Neal Kruis
Neal Kruis
by
Nathanael J. F. Kruis
Doctor of Philosophy
2015
UMI Number: 3704744
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.
UMI 3704744
Published by ProQuest LLC (2015). Copyright in the Dissertation held by the Author.
Microform Edition © ProQuest LLC.
All rights reserved. This work is protected against
unauthorized copying under Title 17, United States Code
ProQuest LLC.
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346
This thesis entitled:
Development and Application of a Numerical Framework for Improving Building Foundation Heat
Transfer Calculations
written by Nathanael J. F. Kruis
has been approved for the Department of Civil, Environmental and Architectural Engineering
Michael Brandemuehl
Harihar Rajaram
Marcus Bianchi
Michael Deru
Date
The final copy of this thesis has been examined by the signatories, and we find that both the
content and the form meet acceptable presentation standards of scholarly work in the above
mentioned discipline.
iii
Development and Application of a Numerical Framework for Improving Building Foundation Heat
Transfer Calculations
Heat transfer from building foundations varies significantly in all three spatial dimensions
and has important dynamic effects at all timescales, from one hour to several years. With the ad-
ditional consideration of moisture transport, ground freezing, evapotranspiration, and other physi-
cal phenomena, the estimation of foundation heat transfer becomes increasingly sophisticated and
computationally intensive to the point where accuracy must be compromised for reasonable compu-
tation time. The tools currently available to calculate foundation heat transfer are often either too
limited in their capabilities to draw meaningful conclusions or too sophisticated to use in common
practices.
This work presents Kiva, a new foundation heat transfer computational framework. Kiva pro-
vides a flexible environment for testing different numerical schemes, initialization methods, spatial
and temporal discretizations, and geometric approximations. Comparisons within this framework
provide insight into the balance of computation speed and accuracy relative to highly detailed
reference solutions.
The accuracy and computational performance of six finite difference numerical schemes are
verified against established IEA BESTEST test cases for slab-on-grade heat conduction. Of the
schemes tested, the Alternating Direction Implicit (ADI) scheme demonstrates the best balance
Kiva features four approaches of initializing soil temperatures for an annual simulation. A
new accelerated initialization approach is shown to significantly reduce the required years of pres-
imulation.
iv
context further improve computational performance. A new approximation called the boundary
layer adjustment method is shown to improve accuracy over other established methods with a
negligible increase in computation time. This method accounts for the reduced heat transfer from
concave foundation shapes, which has not been adequately addressed to date. Within the Kiva
framework, three-dimensional heat transfer that can require several days to simulate is approxi-
mated in two-dimensions in a matter of seconds while maintaining a mean absolute deviation within
3%.
v
Acknowledgements
This work was funded in part by an American Society of Heating Refrigeration and Air
Simulations were performed on the Janus supercomputer, which is supported by the National
This work would not be possible without the help and guidance of my advisor Dr. Moncef
Krarti. You always encouraged me to explore deeper and develop a broader understanding.
I would like to thank Peter Ellis, President of Big Ladder Software, for allocating time and
resources to enable this work. Also, thanks to Joe Huang President of White Box Technologies for
A special thanks to Ben Rohrs, for his friendship, support, and accountability, and Becky
To Dr. Mary Poel and Dr. Rick Kruis, without whom I would not have made it where I have
(let alone existed). Thank you, mom and dad, for your love, support, and the opportunities you
Most importantly I want to thank my wife, Cameron Frothingham Kruis, for the support
and encouragement she provided throughout working on this thesis. You demonstrated confidence
Chapter
1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Literature Review 5
2.1 Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3.5 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.5.4 eQUEST . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
2.5.6 TRNSYS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3 Tool Development 50
3.1 Kiva . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.4.2 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.5 Outputs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4 Solution Verification 76
Bibliography 149
Appendix
Table
4.2 Mesh detail parameters selected by each tool for IEA BESTEST analyses . . . . . . 83
4.4 BESTEST steady-state test case results: Slab heat loss [W] . . . . . . . . . . . . . . 88
4.7 BESTEST GC40a initialization method percent difference from “ideal” for seven
years of simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.8 BESTEST harmonic (unsteady) test case results: Annual slab heat loss [MWh] . . . 96
4.9 BESTEST GC40 harmonic parameters for each solution (shading indicates deviation
4.11 BESTEST harmonic (unsteady) test case results: Simulation wall time [s] . . . . . . 101
5.4 Average simulation wall time for two-dimensional approximation methods . . . . . . 140
6.1 Recommended methods and numerical parameters for foundation heat transfer . . . 145
Figures
Figure
2.4 Illustration of ITPE zones (1, 2, and 3) for a basement foundation (adapted from
2.5 Concentric heat flow paths (adapted from Latta and Boileau [79]) . . . . . . . . . . 49
3.3 Plan view illustration of far-field boundary definition in three-dimensions and foun-
3.6 Boundary cells and resulting zero-thickness cells. Black nodes = normal solid cells,
white nodes = boundary cells, gray nodes = zero-thickness cells, hollow nodes = air
cells. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.7 Example finite difference scheme (black nodes are known values, white nodes are
unknown values) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.13 Typical implicit matrix sparsity pattern for three dimensions (up to seven unknowns
per equation) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.14 ADI matrix sparsity pattern for three dimensions (up to three unknowns per equation) 71
4.2 BESTEST GC10a slab heat loss sensitivity to mesh detail parameters . . . . . . . . 82
4.3 BESTEST GC10a simulation wall time sensitivity to mesh detail parameters . . . . 82
4.6 BESTEST GC10a slab heat loss and simulation wall time sensitivity to linear solver
tolerance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.8 BESTEST steady-state test case simulation wall times (note logarithmic y-axis) . . 88
4.9 BESTEST GC40a comparison of initialization methods: First year slab heat loss . 92
4.10 BESTEST GC40a comparison of initialization methods: Annual slab heat loss for
4.11 BESTEST GC40a comparison of first year slab heat loss for traditional and accel-
4.12 BESTEST harmonic (unsteady) test case annual slab heat loss (results for GC50b
4.14 BESTEST harmonic (unsteady) test case simulation wall times . . . . . . . . . . . . 100
xv
5.4 Dimensions of a rounded rectangle with equivalent area and perimeter . . . . . . . . 110
5.6 Plan view cross-section at grade height of heat flux near a concave slab corner (ex-
5.8 Reduction in heat loss between two parallel concave edges over a range of area-to-
5.9 Reduction in heat loss between two parallel concave edges over a range of soil con-
5.10 Reduction in heat loss between two parallel concave edges over a range of insulation
5.11 Illustration of a heat flux boundary layer profile for a slab with exterior horizontal
insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.12 Example boundary layer profile for a slab with exterior horizontal insulation (ksoil =
5.13 Shapes with equal area and perimeter but different number of concave corners . . . . 118
5.14 Reduction in heat loss due to corners over a range of area-to-perimeter ratios (ksoil =
5.15 Reduction in heat loss due to corners over a range of soil conductivity (A/P = 4.6
5.16 Reduction in heat loss due to corners over a range of insulation R-values (ksoil = 1.0
5.17 Illustration of corner bisector and unaffected edge boundary profile cross-sections . . 121
5.22 Annual outdoor dry-bulb temperatures from weather data files for selected climates.
Alaska . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
5.26 Shapes simulated in three dimensions for approximation method testing. Each plot
5.27 Comparison of heat flux from a range of uninsulated foundation shapes . . . . . . . 135
5.28 Heat flux reduction relative to uninsulated foundations in mixed climate for a range
5.29 Comparison of annual total heat transfer between three-dimensions and two-dimensional
5.30 Comparison of annual total heat transfer between three-dimensions and two-dimensional
5.31 Comparison of annual total heat transfer between three-dimensions and two-dimensional
5.34 Deviation from three-dimensional annual heat transfer vs. simulation wall time for
B.1 Percent difference in annual heat transfer between three-dimensional square founda-
B.3 Percent difference in annual heat transfer between three-dimensional narrow-H foun-
B.4 Percent difference in annual heat transfer between three-dimensional wide-H foun-
C.1 Comparison of heat flux from a range of uninsulated foundation shapes in the cold
climate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
C.2 Heat flux reduction relative to uninsulated foundations in mixed climate for a range
C.3 Percent difference in annual heat transfer between three-dimensional square founda-
C.5 Percent difference in annual heat transfer between three-dimensional narrow-H foun-
C.6 Percent difference in annual heat transfer between three-dimensional wide-H foun-
D.1 Comparison of heat flux from a range of uninsulated foundation shapes in the hot
climate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
D.2 Heat flux reduction relative to uninsulated foundations in mixed climate for a range
D.3 Percent difference in annual heat transfer between three-dimensional square founda-
D.5 Percent difference in annual heat transfer between three-dimensional narrow-H foun-
D.6 Percent difference in annual heat transfer between three-dimensional wide-H foun-
E.1 Comparison of annual minimum heat transfer between three-dimensions and two-
E.2 Comparison of annual maximum heat transfer between three-dimensions and two-
E.3 Comparison of annual minimum heat transfer between three-dimensions and two-
E.4 Comparison of annual maximum heat transfer between three-dimensions and two-
E.5 Comparison of annual minimum heat transfer between three-dimensions and two-
E.6 Comparison of annual maximum heat transfer between three-dimensions and two-
Introduction
1.1 Motivation
Building energy modeling has become an integral part of the design process for both new
building projects and existing building retrofit projects. The tools used to simulate building per-
formance allow designers to make informed, economic decisions on how to best invest in energy
Simulation tools can aid in the design of building foundations, just as they serve to inform
the design of other components in buildings (walls, windows, lighting, HVAC equipment, etc.).
To accomplish this, simulation tools must be capable of estimating the heat loss through building
Heat loss from building foundations varies significantly in all three spatial dimensions and
has important dynamic effects at all timescales, from one hour to several years. With the additional
consideration of moisture transport, ground freezing, evapotranspiration, and other physical phe-
nomena, the estimation of foundation heat transfer becomes increasingly sophisticated and compu-
tationally intensive to the point where accuracy must be compromised for reasonable computation
time.
problem entirely because the overall magnitude of foundation heat fluxes is considered to be small
relative to the other thermal loads in a building. However, as buildings become progressively more
efficient, foundation heat loss begins to represent a larger portion of the total building load, thus
2
increasing the importance of accuracy in the calculations. It is estimated that heat transfer through
foundations can account for 30%–50% of the total heating load in certain buildings [39].
Unfortunately, the tools currently available to calculate foundation heat transfer are often
either too simplistic and limited in their capabilities to draw meaningful conclusions from the results
to ignore the problem altogether. Many practitioners use an adiabatic boundary condition for
ground-coupled building surfaces in their models. As a result, the foundation of the building
remains underdesigned relative to the rest of the energy-related features captured in the building
simulation.
Another practice is to avoid simulating the heat transfer from the building foundation by
simply overdesigning the foundation insulation. The Passivhaus strategy can specify up to 25.4
cm (10 in) of rigid foam insulation beneath the entire slab foundation [83]. Although some level
of design does go into the Passivhaus specification, the level of detail and accuracy involved in the
In one extreme, there is the possibility of unrealized energy savings, and in the other, the
Heat transfer from building foundations remains a large uncertainty in the estimation of
building energy use. An informal survey of energy modeling professionals indicates that the current
practice with regard to modeling foundation heat transfer is either oversimplified or ignored entirely
leading to low confidence in both simulation results and foundation design decisions. Many modelers
are under the impression that foundation heat transfer is negligible and, therefore, justly ignored.
However, when seeking a reference for such a claim, the author has found no such evidence.
