario experimental wall data


Jan. 11, 2013 − Feb. 8, 2013

60 20

50
10
40

30 0

Heat flux (W m−2)


Temperature (°C)

20
−10
10
−20
0

−10 −30

−20
−40
−30

−40 −50
Jan-11

Jan-13

Jan-15

Jan-17

Jan-19

Jan-21

Jan-23

Jan-25

Jan-27

Jan-29

Jan-31

Feb-02

Feb-04

Feb-06

Feb-08
Interior Interior surface
Sprayfoam-Plywood Plywood-VIP
VIP-VIP VIP-EPS
Exterior surface Exterior
Fig. 18 Temperature profile through the wall assembly with vacuum insulation panels (VIPs) over the 4 weeks testing period in Winter 2013.

Fig. 19 Vacuum insulation panel (VIP) layer of the test wall being installed within the guarded hot-box surround panel for laboratory testing.
788 Insulation Materials

30

20

2SF PLY

Temperature (°C)
10
2 PLY VIP
0 2 VIP VIP
0 5 10 15 2 VIP EPS
−10 Climate side
Metering box
−20

−30
Time (h)
Fig. 20 Steady state temperature profile for a 15 hour test in the guarded hot-box. EPS, expanded polystyrene; VIP, vacumm insulation panel.

This wall specimen was instrumented to provide a temperature profile at the same cross-sections as the in situ testing, with the
temperature profile over the length of the test shown in Fig. 20. Instead of using a weighted average, the total heat transfer through
the wall was measured. Coupled with the measured interior and exterior temperatures, the overall effective thermal resistance
of the wall assembly was found to be 8.2 m2 K W1 (R-46.5 h 1F ft2 BTU1). This value was within the experimental uncertainty of
the test, validating both the in situ test results and updated thermal model in THERM.

2.24.6.3 Building Envelope Construction


Following the detailed analysis of the proposed wall assembly, this design was selected as the complete house. Although the
thermal performance was optimized and fit within the parameters, a number of construction challenges remained. A detailed
panel layout was designed based on the proposed geometry of house, including the size and location of windows, doors, and
required mechanical penetrations. This was required to maximize the VIP coverage in the wall assembly. In areas where the house
dimensions did not match the VIP dimensions, changes were proposed to the architects to change the location or size of the
feature, and in many cases, a compromise was found between the engineers and the architects. This demonstrated an important
aspect of designing with new materials. After the panel layout had been developed, the few remaining areas that did not have VIP
coverage were filled with EPS.
The building envelope was then constructed as per the designed specifications. Special care was required in the storage and
handling of the panels. This included keeping the panels in dry conditions, avoiding materials being stacked on top of them, and
away from tools that could potentially damage the panels. When installing the panels, it was imperative that the panels were not
dropped or left on the ground around the exterior of the building. To improve the building envelope performance, the seams
between the panels were taped, reducing potential air infiltration into the building. The installed panels on the exterior of the
structural members are shown in Fig. 21. Once the VIPs were installed on the exterior of the structural members, strapped EPS was
installed on the exterior of the VIPs (Fig. 22), allowing the cladding to be installed on the exterior of the EPS (Fig. 23). A picture of
the finished exterior is shown in Fig. 24.

2.24.6.4 Lessons Learned


This project successfully showed that VIPs could be utilized in a residential building. A building envelope was designed, tested, and
constructed with an overall effective thermal resistance of 8.2 m2 K W1 (R-46.5 h 1F ft2 BTU1) in a total thickness of 300 mm
(12 in.), which could be easily reduced if not for some of the unique design criteria required for a Solar Decathlon Project. With
current design and building practices, the design process when using VIPs is far more complex and involved when compared to the
design of a traditional house. As more demonstration projects are completed successfully, and the knowledge base is increased, this
process will become more streamlined, and consequently, reduce the cost of implementing VIPs into a residential building envelope.
Once successful, based on the outcomes of this project, the implementation of VIPs into residential buildings could allow for
the energy consumption required for space heating and cooling to be reduced by half, while keeping the same overall wall
thickness and therefore not compromising the interior floor space of the building.