The research presented in this thesis holds the following premise: Heat transfer through
building foundations is not always negligible and, therefore, warrants the careful consideration of
1.2 Objectives
The primary hypothesis of this research is as follows: A foundation heat transfer simulation
tool can be developed to balance capability, complexity, computational performance, and algorith-
mic confidence to provide building designers with a tool that will allow them to make informed
The list below states the research questions that are answered within the content of this
thesis:
(1) What algorithmic assumptions and simplifications can reduce computation time while re-
(2) What are the accuracy and speed implications of these assumptions and simplifications?
(1) Develop a detailed numerical truth standard as a benchmark for all subsequent assumptions
and simplifications.
efficient calculations of building foundation heat transfer within the context of whole-
Chapter 2: Literature Review: A review of the foundation heat transfer methods and tools
Chapter 3: Tool Development: A description of the tool developed for the work in this thesis.
4
Chapter 4: Solution Verification: Verification of the six solution methods implemented in said
tool. This chapter also investigates solution sensitivity to selected numeric parameters and
the efficiency of several initialization techniques within the context of a well-defined test
specification.
Chapter 6: Conclusions and Future Work: Concluding statements and suggestions for future
work.
Chapter 2
Literature Review
This chapter contains four sections, each describing a different aspect of foundation heat
transfer simulations:
(1) Terminology
A final section provides a comparison of prominent methods and tools found in both the
2.1 Terminology
For the sake of consistency, standard terminology used throughout this thesis is defined below.
These definitions are adapted from Fritzson [45] and Judkoff and Neymark [60]:
System: A system is an object or collection of objects whose properties are studied. The system
under consideration in this thesis is the physical construction of the building foundation
Model: A model is an approximate representation of a system. The model is the set of input
and output data values relevant to the study of the system. In this case, the model is the
Simulation: Simulation is the process of applying controlled inputs to the model and observing
Solution Method: The solution method describes the algorithmic translation between model in-
put and model output. For a foundation and ground system, the solution method generally
refers to how the conduction-diffusion partial differential equations are solved including the
Tool: A software implementation of a solution method (or multiple methods) capable of performing
The following sections provide a review of modeling capabilities, both available in and missing
from current foundation heat transfer tools. This informs which capabilities are important for the
development of new tools and the challenges associated with implementing them.
Foundation design capabilities are the tool’s ability to define the physical construction details
Figure 2.1 shows a range of foundation types found throughout the architectural domain.
Most tools can model the three standard foundation types (Figures 2.1(a)–2.1(c)). Even among
these common foundation types, tools are often limited to specific foundation depths. Few tools
(d) Walk-out basement (e) Split-level (f) Sub-basement (g) Underground structure
A few case studies in literature investigate heat loss from underground structures [23, 84] and
slabs with a stepped-exterior grade [22], but these capabilities are rarely integrated into publicly
available tools.
Accommodating many of these additional foundation types requires a more detailed descrip-
For a given foundation type, a continuous range of potential foundation designs exists. Figure
2.2 shows examples of typical insulation designs for foundations. While Figure 2.2 represents only
a small subset of the possible insulation designs, the full parameter space includes combinations of
• The location of the insulation (whole slab, interior horizontal, interior vertical, slab gap,
interior vertical, in-wall, exterior vertical, exterior horizontal, and all combinations thereof)
• Insulation dimensions (depth of vertical insulation, width and depth of horizontal insula-
tion)
(a) Uninsulated (b) Whole-slab insulation (c) Interior and slab gap in-
sulation
(d) Interior half-wall insula- (e) Interior wall insulation (f) Interior wall and floor in-
tion sulation
(g) Insulated concrete forms (h) Exterior wall insulation (i) Exterior vertical and hor-
izontal insulation
Many simulation tools only model a discrete set of insulation designs either because they do
not generalize a meshing strategy for a continuous range of dimensions or because they are based
Relatively few researchers have investigated the impact of foundation shape on heat loss.
Many of the tools used today apply one-dimensional or two-dimensional approximations based on
the area and perimeter of the foundations. Few tools have the capability to simulate heat transfer
range of foundation shapes in three dimensions is necessary to justify approximations with fewer
Common simplifications for foundation heat transfer calculations only account for heat trans-
fer from the surfaces of a single thermal zone, neglecting heat transfer between thermal zones
Capturing these effects requires an explicit three-dimensional model with distinct boundary
Few examples of foundation heat transfer between multiple interior thermal zones exist in the
literature. The temperature difference between interior thermal zones is often not great enough to
cause significant heat transfer. Xie et al. [119] and McDowell et al. [86] describe tools with multiple
interior zone boundaries; however, they do not quantify the magnitude of heat transfer between
the zones.
10
nificant concern. The placement of insulation to eliminate potential heat flow paths through the
bridging is common along the perimeter—at the interface of the above-grade foundation wall with
the exterior environment. Many studies investigate the optimal configuration of insulation on both
exterior and interior portions of the foundation wall to minimize heat loss through thermal bridges
such as the edge of the slab and sill plate (Figure 2.3).
The construction elements near the perimeter, which lead to thermal bridges, require a high-
resolution description of the foundation details. Such high resolutions are time consuming to model
and simulate. Many of the computational methods employed in foundation heat transfer models
do not describe the geometry at a resolution that adequately represents thermal bridging.
Wang [115] found that neglecting the heat transfer through the sill plate interface between
the above-grade and below-grade walls can result in a 46% underprediction in the basement wall
11
heat transfer. He also found that exterior foundation wall insulation can have a 25%–32% difference
A number of other potential thermal bridges exist along foundation perimeters. Rock [99]
explored the effect of expansion joints on thermal bridging through the slab edge.
Most simulation algorithms assume one-dimensional heat transfer for above-grade walls.
However, select studies show that the two-dimensional effects near the interface of above-grade
walls and the foundation can have a significant impact on the overall building heat balance [67].
There are a number of systems that are designed to exchange heat with the ground near
the building. These systems are rarely considered when calculating the heat transfer through the
Inclusion of such systems in the foundation heat transfer model requires an iterative solution
to ensure that the corresponding heat fluxes across the boundary between the system and the
One of the largest challenges in creating a ground domain model is establishing the thermal
properties of the soil. The thermal properties at any location within the ground domain are largely
• Quartz
• Clay minerals
• Organic matter
• Water
• Ice
• Air
The relative fraction of each of these components can vary significantly throughout the ground
domain. Because these variations are both difficult to measure and burdensome to describe within
a model, a majority of the work found in literature assumes that soil properties are spatially
homogeneous.
A few authors have explored the impact of spatial variation on foundation heat transfer calcu-
lations. Krarti [65] explored the effect of lateral variation in soil types and Rees et al. [96] explored
the effect of vertical variation of moisture content based on a static one-dimensional moisture pro-
file. Deru [39] demonstrated that spatial variations of moisture content can be approximated using
The moisture distribution in soils is not static but instead is driven by determinable forces.
Simulating the actual distribution of moisture throughout the ground domain requires calculation
Few authors have explored the impact of both heat and moisture balances on foundation
heat transfer. However, two notable contributions, Deru [39] and Janssen [56], have developed
solutions incorporating a moisture balance that drives spatial and temporal changes to the soil
thermal properties.
calculations have found that the heat-transfer-only calculation underpredicts foundation heat trans-
When simulating the effect of rainfall, authors have shown disparate results. Though most
published results range between 0.1% and 10% difference from heat-transfer-only calculations [39,
52, 56], dos Santos and Mendes [42] show up to a 100% difference.
An important aspect of structural design for building foundations is the construction of the
foundation footer below the frost line. Though it is a common consideration in structural design,
the impact of ground freezing is rarely considered in building energy analysis. Freezing provides the
building with additional thermal storage in the form of latent energy and influences the apparent
thermal conductivity of the soil. Ice within soil voids also impedes moisture transfer.
A number of authors have discussed heat and mass transfer in frozen soils, though the theory
has primarily been applied to frost heaving and structural design [61, 81]. Few foundation heat
transfer tools calculate freezing, and the overall impact of freezing has not yet been thoroughly
quantified.
14
Interior convection is usually assumed to be buoyancy driven and is thus a function of the
temperature difference between the surface and the nearby air. Though several correlations exist
for defining natural convection coefficients, the correlations developed by Walton [113] are widely
• β is the tilt of the surface (from 0 at horizontal facing up, to π at horizontal facing down)
[radians],
Unstable heat flow refers to situations where the heat flow is upward and enhances buoyancy—the
surface is tilted up and is hotter than the ambient air, or the surface is tilted down and is colder
than the ambient air. Stable heat flow refers to situations where the heat flow is downward.
exact solution requires an iterative approach. Some tools calculate the convection coefficient using
15
the surface temperature from a previous timestep or set the convection coefficient to a constant
value.
The ambient air temperature is also determined by a heat balance performed on the adjacent
thermal zone. For the case of conditioned spaces, it is acceptable to assume a constant space
temperature; however, in cases where the space is unconditioned or the HVAC system is not able
to meet the set-point, the resulting ambient and surface temperatures must be calculated using an
iterative approach.
There are relatively few existing models that incorporate a direct coupling between the ground
Long-wave radiation is exchanged among the interior surfaces of a room based on the temper-
atures, emissivities of the surfaces, and the view factors between the surfaces. Calculating radiation
exchange becomes increasingly expensive with more surfaces, as it requires calculating all the sur-
face temperatures simultaneously in conjunction with the convective and conductive heat transfer
calculations. The EnergyPlus simulation engine [111] performs these calculations using an algebraic
method called the script-F method described by Hottel and Sarofim [53]. Cogil [30] developed a
tool for calculating heat transfer from basements, which used a similar algebraic method using
surface radiosities. Alternative algebraic methods are described by Clark and Korybalski [28].
Most tools do not solve the set of algebraic equations required for a complete radiation heat
balance. A common simplification is to treat all radiation from a surface as though it were only
q̇rad,lw = εs σ Ts4 − Tair
4
(2.2)
hrad = εs σ Ts2 + Tair
2
[Ts + Tair ] , (2.3)
where
16
Like the natural convection coefficient, the radiative coefficient is a non-linear function of the
surface temperature. To calculate radiation within the context of a linear system of equations,
the coefficient is calculated using the previous time-step’s surface temperature at the loss of some
accuracy.
Many tools, for example DOE-2 [80], are not capable of simulating dynamic coefficients and
use a single constant coefficient to represent both convection and radiation from a surface.
Interior foundation surfaces are sometimes exposed to short-wave solar radiation transmitted
through windows. Passive solar buildings are designed and oriented specifically to expose slab
foundation surfaces to incoming solar radiation to store heat. Most whole-building simulation tools
lack the coupling between the foundation heat transfer and interior short-wave radiation to capture
The effect of short-wave radiation incident on interior surfaces is calculated using a constant
heat flux. However, calculating the effect of transmitted solar radiation on foundation surfaces
requires integration with a more comprehensive building simulation tool to capture the effects of
is also a natural buoyancy-driven component. Numerous convection correlations appear in the lit-
erature with varying complexity. EnergyPlus combines several correlations to account for a wide
range of conditions into the somewhat misleadingly termed the “DOE-2 Model.” The algorithm
combines the MoWiTT correlations developed by Yazdanian and Klems [120] to account for wind
speed and wind direction, the natural convection algorithm described by Walton [113], some ad-
justments made by Booten et al. [18] to correct for the height of the surface, and an addition from
where
• hconv,smooth is the convection coefficient for a perfectly smooth surface [W/m2 -K] defined
by Equation 2.5:
hconv,smooth = h2n + h2f , (2.5)
where hf is defined by
hf = av b , (2.6)
where
• a is a correlation constant of
• b is a correlation constant of
The roughness factor described in Equation 2.4 is a qualitatively defined factor based on example
No references were found to provide surface roughness factors for roughnesses beyond the
range shown here. As none of the example surfaces describe typical ground coverings, the roughness
factor will have to be extrapolated to account for soils, asphalts, grasses, and other vegetation. As
a crude approximation, the roughness factors of other surfaces beyond the examples given here are
estimated by
Rf = 1 + Rt /Rn , (2.7)
where
The resulting estimated value of Rt for each of the examples is also shown in Table 2.1.