2.24.6.5 Building Enclosure Simulation


Following the experimental analysis at steady state and in situ conditions, a 2D heat transfer computer simulation was conducted.
In the Team Ontario Solar Decathlon House, incorporating the VIPs to the exterior wall insulation of the building introduced a
nonhomogenous thermal resistance to the layer. The temperature deviations that are caused by the VIPs are difficult to visualize
with experimental measurements due to a lack of granularity. However, computer simulations are a useful tool to determine the
temperature gradients and effective RSI-value for multiple cross-sections and designs. For 2D simulations, when thermal resistance
Insulation Materials 789

Fig. 21 Vacuum insulation panel (VIP) layer being installed on the exterior of structural members.

Fig. 22 Strapped expanded polystyrene (EPS) exterior layer installed and ready for siding.

discontinuities exist in the vertical and horizontal directions of a single layer, multiple cross-sections need to be simulated, and a
weighted averaging technique based on the height ratios of the cross-sections should be used.
When integrating VIPs into building envelopes, installing them as a continuous layer, similar to foam board as an exterior
insulation, is not the most efficient method due to cost and protection the VIP from puncture. Recently, the VIPs have been encased
in rigid foam board prior to being fastened to the exterior of the building to simplify the installation method at the construction site.
Fig. 25 shows a panel with VIPs encased in XPS with the top removed and distinct spaces created for a thin, high performance wall,
and to avoid puncturing the VIP before completing construction of the building. However, this does create a nonhomogeneous layer
and requires simulation of multiple cross-sections to determine the effective RSI. A drawing of an exterior wall assembly using VIP
encased in XPS as the exterior insulation is shown in Fig. 26. The wall is built with 38 mm by 89 mm (2  4 lumber) wood studs on
406 mm (16 in.) center spacing with RSI-2.4 m2 K W1 (R-14 h 1F ft2 BTU1) batt insulation in the cavity between the studs.
790 Insulation Materials

Fig. 23 Pine siding installed to provide a rain screen with small gaps to allow for drying.

Fig. 24 Finished product at the competition site in California (photo credit: U.S. Department of Energy).

The simulation consists of modeling two separate cross-sections, one incorporating VIPs and one excluding VIPs. By using
interior and exterior boundary conditions, and the thermal properties of the building materials given from by the manufacturers or
found in the THERM database, simulation outputs valuable information regarding the assembly. The VIP and non-VIP cross-
sections were simulated to have RSI-values of 9.27 m2 K W1 (52.6 h 1F ft2 BTU1) and 4.48 m2 K W1 (25.4 h 1F ft2 BTU1),
respectively. The effective thermal resistance value of the either assembly is found by using the coverage ratios, where 88% of the
wall assembly is covered by VIPs and the remaining 12% is covered by 50 mm (2 in.) of XPS. Therefore, the effective thermal
resistance of the entire wall assembly is simulated to be 8.87 m2 K W1 (50.2 h 1F ft2 BTU1). This shows that when the VIPs are
encased in insulation as a protective measure, they can maintain a high level of thermal performance in the building envelope with
a minimal loss in effective RSI when compared to a complete VIP layer.
In addition to the RSI-value of a cross-section, a desirable output of a simulation is the isotherms of a model. Understanding
how the heat will permeate through the wall assembly is also of interest when the insulating layers contain discontinuities that will
act as thermal bridges. The isotherms present in the wall assembly from the exterior to the interior with the VIP-encased panel are
Insulation Materials 791

Fig. 25 A photo of a sheet of extruded polystyrene (XPS) with embedded vacuum insulation panels (VIPs) to augment the insulation. This was
installed within a guarded hot-box test facility to measure the effective RSI-value.

Stud Batt insulation Drywall Plywood

XPS VIP
Fig. 26 Composition of simulated exterior high RSI-value wall assembly. VIP, vacuum insulation panel; XPS, extruded polystyrene.

Interior

Exterior
Fig. 27 Schematic of a wall assembly evaluated with two-dimensional steady state without isotherms.

Interior
19.4
16.8
14.1
11.5
8.9 6.3

−17.3
Exterior
Fig. 28 Isotherms through the wall assembly. Note the thermal bridge that exists at the extruded polystyrene (XPS) of the composite insulation
panel.

shown in Fig. 27. Even without considering the edge effects caused by the lower thermal resistance along the perimeter of the VIPs,
a large temperature difference along the wood stud–gypsum board interface is caused by the XPS aligned with the wood studs, two
low thermal resistive materials compared to the VIP (Fig. 28).
792 Insulation Materials

2.24.7 Future Directions

With the increase in building efficient and high performance buildings, the use of high R-value materials has been growing. These
novel materials offer a large thermal resistance with a small thickness in order to highly insulate the building without compro-
mising the internal floor area of the home or the building’s footprint caused by increasing the thickness of the building envelope.
The materials that will be discussed can offer as high as 10 times the thermal resistance per unit thickness when compared to the
conventional insulation materials discussed previously [39].

2.24.7.1 Aerogels
Aerogels are very porous solid and lightweight materials that are composed of a silica dioxide gel and a liquid solvent. Through a
chemical process and drying the gel, the liquid pores are replaced with gas, providing the low density solid. Aerogels are an
effective insulator because the small gas-filled pores significantly limit the means of heat transfer. The internal structure limits the
air movement through the material, which reduces convection [40]. Silica is known as having a low thermal conductivity, and the
pores in the material cause many thermal breaks further improving the effective thermal conductivity.
Aerogels are often provided on a roll, and cover the surface of the building envelope. Compared to conventional insulations,
aerogels can provide 2–2.5 times the thermal resistance, which offer a significant reduction in insulation thickness required to
create a high performance envelope. Aerogels have the flexibility and versatility to be mechanically fastened to many facets of the
building, including the interior or exterior walls, windows, doors, attics, among many others. Due to their high compressive stress,
they can also be used as underfloor insulation when the thickness is a limitation [41]. In buildings, the aerogels offered can be
translucent or opaque, which further opens the possibility of applications, specifically large thermal improvements on windows
and skylights [40]. Unfortunately, a high cost per unit area limits its widespread implementation.

2.24.7.2 Vacuum Insulation


VIPs are a highly insulating material that has been implemented into many appliances, such as refrigerators and freezers, and the
transportation industry where refrigeration is required. VIPs consist of an evacuated, open porous core enclosure within a several
metallized laminate layers that provide a high thermal resistance per unit thickness [39]. The thermal resistance benefits are
provided by the lack of convective and conductive currents through the panel caused by the porous structure and vacuum induced
inside.
Some VIP manufacturers have claimed a thermal resistance as high as RSI-234.6 per meter (R-40 in.1) through the center of
the panel. However, the panel does not have a homogenous thermal resistance. Experimental studies have shown that a 32% drop
in thermal resistance may exist along the perimeter of VIPs and this reduction would vary between each panel from different
manufacturers [42]. The thermal resistance variations that exist along the perimeter are caused by the different sealing or folding
processes of the metallic envelope to contain the vacuum. Another factor would be the type of metallic envelope and core material
used to fabricate the VIPs.
As such, the VIPs contain an inherent thermal bridge along the perimeter of the panels. The perimeter has a significant increase
in thermal conductivity when compared to the center of the panel because of the manufacturing technique to seal the seams and
the conductivity of the metallized foil envelope [43]. The thermal bridge can be seen through infrared thermal images taken during
laboratory testing in Fig. 29. The panel edges are distinctly shown in the thermal images by the variation in colors related to the
surface temperature.
Even though there are substantial advantages of VIPs, many disadvantages and challenges exist before they can be implemented
into the building envelope as an insulator. Since the VIP’s low conductivity is based on the vacuum contained within the
enclosure, if they are punctured, or the enclosure is compromised, there is a large reduction in performance. In building con-
struction, there are many hazards, such as mechanical fasteners (e.g., screws, nails, etc.) and the physical handling of the panels,
could lead to puncture before the building envelope construction is complete [44].
Another effect of VIPs when installed inside the building envelope is their degradation in thermal performance over its lifespan.
The panels will experience a drop in RSI-value after fabrication due to increasing internal pressure and moisture accumulation
within the panel. While the VIPs create an impermeable vapor barrier within the building envelope, there can be vapor and air
diffusion into the core of the panel either through the metallic foil or through imperfections created during manufacturing or
installation [39]. Some studies have shown that moisture accumulation will have a greater effect on VIP degradation than
increased air pressure, and a response has been the addition of desiccants to the core material. The core material’s conductivity is
increased as moisture enters the foil and is trapped within the pores. This creates a thermal bridge for heat to travel.
After installation into a building enclosure, the VIPs are introduced the environment and moisture passing through the
envelope, either through diffusion or unintended leakage. The moisture is unavoidable after installation, and the effects should be
taken into consideration during the analysis for the lifetime of the envelope, whether it is 10, 15, 30, or 50 years [39,43]. However,
even though the panel will experience degradation over time, the VIPs will maintain an RSI-value greater than common building
insulations [45].
The high thermal performance of VIPs offer the ability to create thin, high RSI-value building envelopes that comply with
voluntary building standards without sacrificing the internal floor area or increasing the footprint of the building required by the
Insulation Materials 793