Wind speed values are typically measured at a weather station in an open area several meters
above any nearby obstructions and are not appropriate for calculating forced convection near the
ground adjacent to a building. The wind speed, v, in Equation 2.6 is often adjusted from the
weather station measurements provided in standard weather data sets using a procedure from the
αws αlocal
δws zveg
vground = vws , (2.8)
zws δlocal
where
• δws and δlocal are the boundary layer thicknesses [m] (Table 2.2) at the weather station and
• zws and zveg are the heights [m] of the weather station and the local vegetation, respectively,
and
• αws and αlocal are the terrain exponents (Table 2.2) at the weather station and the local
building, respectively.
Long-wave radiation is exchanged among exterior building surfaces, the ground surface, the
outdoor air, and the sky. The exterior long-wave radiation exchange calculation requires that each
The air temperature is often provided by weather data for the building location, and the
ground surface temperatures are calculated through a heat balance. The ground and building
surface temperatures are rarely, if ever, solved simultaneously using algebraic methods similar to
those described for interior long-wave radiation. Instead, long-wave radiation from exterior building
surfaces is calculated assuming that the ground is at the same temperature as the air, and long-
wave radiation from the ground is calculated assuming that the ground does not have a view-factor
Sky temperature is often correlated to information available in standard weather data sets.
The correlation used in both EnergyPlus and DOE-2 is shown in Equation 2.9:
1/4
Tsky = εsky Tair , (2.9)
where
Tdewpoint
εsky = 0.787 + 0.764 ln 1 + 0.0224fc − 0.0035fc2 + 0.00028fc3 , (2.10)
T air
where
where
• Fsky is the view factor from the surface to the sky [0–1],
• fs is the fraction defined by Equation 2.12, which distinguishes radiation to the sky from
• Fground is the view factor from the surface to the ground [0–1], and
and
where β is the tilt of the surface in radians (from 0 at horizontal facing up, to π at horizontal facing
down).
The radiative flux can be simplified when substituting the definition of sky temperature and
assuming that the ground temperature is the same as the air temperature:
q̇rad,lw =εs σ Ts4 − Fsky fs εsky Tair
4 4
+ Fsky [1 − fs ] Tair 4
+ [1 − Fsky ] Tair
=εs σ Ts4 − Tair
4
F∗ , (2.15)
22
hrad = εs σ Ts2 + Tair
2
F ∗ 1/2 Ts + Tair F ∗ 1/4 (2.16)
while using the surface temperature from the previous time-step, the calculation of exterior radia-
q̇rad,lw = hrad Ts − Tair F ∗ 1/4 . (2.17)
Short-wave solar radiation is calculated as a heat flux on the ground surface. Many tools
will simulate solar incidence on the ground, but often the solar absorptivity of the ground surface
is assumed to be homogeneous and does not account for shading from building surfaces or ground
cover.
A couple authors have explicitly simulated the effect of shading on the ground in their tools.
Bahnfleth and Pedersen [11] showed that shade from the building can increase the mean heat loss
from a slab by 27.4% in Phoenix, Arizona and 14% in Medford, Oregon. dos Santos and Mendes [41]
estimated that shading can cause up to a 30◦ C range of variation in the ground surface temperatures
2.2.3.7 Evapotranspiration
Evapotranspiration describes the evaporation of water from ground and plant surfaces. Many
models do not incorporate evapotranspiration despite its cooling effect on the exterior ground
surface boundary.
The rate of evaporative heat loss from the soil depends on the vapor pressure at the soil’s
surface relative to the vapor pressure of the air. In the absence of moisture transfer calculations, the
evaporative effect can only be approximated based on an assumption of the wet-bulb temperature
at the soil’s surface. The actual evapotranspiration rate is between zero and a maximum where the
23
soil surface wet-bulb temperature is the same as the wet-bulb temperature of the air. When the
soil is saturated with moisture, it is safe to assume maximum evapotranspiration. However, for dry
soil the low surface wet-bulb temperatures can lead to unrealistic results.
Using the maximum potential evapotranspiration, Bahnfleth and Pedersen [11] demonstrated
that neglecting the effect can reduce the annual average heat loss by up to 170% (i.e., a heat gain)
32.3% in Medford, Oregon. Deru [39] also showed an 11.2% difference in Fort Collins, Colorado.
Snow cover has several effects on the exterior ground surface boundary condition:
• Snow acts as an insulating layer on the ground. Gilpin and Wong [47] refer to this as the
heat-valve effect.
One significant challenge in simulating snow cover is that snow amounts are not typically
provided in standard weather data. While there are data fields available for snow depth, they are
The far-field boundary represents the distance from the building where there is no longer heat
transfer in the horizontal directions. This may be at a distance sufficiently far from the building
or, alternatively, along a plane of symmetry (e.g., half-way between two similar buildings).
calculated, one-dimensional temperature profile. If the boundary is sufficiently far away from the
24
building, then both approaches should produce the same result. However, the zero-flux boundary
The deep-ground boundary is usually defined as either the distance to the water table or the
distance to a point in the ground deep enough that it is not affected by the annual temperature
If the water table is shallower than the unaffected ground depth, then a constant temperature
condition is applied (usually the water table temperature is approximated as the annual average
outdoor dry-bulb temperature for the location). In a case where the water table is below the
unaffected ground depth, a zero-heat-flux condition is applied, although this is also sometimes
approximated by using a constant temperature condition equal to the annual average outdoor
dry-bulb temperature.
Bahnfleth and Pedersen [11] found that the type of boundary condition (zero-heat-flux versus
constant temperature) only accounts for a 1.3–3.5% change in average annual heat loss. Contrarily,
the depth of the boundary was found to have a fairly significant impact on results. Between depths
Several authors have explored the impact of water table depth on foundation heat transfer
[37, 11, 96, 39]; however, no tool accounts for the annual variation in water table depth observed
Some foundation heat transfer tools are used in the context of a whole-building energy sim-
ulation engine. Current whole-building simulation tools require tens of seconds to several hours
to simulate a single building’s energy use, depending on the complexity of the building model and
the solution methodologies employed. As foundation heat transfer represents only a fraction of the
energy load on the building, the calculations should represent a relatively similar fraction of the
25
overall computation time and not several orders of magnitude longer, which is the case for many
To minimize the computation time devoted to foundation heat transfer, many whole-building
simulation tools employ simplistic methods with limited capability based on preprocessed or re-
Most whole-building simulation tools are not designed for multidimensional heat transfer cal-
culations. Often there is an artificial boundary created between the foundation heat transfer calcu-
model.
The TRNSYS (TRaNsient SYtems Simulation) software program [104] has a very detailed
ground heat transfer library, which can be coupled to a whole-building energy simulation. The
heat balance between the thermal zone and the foundation requires a computationally expensive
iterative solution.
Decoupling is not without its advantages. Al-Anzi and Krarti [6] developed a method that
used decoupling as an opportunity to employ separate solution methods that have advantages at
different time scales. In their work titled “Local-Global Analysis,” the global domain of ground
heat transfer is calculated using a faster analytical model, and the local domain (near the interface
of the building) is calculated using a numerical method with high resolution to capture thermal
bridging effects.
Using whole-building energy simulation tools for foundation insulation design requires coupled
foundation heat transfer calculations that are both faster than the detailed three-dimensional tools
like TRNSYS and more capable than the simplistic regression-based tools.
The real limitation in estimating heat loss from building foundations is not in our under-
standing of the physics but in the shear number of calculations required to solve the heat diffusion
equation for the spatial and temporal domains at a reasonably high resolution of detail.
26
The following sections describe the calculations required to simulate the physics of foundation
An energy balance of heat flow through soil is commonly treated as a completely conductive
problem. In reality, there are several mechanisms that contribute to heat transfer [39]:
• Conduction through soil grains, water, ice, and entrained gasses (e.g., air)
• Latent heat transfer as water freezes and thaws within the soil matrix
Of the mechanisms listed above, conduction is considered the dominant form of heat transfer
for most cases. Depending on the composition of the soil, other mechanisms may play greater roles
A heat balance on an infinitesimally small volume yields a relationship between two indepen-
dent variables, T , temperature (the potential for heat flow), and Ψ, matric potential (the potential
∂T ∂Ψ
CH,T + CH,Ψ = ∇ · [k∇T ] + ∇ · [DH,Ψ ∇Ψ] + Cp,l ṁl · ∇T + q̇, (2.18)
∂t ∂t
where
• t is time [s],
• Ψ is the matric potential at a given spatial position and point in time [m],
Here, CH,T is the thermal heat capacitance, which accounts for the volume-average heat
capacity of the soil and its constituents as well as the internal energy of water vapor associated
where the subscripts s, l, and v refer to the soil, liquid water, and vapor water components,
respectively, and
CH,Ψ is the matric heat capacitance, which accounts for the internal energy of water vapor
where Cv,Ψ is the matric vapor capacitance which, as Janssen [56] explains, has very little influence
The third term in Equation 2.18, ∇ · (k∇T ), represents the Fourier heat conduction.
The effective thermal conductivity, k, represents the combined conductivity of all the soil
constituents (soil grains, water, air, ice, etc.) as well as the effect of vapor distillation [35].
The effective conductivity can be calculated based on the conductivities and volumetric frac-
tions of the constituents as well as the geometry of the soil grains (details of this calculation can
where
• Subscript p indicates the gas-filled (i.e., air and water vapor) portion of the soil matrix
whose conductivity is calculated to include the effect of vapor distillation in Equation 2.22,
and
• ξ is the ratio of the average temperature gradient in the constituent and the average tem-
∂ρvs
kp = ka + φhf g Da , (2.22)
∂T
where
29
3
ξi = . (2.23)
ki /kl + 2
The fourth term in Equation 2.18, ∇·(DH,Ψ ∇Ψ), represents heat transfer caused by moisture
diffusion. Here, DH,Ψ is the matric heat diffusivity, which is proportional to the matric vapor
diffusivity:
where Dv,Ψ is the matric vapor diffusivity defined in Equation 2.29 in Section 2.3.2 on moisture
diffusion.
The fifth term in Equation 2.18, Cp,l ṁl · ∇T , represents the sensible heat transfer by bulk
fluid flow. The assumption here is that the bulk fluid and the surrounding media are in thermal
equilibrium.
∂Ψ ∂T ∂K
CM,Ψ + CM,T = ∇ · [DM,Ψ ∇Ψ] + ∇ · [DM,T ∇T ] + , (2.25)
∂t ∂t ∂z
where
• Ψ is the matric potential at a given spatial position and point in time [m],
• t is time [s],
CM,Ψ is the matric moisture capacitance, which is comprised vapor and liquid matric capac-
itances:
where
∂θl
Cl,Ψ = , (2.27)
∂Ψ T
CM,T is the matric moisture capacitance, which also has a neglibible effect on the solution
[56].
Calculating the derivatives in Equation 2.27 requires moisture retention curves for specific
soil types, which can be difficult to obtain or generalize for a range of soil compositions and grain
geometries.