20.3
14.0 23.0

Fig. 29 Infrared thermal image of interior surface temperature of vacuum insulation panels (VIPs) installed in a building envelope. Reproduced
from Conley B, Cruickshank CA. Evaluation of thermal bridges in vacuum insulation panels assemblies through steady-state testing in a guarded
hot box. In: eSim 2016, Hamilton; 2016.

additional thickness when using common insulations. Even though drawbacks, such as performance degradation, thermal bridges
along the perimeter, fragility and constructability exist, they offer a high-upside solution to thermally inefficient buildings without
increasing wall thickness.

2.24.7.3 Gas-Filled Panels


Gas-filled panels (GFP) are similar to VIPs such that they have a metallic envelope. However, instead of being evacuated, they are
filled with a low conductivity gas, and since the enclosed gas remains at ambient pressure, the porous structure is not required. The
GFPs still contain a baffle structure inside in order to restrict the movement of gas and limit convection [46].
Since the filled gas has a strong effect on the effective thermal conductivity, varying the gas can change the thermal
performance. Some studies have suggested that GFPs would be able to obtain an effective thermal conductivity of
0.035 W m1 K1 for a 25-mm panel thickness when using air as the main gas, and 0.0106 W m1 K1 with krypton [46].
Typically, gases with higher molecular weight and mono-atomic gases will offer a smaller thermal conductivity, and should be
chosen as the fill gas.
The envelope foil must act as an effective gas barrier in two directions, such that it effectively keeps the low conductivity gas
inside and the air and moisture outside the panel. The GFP lifecycle strongly depends on the gas transmission rate through the
envelope foil as the effective thermal conductivity of the panel [46]. Similar to VIPs, the lifespan of the panels are dictated by the
diffusion through the foil.

2.24.8 Closing Remarks

In conclusion, the building enclosure significantly impacts the overall efficiency of a building, in terms of thermal performance
and indoor comfort. The enclosure can be improved through addition insulation applied to various locations in the building
either during the initial construction or during a renovation. Another method to improve the efficiency is to increase the
airtightness of the enclosure, by sealing the seams and imperfections that may exist. Finally, improving the enclosure will affect the
required capacity of the heating and cooling equipment in a building, further improving the amount of energy and costs required
to maintain indoor comfort.
The various types of insulation discussed in this chapter have a number of benefits, installation requirements, and challenges.
The building designer is responsible for selecting a building enclosure that meets the required performance. The insulation
materials or building assemblies may be experimentally tested or simulated to validate the thermal or moisture performance of the
design at specific climate, since climatic zones play a large role in the enclosure requirements. As building designs are trending
toward a lower energy consumption, the enclosure insulation becomes increasingly important. Adopting new and higher per-
forming insulating materials with careful enclosure design will be paramount to the future of buildings.
794 Insulation Materials