The third and fourth terms in Equation 2.25, ∇ · (DM,Ψ ∇Ψ) and ∇ · (Dv,T ∇T ), represent
the diffusion of moisture driven by the matric and temperature gradients, respectively.
where
• θ is the volumetric fraction of a given component (air a, or liquid water l) [m3 /m3 ], and
• θl,c represents the liquid moisture content, below which the liquid water is no longer con-
sidered continuous throughout the porous structure and the relative humidity drops below
100%. This usually occurs between values of 0.1 and 0.2 m3 /m3 .
The fifth term in Equation 2.25, DM,T , is the thermal moisture capacitance, which Janssen
to gravity.
32
The equations established by Deru [39] with the assumptions explained by Janssen [56], lead
∂T ∂Ψ
CH,T + CH,Ψ = ∇ · [k∇T ] + ∇ · [DH,Ψ ∇Ψ] + Cp,l ṁl · ∇T + q̇ (2.30)
∂t ∂t
and
∂Ψ ∂K
CM,Ψ = ∇ · [DM,Ψ ∇Ψ] + . (2.31)
∂t ∂z
Equations 2.30 and 2.31 must be solved simultaneously using an iterative, non-linear solution
method such as the Newton-Raphson method in order to satisfy the conservation of both mass and
energy. Several of the terms in these equations are highly non-linear and can increase the number
of calculations significantly.
Solving the combined heat and moisture diffusion equations requires significantly more in-
formation about the soil properties than is typically available to building energy simulators. A
number of methods have been developed to calculate the thermal and matric properties based on
a soil’s composition [36, 88, 112], but it is not common practice to collect and characterize soil
Most foundation heat transfer tools will simplify the calculations by assuming that the ther-
mal properties are linear (i.e., they do not depend on the temperature of the soil) and without
calculating coupled moisture diffusion. These simplifications reduce the calculations to a single
equation:
∂T
CH,T = ∇ · (k∇T ) + q̇. (2.32)
∂t
The apparent thermal conductivity, k, in Equation 2.32 approximates all mechanisms of heat
diffusion in the porous soil medium and not solely the conduction. The value of k may vary spatially
and temporally while still solving the discretized equation using linear algebra methods; however,
the value of k cannot depend on the calculated temperatures in the domain without invoking
With Equation 2.32, the estimated foundation heat transfer is most sensitive to the value of
soil thermal conductivity [1]. Sterling et al. [105] conducted a survey of the apparent conductivity
of soils found in the United States, which revealed that the value can range between 0.2 W/m-K and
4.0 W/m-K depending on moisture content and soil composition. Rock [99] showed that between
soil conductivities of 0.9 W/m-K and 2.0 W/m-K there can be a 100% difference in the predicted
When it comes to methods for solving the diffusion equations in the context of foundation
heat transfer, there is no “silver bullet.” Many such methods have been presented throughout
the literature, and each method brings unique advantages and disadvantages. For a more in-
depth review of numerical techniques used in solving foundation heat transfer problems, consult
MacArthur et al. [85]. The following sections will provide a brief description and explain the
Lachenbruch [78] demonstrated one of the first analytical solutions for slab-on-grade heat
transfer using Green’s function. Delsante et al. [38] later developed a similar model using Fourier
transforms. These two models, compared by Kusuda and Bean [76], were shown to produce nearly
identical results. The major limitation of both Lachenbruch’s and Delsante’s solutions is that they
Many analytical solutions, including those using Fourier transforms, take the form of an
infinite series. The calculated solution is then an approximation where the series is calculated up
to a large number, N , after which the series is truncated. Larger values of N will improve the
numerical accuracy of the solution but will require more computation time.
Analytical methods are typically faster than the numerical alternatives where finer discretiza-
tions require proportionally more computation time. The major drawback of analytical methods
34
is solutions only exist for a small subset of problems with highly constrained—and not always
Hagentoft and Claesson [48, 26, 51, 49, 50, 52] have developed a number of solutions using
Krarti [62] is another prominent contributing author of analytical solutions. His work is
The finite difference method is one of the more prominent solution methods encountered in
foundation heat transfer tools. Variants of the method can be found in work by Bazjanac et al.
The finite difference method often uses spatial discretization aligned with the axes of the
coordinate system. This imposes a limitation on the range of geometries that can be simulated.
However, this same constraint enables computationally efficient storage and execution of the linear
algebra routines.
Furthermore, building foundations are often constructed using orthogonal surfaces and rarely
require geometries that are not aligned with a Cartesian coordinate system. The restriction that
the finite difference method imposes on the shape of the domain only poses a problem for a small
Finite difference methods are typically categorized as either explicit or implicit schemes de-
pending on how the solution is calculated relative to the solution of the previous time-step. Explicit
schemes do not require the use of linear algebra algorithms and, therefore, require fewer calcula-
tions. Conversely, implicit schemes yield more reliably stable and realistic results. For the standard
explicit scheme formulated in Cartesian coordinates with a uniform spatial discretization and ho-
where
• Δx, Δy, and Δz are the spatial cell size [m] in the x, y, and z directions, respectively.
Several other variants of the finite difference method are discussed in greater detail in Section
3.4.3.
Finite element methods are implicit in nature and require the use of linear algebra algorithms
to solve the set of simultaneous equations resulting from the domain discretization. Computation
Unlike finite difference methods, the spatial discretization does not have to align with the
coordinate system axes. The unconstrained discretization enables the simulation of a wider range
of domain geometries. However, off-axes geometry makes the discretized mesh difficult to generate.
The finite element method is used in the foundation heat transfer tools developed by Mitalas
[87], Baylon and Kennedy [12], Beausoleil-Morrison et al. [17], Deru [39], and Janssen [56].
Response factors are a set of values that relate the heat transfer between two surface bound-
aries to a series of temperature histories. DOE-2 uses response factors to calculate heat conduction
between interior and exterior surfaces of above-grade walls, roofs, ceilings, and floors. EnergyPlus
uses a modified response factor approach called conduction transfer functions or CTFs [111]. A
number of other response-factor-type calculations are employed in foundation heat transfer tools
developed by Davies et al. [34], Krarti et al. [73], Zhang and Ding [121], and Wentzel [116]. Bahn-
fleth and Amber [10] developed Ground Conduction Functions (GTFs) and demonstrated that the
GTFs produce results with 1% of the results from a detailed finite difference model.
36
Response factors must be generated before the simulation, either through analytical calcu-
lations or a numerical preprocessor. Response factors are determined by recording the system’s
response to temperature pulses at the boundaries. Once established, the calculation of heat transfer
based on thermal history can be very computationally effective. The trade off between computation
time and numerical accuracy is determined by the additional computation time required by the
One concern of accuracy when using response factors is the amount of time required to estab-
lish a reasonably consistent thermal history before the results can be valid for use in a simulation.
The ground domain has a very slow response to temperature variations, with some effects verging
on the magnitude of several years. Even if the response factors were to capture an adequately long
thermal history, the conduction calculated in the simulation would still have to be initialized with
Response factors are also generated for a specific set of thermal properties that are assumed
to be static for the duration of the simulation. This approach works for most instances of the
diffusion equation but does not work for any non-linear or time-varying thermal properties.
have an equivalent thermal response. Xie et al. [119] describe a method for creating two (or more)
layered equivalent slabs that represent the foundation heat transfer for the more important diurnal
data. Mitalas [87] developed a regression method that provides steady-state and periodic heat loss
factors to describe below-grade heat loss. Although originally intended for use in deep basements,
the method was later adapted to describe shallow basements and slabs-on-grade. The below-grade
surfaces are separated into subsurfaces, usually an upper and lower wall and a perimeter and core
floor. The heat flux through each section is calculated by steady-state and dynamic heat loss factors
37
derived from a two-dimensional finite element program and tabulated for nearly 100 combinations
Mitalas’s finite element tool was later used to develop the more general BASESIMP regression
by Beausoleil-Morrison and Mitalas [15]. Krarti [64] created a separate regression based on neural
networks, and later compared the methodology to the BASESIMP regressions with consistent
results [14]. Additional regression methods are described by Krarti and Choi [68], Bazjanac et al.
Most regression methods require little computation time, but the modeling capability of
most regression methods is limited by the robustness of the data set that originally informed
the regression. For example, Mitalas’s regression factors were developed for use in Canada, with
consistent snow cover on the ground, limiting the applicability of the regressions to snowy locations.
2.3.5 Discretization
Of the solution methods described above, the analytical methods are too constrained to be
broadly applicable to foundation insulation design. The other four methods are either numerical
inevitably play a role in developing more capable and computationally efficient foundation heat
All numerical approximations of the diffusion equation require that both spatial and tem-
poral domains are divided into discrete cells or time-steps, respectively. The total computational
requirement for a simulation is directly proportional to the number of cells in the spatial domain
multiplied by the number of time-steps in the temporal domain. This section describes how the
numerical discretization is defined and some methods developed to reduce the number of associated
calculations.
38
The spatial domain can be idealized as a semi-infinite solid extending laterally outward and
extending downward from the exterior grade. In a numerical model, the domain is bounded by
the lateral far-field boundaries and the deep-ground boundary. These boundaries must extend
sufficiently far from the building to capture the influence on the soil.
As part of the IEA BESTEST: In-Depth Diagnostic Cases for Ground Coupled Heat Transfer
Related to Slab-On-Grade Construction, Thornton [109] demonstrated that these boundaries had
less than a 0.01% impact when the deep-ground boundary was more than 30 meters below the
building and far-field boundaries were more than 20 meters from the building. A smaller domain
requires fewer discretized cells but may not reflect a semi-infinite nature.
The cells created by the discretization are numerical approximations of the infinitesimally
small elements for which the partial differential equations were derived. As cell size decreases
the approximation approaches the continuous solution. Many authors demonstrate that below a
certain cell size, the results achieve “discretization independence” and are no longer sensitive to
The upper bound on cell size is often determined by the level of detail required in the model.
For the case of foundations, the upper bound is determined by the size of the thinnest component
of the foundation construction represented in the thermal model. For example, a board of rigid
insulation may be as thin as an inch, though cells elsewhere in the domain can be larger.
Often it is necessary to have fine cell spacing near the foundation perimeter, but the spacing
can gradually increase in the direction away from the perimeter. Usually, the cell spacing increases
where fg is a constant growth coefficient, typically somewhere between 1.05 and 1.50. Geometric
The total number of cells in the domain is the product of the number of cells in each of the
coordinate axis directions. Utilizing planes of symmetry eliminates half of the cells in the direction
perpendicular to the plane; however, this only works if the entire problem is symmetric, including
the geometry, thermal properties, and boundary conditions. This means convection algorithms that
are dependent on wind direction and solar incidence on non-horizontal surfaces must be generalized
Another effective way to reduce the number of cells is to eliminate an entire computational
effects near the corners of the foundation. Many tools either neglect the third dimension (along one
system.
Most whole-building energy simulation tools simulate a single year of operation, largely be-
cause they are driven by a single year of weather data. Additionally, most of the energy-related
features of a building do not have effects at longer time scales. This is not the case for features
that interact with the ground. After a building is first heated, the ground can take several years
to come into a steady periodic equilibrium with the building. IEA BESTEST [90] showed that
the initialization period for slab-on-grade simulations can take anywhere from three to ten years to
reach a steady-periodic state, within 0.1% annual heat loss of the previously simulated year. Baz-
janac et al. [13] used an initialization period of four years for the simulations that informed their
regression method. Other mechanisms which transfer heat to the ground, including ground-source
heat pumps [44] and seasonal heat storage [103], also demonstrate similarly long time responses.
For single year analyses, it is best to use simulation results for steady periodic ground conditions
since this represents the state of the building for most of its existence.
40
The number of years required to arrive at steady periodic ground conditions in a simulation
is dependent on how well the initial temperatures reflect the state of the ground at the beginning of
the simulation period. Poor estimates of initial ground temperatures can lead to additional years
Constant temperature: Setting all the cells to a single spatially constant temperature is the
simplest approach to initializing ground temperatures. Many authors have used the annual
mean dry-bulb temperature as specified by Neymark and Judkoff [90] for the IEA BESTEST
work.
Kusuda: Kusuda and Achenbach [75] developed a one-dimensional analytical solution for undis-
turbed soil. This solution is based on heat diffusion into a semi-infinite domain driven by
where
• Ty,ave is the annual average outdoor dry-bulb temperature at the location [K],
• ΔTm is the range of monthly average dry-bulb temperatures throughout a year [K],
• Δtshif t is the time difference between the calendar cycle and the temperature cycle
[s]. This is the time from the beginning of the calendar year to the time of the annual
This approach approximates the entire ground domain as undisturbed by the presence of
a heated building. Though it does not account for lateral variation in soil temperatures,
41
the Kusuda approach is considered a better estimate than a spatially constant temperature
field.
tion in all dimensions, but it does not account for the long-term thermal response provided
Each of these approaches, along with a new initialization method, are compared in Chapter
4 to illustrate the effect each has on reducing the duration of the initialization period.
The discretized time-steps in the temporal domain must be short enough to both capture the
variation in boundary conditions (typically, hourly weather data) and ensure numerical stability in
Judkoff and Neymark [60] characterize three primary methods of evaluating the accuracy of
Empirical Validation: Calculated results from a tool or method are compared to monitored data
Analytical Verification: Outputs from a tool or method are compared to results from a known
analytical solution or generally accepted numerical method for isolated heat transfer under
Each of these methods is discussed in further detail, along with relevant efforts in evaluating
Empirical data provides quantifiable results that are accurate within the uncertainty of the
measurements, but such measurements can be expensive and time consuming. This is particularly
true in the case of foundation heat transfer where measurements may require multiple seasons or
years to produce an adequate data set. A well-designed experiment requires that all unknown
quantities be carefully measured, which in the case of foundation heat transfer would require the
measurement of the thermal properties of the soil at every point in the spatial and temporal domain
of the experiment.
Another disadvantage of empirical validation is that the results are only applicable to the
specific experimental setup at the time of measurement. The range of foundation designs, ground
domain properties, and boundary conditions that affect heat transfer are difficult to capture through
experiment, meaning that the results of the experiments would only serve to validate a small subset
There have been relatively few efforts in the past to validate the accuracy of foundation heat
transfer tools. Experimental validation has traditionally been cost and time prohibitive and has
produced only a limited set of measurements. Two of the more significant efforts of experimental
validation are authored by Rees et al. [94, 93, 107, 97, 98, 108] and Adjali et al. [1, 3, 4]. They
found that in many cases the models produce reasonably accurate results and that most of the
errors were related to uncertainty in the model input (namely the thermal properties of the soil)
and the inability to model the impacts of rain and snow. Additional empirical validation efforts
can be found in Trethowen and Delsante [110] and Rantala and Leivo [91]. A more comprehensive
In many cases, attempts to validate foundation heat transfer tools becomes an exercise in
calibrating a model rather than a true validation of the tool due to the high degree of uncertainty
Analytical solutions can be used to verify the implementation of the algorithms in computer
code. Since analytical solutions are well defined and highly constrained, analytical verification does
not introduce the same problem of input uncertainty inherent to experimental validation. However,
these constraints also limit the application of the verified results to a similarly small subset of the
parameter space.
analytical verification for cases where an analytical solution cannot be derived. These solutions
must first be verified against the analytical solutions, where they exist, to anchor the veracity of
their implementation.
For analytical verification, once the solution has been established, it is relatively quick and
inexpensive to verify other tools against it. However, verification is limited to the ground domain
subsystem of the building where a problem is reasonably well-defined and cannot encompass the
added complexity at the whole-building level without straying from the well defined analytical
anchor.
The most comprehensive example of analytical verification is the International Energy Agency
Building Energy Simulation Test and Diagnostic Method (IEA BESTEST ) for ground coupled
heat transfer related to slab-on-grade construction [90]. The IEA BESTEST procedures include
verification against the analytical solution developed by Delsante et al. [38] and several verified
numerical solutions.
The IEA BESTEST for ground coupled heat transfer related to slab-on-grade construction
establishes test cases for rectangular, uninsulated slabs with constant thermal properties of soil and
constant convective boundary conditions. IEA BESTEST has yet to produce additional test cases
to explore other foundation shapes, types, insulation designs, and boundary conditions.
44
Comparisons of simulation results among different tools or methods reveal inaccuracies where
there is poor agreement. However, agreement under comparative testing does not guarantee ac-
curate results since there is no absolute truth standard like there is in empirical validation and
analytical verification.
Like analytical verification, comparative tests can be executed inexpensively to quickly iden-
tify discrepancies. The tests can also be designed to focus on any portion of the system as well as
Previous whole-building comparative tests described in the Home Energy Rating System
(HERS) BESTEST include four cases highlighting heat loss through foundations [59]. These cases
include insulated and uninsulated slabs, and insulated and uninsulated basements, illustrated by
Deru et al. [40]. Several comparative test cases in the IEA BESTEST specification [90] revealed
disagreements in predicted heat loss between 1%–24% across a number of simulation tools.
The BESTEST procedures have laid the foundation for improving the accuracy of foundation
heat transfer tools but are limited in the range of capability they can test. Additional validation
is still needed for a wide range of foundation designs, ground domain properties, and boundary
conditions.
This section takes a closer look at some of the more prominent foundation heat transfer tools
and methods either frequently used by energy modeling practitioners or cited often in the literature.
The methods and tools discussed in this section are qualitatively compared side-by-side along with
the tool developed as part of this thesis work in Table 2.3. Kiva is described in detail in Chapter
3. Claridge [27], Adjali et al. [2], Krarti [66] and Deru [39] also provide thorough reviews of the
various foundation heat transfer methods and tools found throughout literature.
Table 2.3: Comparison of prominent foundation heat transfer tools and methods
45
Finite
Solution Method Analytical Regression Analytical Difference Difference Regression Difference Difference
Element
(Explicit) (ADI) (Implicit) (Multiple)
Mesh Generation - - - Automatic Automatic - Manual Manual Automatic
46
Some of the oldest foundation heat transfer calculation methods can be found in the ASHRAE
Handbook of Fundamentals [7]. The handbook describes separate methods corresponding to the
type of foundation. Heat flow from insulated basement walls is calculated based on a method
developed by Latta and Boileau [79] in 1969 where steady-state conduction is approximated through
The basement wall and floor are separated into discrete sections, each with its own radial
path with only angular variation in thermal properties. This allows for relatively simple calculation
using polar coordinates where each section of the wall or floor can have different levels of insulation.
Heat transfer through slabs is defined by a perimeter conductance factor found in look-up
tables. The factors are provided for only four slab configurations based on regressions from finite
The ITPE technique developed by Krarti et al. [71, 70, 72] is the most complete analytical
solution for foundation heat transfer. The basic premise of this method is to break up the domain
into zones so that each zone is a rectangular region with unique boundary conditions. The heat
diffusion equation is solved for each zone using an estimated temperature profile as the boundary
condition between zones, which satisfies a continuity of heat flux across the boundary. This method
has been adapted to work in either two- or three-dimensions. Applications of the ITPE technique
Bahnfleth [9] and Bahnfleth and Pedersen [11] established a three-dimensional finite difference
solution for slab-on-grade floors using a standard explicit solution scheme. This solution was
47
originally used to examine sensitivity to shading caused by the building, evapotranspiration, and
Later, Cogil [30] developed a similar solution for basements using an alternating direction
implicit (ADI) finite difference scheme. The solution included a detailed radiation exchange model
and heat balance for the basement space (including heat losses through the sill box) that was used
Bahnfleth’s and Cogil’s tools were adapted by Clements [29] to provide preprocessed results
for use in the EnergyPlus simulation engine [111]. In the process of adapting the code for use in
EnergyPlus, the shading algorithm in the slab program and the radiative exchange model in the
basement program were removed to reduce computation time. The preprocessors output monthly-
average surface temperatures for ground-coupled surfaces. These temperatures serve as boundary
conditions for the Conduction Transfer Functions (CTFs) used to calculate one-dimensional un-
2.5.4 eQUEST
eQUEST (the QUick Energy Simulation Tool) is a popular graphical user interface to the
DOE-2 simulation engine [55]. Huang et al. [54] integrated a fully implicit, two-dimensional finite
difference solution into DOE-2 to establish foundation insulation guidelines for the Building Foun-
dation Design Handbook [77]. A large set of simulation data, 88 foundation configurations in 13
cities, was generated and used to establish the regression method employed by Winkelmann [118],
where perimeter conduction factors characterize the effective steady-state resistance of an under-
ground surface. The dynamic effects of heat transfer through surfaces in DOE-2 are calculated
using a form of response factors. As implemented in DOE-2, the response factors are limited to a
thermal history of 100 hours and are unable to capture the longer time-frame responses typically
seen in foundation heat transfer [13]. To best represent the thermal response under this constraint,
the actual wall is modeled as a layered construction with one foot of soil.
48
Krarti and Choi [69] found that this approach tends to underestimate the mass effect of the
ground and cannot properly represent indoor temperature changes within the course of a day.
Deru [39] developed one of the most complete solutions with coupled heat and moisture trans-
fer. His solution employs a two-dimensional finite element method with moisture and temperature
dependent thermal properties. The tool can also simulate the effects of evapotranspiration and
ground freezing.
2.5.6 TRNSYS
TESS (Thermal Energy Systems Specialists) created a ground coupling library for use in the
TRNSYS simulation engine [104]. The library employs a three-dimensional finite difference model
that can interface with any number of zones in a building. The TRNSYS ground-coupling library is
currently one of the only tools where the foundation heat transfer calculations are directly coupled
to the zone heat balance model through iterative simulation. TRNSYS is the only whole-building
simulation tool with a ground heat transfer solution that was used as a secondary truth standard
in IEA BESTEST for ground-coupled heat transfer [90]. TESS also compared TRNSYS to several
Figure 2.4: Illustration of ITPE zones (1, 2, and 3) for a basement foundation (adapted from Krarti
et al. [70])
Figure 2.5: Concentric heat flow paths (adapted from Latta and Boileau [79])
Chapter 3
Tool Development
3.1 Kiva
Kiva is a free and open-source C++ program developed by the author for the purposes of
this thesis and the source code is currently hosted at GitHub.com [74]. The name comes from the
subterranean ceremonial structures used by the ancient Puebloan peoples of the present-day states
of Colorado, Utah, Arizona, and New Mexico. Kivas are among the earliest constructed structures
The need for Kiva was established by observing the difficulty of the current tools to achieve
fidence. Such a balance is a subjective assessment and will vary depending on the application of
the tool. However, in the context of whole-building energy simulation, there are some precedents
Capability: The tool should provide methods of describing a large range of potential foundation
insulation designs. Such a range cannot be objectively defined, but some of the more
• Continuously variable foundation depth (for different basement and crawlspace heights).
51
soil).
Complexity: The level of complexity required to describe a foundation system should be within
the expected domain knowledge of the user. In this case, users are typically mechanical
engineers or architects with domain knowledge pertaining to buildings and heat transfer.