References

[1] International Energy Agency. Key world energy statistics. Paris: International Energy Agency; 2016.
[2] Office of Energy Analysis. U.S. Energy Information Administration. International energy outlook 2016. Washington, DC: U.S. Department of Energy; 2016.
[3] Natural Resources Canada – Office of Energy Efficiency. Energy use data handbook 1990 to 2013. Ottawa, ON: Natural Resources Canada; 2015.
[4] EuroStat. Final energy consumption, EU-28, 2014, European Commission; 2016.
[5] U.S. Department of Energy. Furnaces and boilers. Available from: http://energy.gov/energysaver/furnaces-and-boilers.
[6] International Code Council. 2012 International energy conservation code. Illinois: International Code Council Inc.; 2011.
[7] ASHRAE. Energy standard for buildings except low-rise residential. Atlanta, GA: ASHRAE; 2013.
[8] International Energy Agency. Energy efficiency requirements in building codes, energy efficiency policies for new buildings. Paris: International Energy Agency;
2008.
[9] National Research Council of Canada. National energy code of canada for buildings 2011. Available from: http://www.nrc-cnrc.gc.ca/eng/publications/codes_centre/
2011_national_energy_code_buildings.html; 2016.
[10] Passive House Institute US. Available from: http://www.phius.org; 2016.
[11] U.S. Green Buildiing Council. LEED. Available from: http://www.usgbc.org/leed; 2016.
[12] Green Building Initiative. Available from: http://www.thegbi.org/; 2016.
[13] U.S. Department of Energy. Zero energy ready home. Available from: http://energy.gov/eere/buildings/zero-energy-ready-home; 2016.
[14] Wilkinson J, Ueno K, De Rose D, Straube J, Fugler D. Understanding vapour permeance and condensation in wall assemblies. In: 11th Canadian conference on building
science and technology, Banff, Alberta; 2007.
[15] Lstiburek J. Understanding vapour barriers. Westford, MA: Building Science Corporation; 2005.
[16] Straube J. The influence of low-permeance vapour barriers on roof and wall performance. In: Proceedings of performance of whole buildings; 2011.
[17] Brown W, Chown G, Poirier G, Rousseau M. Designing exterior walls according to the rainscreen principle. Ottawa, ON: National Research Council of Canada;
1999.
[18] Lstiburek J. Moisture control for buildings. ASHRAE J 2002;44(2):36–41.
[19] Canada Mortgage and Housing Corporation. Energy efficiency building envelope retrofits for your house. Canada: Canada Mortgage and Housing Corporation;
2015.
[20] US Department of Energy. Blower door tests. Available from: http://energy.gov/energysaver/blower-door-tests; 2016.
[21] Canada Mortgage and Housing Corporation. Before you start your energy efficiency retrofit – The building envelope. Canada: Canada Mortgage and Housing Corporation;
2015.
[22] Natrual Resources Canada. Heat/energy recovery ventilators; 2016.
[23] US Department of Energy. Insulation. Available from: http://energy.gov/energysaver/insulation; 2016.
[24] US Department of Energy. Insulation materials. Available from: http://energy.gov/energysaver/insulation-materials; 2016.
[25] US Department of Energy. Types of insulation. Available from: http://energy.gov/energysaver/types-insulation; 2016.
[26] Mullens M, Arif M. Structural insulated panels: impact on the residential construction process. Constr Eng Manag 2006;132(7):786–94.
[27] Medina MA, King JB, Zhang M. On the heat transfer rate reduction of structural insulated panels (SIPs) outfitted with phase change materials (PCMs). Energy 2008;33
(4):667–78.
[28] Canada Mortgage and Housing Corporation. Insulating your house. Canada: Canada Mortgage and Housing Corporation; 2009.
[29] Natural Resources Canada. Keeping the heat in – chapter 7: walls, Natural Resources Canada; 2012.
[30] US Department of Energy. Where to insulate in a home. Available from: http://energy.gov/energysaver/where-insulate-home; 2016.
[31] Natural Resources Canada. Keeping the heat in – chapter 5: roofs and attics, Natural Resources Canada; 2012.
[32] Natural Resources Canada. Keeping the heat in – chapter 6: basement insulation, Natural Resources Canada; 2012.
[33] ASTM International. Standard test method for steady sate heat flux measurements and thermal transmission properties by means of a guarded hot plate apparatus. West
Conshohocken, PA: ASTM International; 2013.
[34] ASTM International. Standard test methods for water vapor transmission of materials. West Conshohocken, PA: ASTM International; 2015.
[35] ASTM International. Standard test method for thermal performance of building materials and envelope assesmblies by means of a hot box apparatus. West Conshohocken,
PA: ASTM International; 2011.
[36] ASTM International. Standard practice for determining thermal resistance of building envelope components for in-situ data. West Conshohocken, PA: ASTM International;
2007.
[37] Hammond G, Jones C. Inventory of carbon & energy (ICE). Available from: http://www.circularecology.com/embodied-energy-and-carbon-footprint-database.html; 2011
[accessed 01.08.17].
[38] THERM. Lawrence Berkeley National Laboratory (LBNL). Available from: http://windows.lbl.gov/software/therm/therm.html; 2016.
[39] National Renewable Energy Laboratory. Energyplus. Available from: https://energyplus.net/; 2016.
[40] Mukhopadhyaya P, Kumaran K, Ping F, Normandin N. Use of vacuum insulation panel in building envelope construction: advantages and challenges. In: 13th Canadian
conference on building science and technology, Winnipeg, MB; 2011.
[41] Baetens R, Jelle B, Gustavsen A. Aerogel insulation for building applications: a state-of-the-art review. Energy Build 2011;43(4):761–9.
[42] Fickler S, Milow B, Ratke L, Schnellenbach-Held M, Welsh T. Development of high performance aerogell concrete. In: 6th International building physics conference;
2015.
[43] Conley B, Cruickshank CA. Evaluation of thermal bridges in vacuum insulation panels assemblies through steady-state testing in a guarded hot box. In: eSim 2016,
Hamilton; 2016.
[44] Fricke J, Heinemann U, Ebert H. Vacuum insulation panels – from research to market. Vacuum 2008;82(7):680–90.
[45] Baetens R, Jelle BP, Thue JV, et al. Vacuum insulation panels for building applications: a review and beyond. Energy Build 2010;42):147–72.
[46] Kalnaes S, Jelle B. Vacuum Insulation panel products: a state-of-the-art review and future research pathways. Appl Energy 2014;116:355–75.