Users should be able to confidently define a foundation system without being confronted
by the numerical settings that are only tangentially related to estimating building energy
use (e.g., spatial discretization, or meshing, and convergence tolerance for a linear solver).
While it is important for these settings to be available to a user, the settings should not be
required inputs that distract from modeling information relevant to the foundation design.
Computational Performance: Current whole-building simulation tools require from tens of sec-
onds to several hours to simulate a single building’s energy use, depending on the complexity
of the building model and the solution methodologies employed. As foundation heat trans-
fer represents only a fraction of the energy load on the building, the calculations should
represent a relatively similar fraction of the overall computation time and not several orders
of magnitude longer, which is the case for many of the more detailed numerical solutions.
Computation performance in this paper is measured in wall time, or “real world” time,
as it is perceived by the user. This is used to differentiate from CPU time, which is the
Algorithmic Confidence: The calculations performed should reflect knowledge of the physical
phenomena of the heat transfer process. Several methods and tools attempt to simplify the
calculations without justification. There are currently few highly-detailed tools available
heat transfer. The calculation of foundation heat transfer is comprised of several components
including:
• the spatial and temporal discretization of the governing partial differential equation,
• the numerical scheme used to progress the simulation from one time-step to the next,
Each component listed can be addressed using a variety of different approaches, with each approach
having different impacts on the computational performance and accuracy of the solution. The Kiva
framework allows for testing of different combinations of approaches while maintaining consistency
The purpose of Kiva is to provide a framework with a highly capable and detailed reference
solution against which the approximations and simplifications of subsequently developed solutions
can be verified. At the time of this work, Kiva is a standalone computer program which simulates
heat transfer through building foundations into the surrounding ground. Though it is not cur-
rently integrated into any whole-building simulation engines, the framework has been generalized
For the most part, tools used for building energy simulation have been developed in a closed
environment. This approach has led to several overlapping efforts to provide many of the same
capabilities (e.g., DOE-2, EnergyPlus, and TRNSYS). There are occasions where this redundancy
is beneficial, such as comparative testing. However, it also means that researchers must often
reinvent the wheel. This is especially true for foundation heat transfer tools.
Several tools have been developed for graduate thesis research, including but not limited to
Bahnfleth [9], Davies [33], Deru [39], Janssen [56], Krarti [62], Shen [101]. The code from several
of these theses has since been incorporated in some fashion into currently used simulation engines:
EnergyPlus (Bahnfleth [9]), DOE-2 (Shen [101]), SUNREL (Deru [39]), and APACHE (Davies [33]).
Without available source code, each of these authors spent a significant amount of time
developing a software tool to perform their analyses. If the code were to be made available in an
open-source environment, future researchers could use more of their time improving and enhancing
the code that already exists in collaboration with other researchers rather than investing time
Many of the existing foundation heat transfer tools are written in Fortran, which is well
established for its computational performance as a compiled language. However, the popularity,
usage, and support of the Fortran language has declined considerably over the years as more
modern languages, such as C++, have introduced improvements such as object-oriented design
with a relatively small difference in computational performance. Using a more supported and
The following sections describe the modeling and simulation capabilities of Kiva. As an open-
source framework, anyone is free to adapt Kiva to include additional features and capabilities.
54
The description of the foundation design is provided within the two-dimensional context
illustrated in Figure 3.1. This profile is applied along the entire perimeter of the foundation.
where Δzwall-above-grade and Δzf oundation are foundation design user-inputs and Δzdeep-ground is a
The foundation type is defined by the combination of Δzwall-above-grade and Δzf oundation
The foundation insulation and structural design are described by six elements illustrated
in Figure 3.2. The components of the insulation and structural design elements each reference a
55
set of thermal properties: thermal conductivity, specific heat, and density. The locations of the
Foundation Wall: The composition of the foundation wall is defined by a series of layers extending
away from the interior wall surface. Although the footer is not explicitly described, the
wall height user-input, Δzwall , can be extended to represent the footer depth.
Slab: The slab is comprised of a series of layers extending below zf loor . Whole-slab insulation is
Interior Vertical Insulation: Interior vertical insulation begins at zwall-top and extends down-
ward and inward to a user-specified depth and thickness. The depth can be specified to
Interior Horizontal Insulation: Interior horizontal insulation begins at the wall’s interior sur-
face and extends inward and downward to a user-specified width and thickness at a user-
Exterior Vertical Insulation: Exterior vertical insulation begins at zwall-top and extends down-
ward and outward to a depth (below grade) and thickness specified by the user.
Exterior Horizontal Insulation: Exterior horizontal insulation begins at the wall’s exterior sur-
face and extends outward and downward to a user-specified width and thickness at a user-
The slab and wall elements also describe the long-wave emissivity and short-wave absorptivity of
the surface.
The level of specification available for each of the insulation and structural elements allows
the user to describe the foundation system in great enough detail to simulate some of the important
meaning that it must be comprised of right angles. However, circles can also be simulated in two-
list of x-y Cartesian vertices tracing the foundation perimeter in a clockwise fashion (see Figure
3.3).
The perimeter of this polygon defines the location of the interior surface of the foundation
wall in the profile shown in Figure 3.1. The positioning of the foundation insulation and structural
The two-dimensional context for the foundation system description reduces the complexity
for the user to create a fully three-dimensional spatial domain. However, this convenience imposes
• It is not possible to simulate heat transfer between multiple thermal zones through the
The far-field boundaries define the extent of the domain in all four of the lateral directions.
These boundaries are created at a distance extending outward from a rectangle bounding the
polygon specified by the user (Figure 3.3). The far-field distance is a simulation input that defaults
to 40 meters.
Figure 3.3: Plan view illustration of far-field boundary definition in three-dimensions and founda-
tion shape vertex definition
A reference to material definition sets the thermal properties (i.e., thermal conductivity, spe-
cific heat, and density) of the soil. Currently, the domain soil has spatially homogeneous properties
and does not account for effects such as moisture content or freezing.
58
The following sections describe the algorithms used to determine boundary conditions in
Kiva.
The user assigns a temporally constant temperature for the interior thermal zone. This
temperature is used to calculate interior convection according to Equation 2.1. The user can also
override the interior convection coefficient to a constant value to replicate the idealistic constraints
Interior long-wave radiation calculation does not utilize an algebraic method to predict radia-
tive exchange among interior surfaces. Instead, surface radiation is approximated as an exchange
Kiva is programmed to account for incident short-wave solar radiation on the interior foun-
dation surfaces. As a stand-alone tool, Kiva does not calculate transmitted solar through windows
The boundary at the top of the foundation wall is adiabatic. In the context of a whole build-
ing simulation, conduction through above-grade walls can still be simulated as a one-dimensional
problem, implying that multi-dimensional impacts at the wall interface are neglected. The wall
boundary can also be set as a constant linear temperature profile to replicate the idealistic con-
Weather data in the EnergyPlus Weather (EPW) file format drives the exterior boundary
conditions in Kiva. Exterior convection is simulated using the relationships described in Section
2.2.3.4. Kiva assumes that the weather station is located in flat, open country at a height of 10
meters and adjusts the wind speed using a user-specified boundary layer thickness and terrain
exponent from Table 2.2. A user-specified vegetation height defines the height for the wind speed
Exterior long-wave radiation is calculated according to Equation 2.16 to account for radiation
exchange between the surface and the ground, air, and sky.
The EPW file provides the direct normal and diffuse horizontal components of solar radiation.
The location latitude and longitude in the EPW file are used to calculate the solar incidence angle
according to the solar position equations provided by Duffie and Beckman [43]. Kiva contains an
experimental capability for simulating the effect of shading on exterior surfaces. This technique
uses the graphics processing unit of the computer to count the pixels of a surface rendered from the
perspective of the sun. The application of pixel counting for use in solar shading calculations was
originally presented by Jones and Greenberg [58]. While Jones and Greenberg [58] demonstrated
that this technique greatly improves the speed of shading calculations, their tests were applied to
cases where the number of shading surfaces was high relative to the number of receiving surfaces.
Preliminary tests of this technique applied to Kiva, where there are potentially thousands of re-
ceiving surfaces (one for each exterior surface cell), yielded prohibitively long computation times.
The effect of shading is not explored further in this thesis, but remains an area for improvement
within the Kiva framework. For the results presented here, Kiva applies unimpeded solar incidence
The far-field boundary is applied as a zero-flux boundary. By default, the deep-ground bound-
ary is a zero-flux boundary, but the user may choose to apply a constant temperature boundary
As an open-source framework, relatively little overhead effort is required to reuse the code in
the context of an established whole-building energy simulation tool. Portions of the code have been
written explicitly to accommodate a more dynamic interaction between a building and the ground
domain, including variable interior temperatures, incident short-wave heat fluxes, and internal heat
This section describes the solution methods used in Kiva to calculate heat transfer from
Heat transfer in the ground domain is solved using the diffusion equation
∂T
CH,T = ∇ · [k∇T ] (3.4)
∂t
which assumes there is no moisture flow. The thermal properties, k and CH,T = cp ρ, are assigned
by the materials referenced for the soil and the foundation design elements.
61
3.4.2 Discretization
Equation 3.4 is solved for the solid portion of the domain bounded vertically by zwall-top
and zdeep-ground , and laterally by the far-field boundaries illustrated in Figure 3.3. For three-
dimensional simulations the user can request that the domain use planes of symmetry if they exist
for the foundation shape defined by the user. Kiva automatically detects symmetry in both the x
and y directions and applies zero-flux boundaries to reduce the domain size. Orientation-specific
boundary conditions, such as solar incidence and wind driven convection, are applied to the reduced
focus of Chapter 5.
The spatial discretization is defined for four distinct regions of the domain. The region
bounding the foundation wall and insulation elements defines the near-field region. All other
regions are defined either laterally (interior and exterior regions) or vertically (deep region) relative
to the near-field region. The near-field, interior, exterior, and deep regions are illustrated in Figure
3.4.
Cells grow geometrically towards the far-field, deep-ground, and symmetry boundaries. Cells
grow towards the center of each element within regions that do not border one of the domain
boundaries. Figure 3.5 illustrates cell growth and distribution for two examples, one in profile view
Cell size and growth are determined by five parameters: four which determine the cell growth
coefficient in each of the regions and another parameter defining the minimum cell size for the entire
domain. The minimum cell size defines the allowable number of cells between element or region
boundaries. The cells’ sizes are allowed to increase to fit within the element or region boundaries
according to the growth and distribution of the cells. For example, the length of N consecutively
62
(a) Profile: basement with exterior vertical and (b) Plan: uninsulated, J-shaped slab. Growth coef-
horizontal insulation. Growth coefficient = 1.20 ficient = 1.20 (for all regions), minimum cell size =
(for all regions), minimum cell size = 0.01 meters 0.1 meters
N
l = Δxmin + Δxmin fg + Δxmin fg2 + . . . + Δxmin fgN = Δxmin fgi , (3.5)
i=0
63
where Δxmin is the minimum cell size [m] and fg is the growth coefficient. If l is greater than the
distance between two element or region boundaries at N but not at N − 1, then the actual cell size
is set using
l
Δxactual = N
−1
. (3.6)
fgi
i=0
Kiva creates extra cells along the boundaries to represent surface temperatures and heat
fluxes. These cells have no thickness and are completely planar with the surface boundaries. The
other cells within the same plane as these boundary cells also have no thickness and require modified
material properties if the normal cells on either side of them are different materials. Figure 3.6
exemplifies zero-thickness cells resulting from the wall and floor boundaries. In this figure, a node
represents the center of each cell. The interior air cells are set to the interior air temperature and
Figure 3.6: Boundary cells and resulting zero-thickness cells. Black nodes = normal solid cells,
white nodes = boundary cells, gray nodes = zero-thickness cells, hollow nodes = air cells.
64
Kiva can simulate any number of years of building operation. The user can select any
time-step duration. There are four initialization methods available: three traditional initialization
methods (constant temperature, Kusuda, and steady-state) and an accelerated initialization method
with several variants. The effectiveness of each of these initialization methods at predicting the
Of the various mathematical solutions employed by other tools and methods described in
Table 2.3, the finite difference methods offer several benefits over the analytical and finite element
methods. Finite difference methods allow for more flexibility in describing boundary conditions
than analytical methods and are less mathematically abstract than finite element methods.
In general, the variants of finite difference methods are characterized by how the new time-
Consider a very simple one-dimensional problem, discretized into three cells (one internal
i-1 i i+1
t
t+1
Figure 3.7: Example finite difference scheme (black nodes are known values, white nodes are
unknown values)
65
The problem is always defined at the previous time-step, t (as defined by the initial conditions
or subsequent time-step solutions). For some finite difference schemes, the solution for the new time-
step, t+1, can depend solely on previous time-step’s known values and can be solved independently
(explicitly) to find the solution for the new time-step. Other solutions may depend on other
unknown values and must be solved simultaneously (implicitly) using linear algebra techniques.
The manner in which each scheme solves the discretized solution will affect the computation
time and the stability of the solution as well as how physically realistic the solution is. Though
descriptions of each scheme can be found in the literature, they are rarely illustrated together using
consistent notation. Understanding the differences between the schemes’ mathematical implemen-
The simplest explicit scheme is one which only uses values from the previous time-step to solve
for the new time-step. Each of the three unknowns shown in Figure 3.7 can be solved independently
and sequentially based on the previous values of the node and its neighbors (indicated by the arrows)
t+1
Ti−1 = f T ti−1 , T ti
Tit+1 = f T ti−1 , T ti , T ti+1 (3.7)
t+1
Ti+1 = f T ti , T ti+1 ,
The simple explicit scheme is conditionally stable and does not always provide physically
realistic solutions. The solution becomes more stable with smaller time-steps and larger cell sizes.
In an implicit scheme, all unknowns must be solved simultaneously since they cannot be
explicitly calculated from known values. In the basic implicit scheme (often referred to as “fully”
implicit), each unknown in the new time-step is calculated from its previous time-step value and
the values of its neighbors in the new time-step. In Figure 3.9, there is only one step since all
i-1 i i+1
t
t+1
t+1
Ti−1 = f T ti−1 , Tit+1
Tit+1 = f T ti , Ti−1
t+1 t+1
, Ti+1 (3.8)
t+1
Ti+1 = f T ti+1 , Tit+1
Since each equation has multiple unknowns, implicit schemes require linear algebra techniques
to solve the simultaneous set of equations. The linear system can be solved either directly (using
Gaussian elimination) to arrive at an exact solution (within the floating point precision of the
machine), or they can be solved iteratively within a specified convergence tolerance. In general,
direct solvers are more robust but can require substantially more memory. While Kiva has the
The fully implicit scheme is both unconditionally stable and always provides physically real-
istic solutions.
The Crank-Nicolson scheme is another implicit scheme, but it relies on the average neighbor-
ing values from both the previous time-step and the new time-step as a better approximation of
i-1 i i+1
t
t+1
t+1
Ti−1 = f T ti−1 , T ti , Tit+1
Tit+1 = f T ti−1 , T ti , T ti+1 , Ti−1
t+1 t+1
, Ti+1 (3.9)
t+1
Ti+1 = f T ti , T ti+1 , Tit+1
A promising explicit scheme is the alternating direction explicit (ADE) scheme developed by
Saul’yev [100]. This scheme is similar to the simple explicit scheme except that it uses the known
values of the subsequently calculated neighbor nodes to calculate the value for the new time-
step. The unknown values are solved independently using sweeps starting from each boundary (see
Mathematically, the dependencies for the upward sweep temperatures, U , and the downward
t+1
Ui−1 = f U ti−1 , U ti
Uit+1 = f U t+1
i−1 , U i , U i+1
t t
(3.10)
t+1
Ui+1 = f U t+1
i , U i+1
t
69
t+1
Vi+1 = f V ti , V ti+1
Vit+1 = f V ti−1 , V ti , V t+1
i+1 (3.11)
t+1
Vi−1 = f V ti−1 , V t+1
i
The final temperature, T , is then the average of the two independent solutions:
U +V
T = (3.12)
2
Because each solution is independent, they can be solved in parallel to further reduce com-
putation time.
The ADE scheme has been shown to be unconditionally stable and give physically realistic
results when used with a uniform grid. However, Fukuyo [46] has shown that the ADE method is in
fact conditionally stable for nonuniform grids with some fairly loose stability requirements related
to cell size and the relative growth of cells. For applications of foundation heat transfer, the ADE
Each of the schemes described so far can be explained with a simple one-dimensional exam-
ple. The alternating direction implicit (ADI) scheme applies only to problems with two or more
70
dimensions. In the typical implicit schemes described (i.e., fully implicit and Crank-Nicolson), each
nonboundary cell requires an equation with three unknowns (i.e., the cell itself, the cell before, and
the cell after). In a three-dimensional solution, each cell requires an equation with seven unknowns
A common way to represent these problems is to show a matrix sparsity pattern. This plot
shows a row for every equation and a column for every unknown in the discretized system. A block
is filled in for each unknown that appears in the given equation. The sparsity pattern for a typical
Figure 3.13: Typical implicit matrix sparsity pattern for three dimensions (up to seven unknowns
per equation)
Unlike the ADE method, where “alternating direction” refers to the order in which the equa-
tions are solved (using alternating sweeps), the alternating direction in ADI refers to something
entirely different. Here, each time-step is divided into smaller sub-time-steps: one for each dimen-
sion in the problem. Each sub-time-step is solved with only the unknowns along the corresponding
dimension (e.g., x, y, or z) being solved implicitly (at the new sub-time-step), with the values of the
other directions using the values from the previous time-step similar to the simple explicit scheme.
The equations for each sub-time-step can be arranged so that all the unknowns fall along the
three inner-most diagonals in the sparsity pattern (as shown in Figure 3.14).
71
Figure 3.14: ADI matrix sparsity pattern for three dimensions (up to three unknowns per equation)
This linear algebra problem can be solved quickly and directly using a standard tri-diagonal
matrix algorithm. This algorithm is significantly faster than an iterative solver and other more
generalized direct solvers. It is, however, less efficient than the explicit schemes because it performs
two serial passes over the data elements (i.e., it cannot be parallelized like the ADE approach), and
it must calculate additional sub-time-steps within each time-step corresponding to each dimension
in the problem.
The specific ADI scheme used in this work is described by Chang et al. [21]. This scheme
introduces some nuances that can improve solution stability and accuracy, even at larger time-steps.
Although one cannot guarantee stability for applications of foundation heat transfer, the stability
The formulations of the discretized approximations of Equation 3.4 for each of the finite
difference methods described above are provided in Appendix A for a two-dimensional axisymmetric
cylindrical coordinate system. Formulations for Cartesian two- and three-dimensional coordinate
systems can be easily extended by omission or addition of certain terms as described in the appendix.
72
All implicit finite difference schemes require linear algebra techniques to solve a set of si-
multaneous equations resulting from discretized equations. The implicit schemes available in Kiva
include fully implicit, Crank-Nicolson, and Alternating Direction Implicit (ADI) schemes. As pre-
viously mentioned, the tri-diagonal nature of the ADI scheme lends itself to a straightforward,
computationally efficient solution for linear system of equations. The other schemes utilize iter-
ative linear solvers available in the Library of Iterative Solvers (Lis) [106]. Lis is an open-source
• Orthomin,
• Jacobi,
• Gauss-Seidel, and
Lis also provides several preconditioners for accelerating the solution process including
• Jacobi,
Each combination of a solver and, if selected, a preconditioner impacts the efficiency of the
iterative process differently. A limited test of several combinations revealed that the biconjugate
gradient stabilized (BiCGSTAB) solver, coupled with an incomplete LU factorization (ILU) precon-
ditioner, provided the fastest convergence and most consistent results relative to other combinations
3.5 Outputs
Kiva has the capability of providing both time-series and graphical output. The time-series
where effective temperature is the temperature that can be applied as a preprocessed boundary con-
dition for foundation surfaces in a whole-building simulation tool. The optional surfaces available
• slab core,
• foundation wall;
where the core and perimeter surfaces of the slab are defined by a user-specified distance inward
Kiva can output images representing any two-dimensional slice of the spatial domain and any
point within the temporal domain. A series of images can be created at a user-specified frequency
for compilation into an animation. Figure 3.15 demonstrates this capability with a profile slice and
a plan slice for the same domains illustrated in Figure 3.5. Kiva can create plots of temperature,
heat flux magnitude, and heat flux in the direction of any of the coordinate system axes.
(a) Profile: basement with exterior vertical and hori- (b) Plan: uninsulated, J-shaped slab at 3 meters
zontal insulation (isotherms equally spaced at 2.5◦ C) depth (isotherms equally spaced at 5.0◦ C)
Below-grade
1.0 0.3
crawlspace
Shallow below-grade
0.6 0.3
crawlspace
Above-grade
1.0 1.0
crawlspace
Solution Verification
Simulation validation describes the effort of determining the accuracy of simulation outputs when
compared to carefully measured results from a controlled experiment. Solution method verification
describes the process of testing the correctness of the algorithms as programmed in a tool.
While, ultimately, an empirical validation may provide the best truth standard for deter-
mining a solution’s accuracy, it is difficult to design an experiment where the properties of the
foundation and ground domain are known within a reasonable degree of uncertainty. When creat-
ing a model of the foundation and ground system, this uncertainty propagates through the solution
resulting in large error bounds on the output results of the simulation. Most attempts to validate
a simulation of a foundation heat transfer result in a calibration of uncertain model inputs rather
For this reason, validation is not an objective of this thesis. This is not intended to undermine
the importance of validation, but given the challenges it presents and the evident need to improve on
the computational performance and algorithmic confidence of current solution methods, validation
Kiva provides a framework with several variants of the finite difference method allowing direct
comparisons of the accuracy, computation time, and stability of each method as they are applied to
solving the foundation heat transfer problem. Such comparisons have not been documented before
in the literature.
77
Before the Kiva framework can be used to develop faster simplified solutions, each of the
finite difference solution methods implemented in Kiva must be independently verified against an
existing mathematical truth standard. This will establish Kiva as an accuracy benchmark for
This chapter focuses on the verification of the finite difference solution methods implemented
in Kiva and discussed in Section 3.4.3. This chapter also investigates solution sensitivity to selected
numeric parameters as well as the efficiency of several initialization techniques within the context
As all numerical discretizations are approximations of the continuous nature of heat diffusion,
the goal is to compare the accuracy of these approximations against a known analytical solution. For
foundation heat transfer, the known analytical solutions tend to be highly constrained, mathemat-
ically idealized, physically unrealistic cases that can only provide a starting point for verification.
Beyond these solutions, the only available comparisons are to other independently developed nu-
merical solutions. While these numerical solutions cannot be considered truth standards like the
analytical solutions, a general consensus among detailed numerical solutions can establish a general
range of acceptance.