Further Reading
ASHRAE. ASHRAE handbook of fundamentals. Atlanta, GA: ASHRAE; 2013.
Jelle B. Traditional, state-of-the-art and future thermal building insulation materials and solutions – properties, requirements and possibilities. Energy Build 2011;43
(10):2549–63.
Modera MP, Persily AK. Airflow performance of building envelope components. West Conshohocken, PA: ASTM International; 1995.
Moran MJ, Shapiro HN, Boettner DD, Bailey MB. Fundamentals of engineering thermodynamics. vol. 7. Hoboken, NJ: John Wiley & Sons; 2010.
Staube J. High performance building enclosures. Sommerville, MA: Building Science Press; 2012.
Insulation Materials 795

Relevant Websites

https://ashrae.org/
American Society of Heating, Refrigeration and Air-Conditioning Engineers (ASHRAE).
https://buildingscience.com/
Building Science Corporation.
https://www.cmhc-schl.gc.ca
Canadian Housing and Mortgage Corporation (CMHC).
BuildingGreen.com
Green Building Information.
http://www.nrcan.gc.ca/home
Natural Resources Canada (NRCan).
http://www.nrcan.gc.ca/energy/offices-labs/office-energy-efficiency
NRCan Office of Energy Efficiency.
http://energy.gov/
United States Department of Energy (US DoE).

Вам также может понравиться