The International Energy Agency Building Energy Simulation Test and Diagnostic Method
(IEA BESTEST ) [90] describes a methodology to verify solutions for foundation heat transfer
against analytical solutions and verified numerical solutions. A single analytical solution developed
by Delsante et al. [38] was used as a steady-state mathematical truth standard for a single test
case. For the same test case, three detailed numerical solutions were developed and shown to give
very close agreement with the analytical solution. These three solutions, TRNSYS [109], FLUENT
[89], and MATLAB [31], were then used as secondary mathematical truth standards for the other
steady-state and harmonic (unsteady) test cases. These are all idealized test cases with several
inherent limitations:
78
• The test cases are for slab-in-grade foundations only. They do not test slabs-on-grade,
• The test cases do not test the sensitivity to any level of insulation. The slabs are all
uninsulated.
• The entire domain has homogeneous soil properties, meaning that the concrete of the slab
and foundation wall has the same thermal properties as the soil.
• The boundary conditions are approximated with constant film coefficients. There is no
Results from the analytical truth standard and the three verified numerical solutions are
compared to the solution methods used in Kiva for each of the BESTEST test cases. Each case
uses symmetry to reduce the size of the domain and, consequently, the number of computations
(to 1/4). The BESTEST procedures do not specify any sort of pass/fail criteria for the solutions
outside of subjective comparison. For the purposes of this paper, verification implies there are no
Before this work, the accuracy and computation time of different finite difference schemes have
not been compared side-by-side in a common framework for foundation heat transfer calculations.
Traditionally, researchers have selected schemes based on (1) ease of implementation (typically
explicit schemes), (2) stability (fully implicit methods), (3) or availability of commercial software
(usually finite element). Each of the three secondary mathematical truth standards uses some form
The original source of secondary mathematical truth standards is not available to measure
their computation performance under a consistent computational environment, nor were the simu-
lation times provided in the BESTEST literature. However, due to their implicit nature, each of
the solutions is expected to require computation times on the same order of magnitude as Kiva’s
Table 4.1 shows the specification of each of the steady-state test cases. The only distinction
between cases not listed in Table 4.1 is that the boundary at the wall/ground interface is a constant
temperature profile for the analytical test case (GC10a) and adiabatic for all other cases.
Slab dimensions hint hext Ground depth Far-field width Soil conductivity
Case
[m x m] [W/m2 ·K] [W/m2 ·K] [m] [m] [W/m·K]
GC10a 12 x 12 ∞ ∞ ∞ ∞ 1.9
GC30a 12 x 12 ∞ ∞ 30 20 1.9
GC30b 12 x 12 100 100 15 15 1.9
GC30c 12 x 12 7.95 ∞ 15 8 1.9
GC60b 12 x 12 7.95 100 15 15 1.9
GC65b 12 x 12 7.95 11.95 15 15 1.9
Other characteristics similar to all steady-state test cases include the following:
The IEA BESTEST User’s Manual specification (Part I of Neymark and Judkoff [90]) for
Demonstrate that the current level of modeling detail yields negligible (≤ 0.1%)
change in results versus a lesser level of detail; examine the effects of shallower deep
80
ground boundary, shorter far-field boundary distance, and less detailed mesh (if
applicable). For ground depth and far-field length variations, increase the number
of mesh nodes proportionally to the increase in soil volume being modeled.
The following sections describe the sensitivity of the GC10a results to changes in mesh detail,
boundary distances, and solver convergence tolerance. These analyses were performed using the
steady-state finite difference solution. Other solutions are expected to have similar results, when
scaled accordingly. The results from the sensitivity analysis are used to inform the numerical
Each of the three secondary mathematical truth standards uses a nonuniform mesh where the
mesh size increases geometrically with the distance from the foundation wall boundary. This is also
the case with Kiva where the mesh is uniform along the wall boundary (in the x and y directions)
and increases geometrically downward (towards the deep-ground boundary), inward (towards the
center-of-slab planes of symmetry), and outward (towards the far-field boundaries). Figure 4.1
shows a profile of the mesh generated by Kiva near the wall boundary for the GC10a test case.
Sensitivity to mesh detail was tested by varying the minimum allowable cell dimension,
Δxmin , and the geometric growth factor, fg . The size of each cell, i cells away from the boundary
The minimum allowable cell size is used to set the size and number of cells in the uniform
mesh region. The actual cell size is allowed to increase in order to fit uniformly into the dimensions
allotted (in this case, the width of the wall boundary). Similarly, the growth factor is allowed to
To demonstrate that the tested mesh detail yields negligible change in results versus a less
detailed mesh, Kiva was tested under a combination of different minimum cell dimensions and
growth factors over a range of values. The results are shown in Figure 4.2. The approximated slab
81
Figure 4.1: BESTEST GC10a mesh growth near the wall boundary
heat loss increases as the geometric growth factor decreases before settling near the value of the
analytical solution at geometric growth factors less than 1.3. The slight peak at growth factors
of 1.3 may be related to the actual sizing of the cells relative to the input minimum cell size and
maximum growth factor. Simulations at higher levels of detail (i.e., fg < 1.10) were beyond the
memory limitations of the computing hardware, so it is difficult to say definitively whether the
No report from the three secondary mathematical truth standards indicated a complete
parametric sweep across ranges of these two parameters. Instead, mesh independence was tested
either with a single variable sweep (i.e., either Δxmin or fg , but not both) or with a small subset
of combinations. Therefore, it is difficult to know whether the other tools demonstrated similar
Figure 4.3 shows how the simulation wall time increases significantly as the geometric growth
2500
Min. cell dim., ¢xmin
2480 3 mm 9 mm Analytical 2%
2440
0%
2420
-1%
2400
-2%
2380
2.0 1.9 1.8 1.7 1.6 1.5 1.4 1.3 1.2 1.1
Geometric growth factor, fg
Figure 4.2: BESTEST GC10a slab heat loss sensitivity to mesh detail parameters
35
Min. cell dim., ¢xmin
30
Simulation wall time [s]
3 mm
6 mm
25
9 mm
12 mm
20
15
10
0
2.0 1.9 1.8 1.7 1.6 1.5 1.4 1.3 1.2 1.1
Geometric growth factor, fg
Figure 4.3: BESTEST GC10a simulation wall time sensitivity to mesh detail parameters
The combination of fg and Δxmin selected for the Kiva BESTEST solutions is based on
simulation wall time as well as the values selected by the other secondary mathematical truth
83
standards. These parameters are shown in Table 4.2 along with the selected values by the other
tools.
Table 4.2: Mesh detail parameters selected by each tool for IEA BESTEST analyses
The analytical solution for GC10a represents the heat transfer through the slab into ground
that extends infinitely outward and downward, so the numerical representation must demonstrate
that variation in the boundary distances (i.e., the deep-ground depth and far-field width) yields
The sensitivity to the boundary distances (far-field width and deep-ground depth) are shown
in Figure 4.4. Generally, there is little sensitivity when the boundary distances are both 20 m or
greater.
Figure 4.5 shows that simulation wall time has relatively little sensitivity to the boundary
distances because the geometrically increasing mesh size requires very few additional cells near
the boundary edges for increased boundary distances. Since there is little computation cost for
improved accuracy, both boundary distances were set to 40 m for the GC10a test case.
IEA BESTEST specifies a similar test for the convergence of iterative solvers [90]:
2500
2480 2%
2440
0%
2420
Figure 4.4: BESTEST GC10a slab heat loss sensitivity to boundary distances
35
Far-field width
30 10 m
Simulation wall time [s]
20 m
25 30 m
40 m
20
15
10
0
10 15 20 25 30 35 40
Deep-ground depth [m]
Figure 4.5: BESTEST GC10a simulation wall time sensitivity to boundary distances
Kiva uses the Library of Iterative Solvers (Lis) [106] to solve the system of equations at
each time-step for the fully implicit, Crank-Nicolson, and steady-state solutions. As with any
85
iterative solver, it is necessary to specify a convergence tolerance. For the linear system, Ax = b,
Figure 4.6 shows the sensitivity of the slab heat loss and simulation wall time to tighter
tolerances. By a tolerance of 10−6 there is less than a 0.1% change in results. Tighter convergence
2500 35
Heat flow
Analytical 30
Time 25
2460
20
2440
15
2420
10
2400 5
2380 0
10-4 10-5 10-6
Tolerance
Figure 4.6: BESTEST GC10a slab heat loss and simulation wall time sensitivity to linear solver
tolerance
The previous sections demonstrate that the level of modeling detail for the GC10a test case
The numerical sensitivity results were derived from steady-state solutions of case GC10a.
Extrapolating the results beyond this case implies the following assumptions:
86
• The accuracy for other cases, and other solution schemes, will show similar sensitivity.
• The computation wall time for other numerical schemes will demonstrate similar sensitivity
The selected parameters affecting the numerical approximation are shown in Table 4.3 and
are used for all the remaining test cases and analyses unless otherwise specified.
Parameter Value
Time-step, Δt [min] 60
Minimum cell dimension, Δxmin [mm] 6.0
Maximum cell growth factor, fg 1.20
Far-field width [m] 40
Deep-ground depth [m] 40
Tolerance 10−6
The slab heat loss for the steady-state BESTEST cases are shown in Figure 4.7 and in Table
4.4. It is evident that each of Kiva’s solution methods produce results within the statistical variation
All of Kiva’s solution methods provided accurate results, but, in some cases, the discretiza-
tion parameters had to be relaxed in order to ensure stable solutions. If instabilities arose for a
particular solution, the time-step was reduced first to ensure that there were no changes to the
spatial discretization between solution methods. If a minute-long time-step was not enough to
satiate the stability requirement, subsequent changes were made to the minimum cell dimension
(Δxmin ).
For all the steady-state cases, the explicit solution method required a minute-long time-step
and a minimum cell dimension of 20 mm. The ADE solution method required 30 minute time-
steps for the cases with infinite convective coefficients (i.e., the a and c cases). These additional
87
3000
2500
1500
1000
500
0
GC10a GC30a GC30b GC30c GC60b GC65b
Analytical FLUENT Kiva: Steady-State Kiva: Explicit Kiva: Implicit
TRNSYS MATLAB Kiva: ADE Kiva: ADI Kiva: Crank-Nicolson
time-steps resulted in a proportional increase in simulation wall time. Figure 4.8 (note logarithmic
y-axis) and Table 4.5 show the total simulation time required for each solution method and for each
case. Each Kiva simulation was performed on a 24 core machine with 96 GB of available memory.
The CPUs are Intel Xeon X5660 2.8 Ghz with 12 MB cache.
The IEA BESTEST User’s Manual (Part I of Neymark and Judkoff [90]) specification for
Initially (time < 0) the ground and slab are at 10 ◦ C (50 ◦ F) throughout; at the
beginning of the simulation (time = 0) the zone air temperature steps to 30 ◦ C (86
◦ F) and slab and soil temperatures begin to change accordingly.
and
If possible for the software being tested, the simulation should be run long enough
that there is ≤ 0.1% variation between the floor slab conduction for the last hour
of the last year of the simulation and the last hour of the preceding year of the
simulation; i.e., ≤ 0.1% variation over an interval of 8760 hours.
88
Table 4.4: BESTEST steady-state test case results: Slab heat loss [W]
105
Simulation wall time [s]
104
103
102
101
100
GC10a GC30a GC30b GC30c GC60b GC65b
Kiva: Steady-State Kiva: Explicit Kiva: Implicit
Kiva: ADE Kiva: ADI Kiva: Crank-Nicolson
Figure 4.8: BESTEST steady-state test case simulation wall times (note logarithmic y-axis)
Table 4.5: BESTEST steady-state test case simulation wall times [s]
meaning that, as specified, the steady-state test cases must use a constant initialization method,
simulated over as many years as required to reach the ≤ 0.1% criteria. However, the user’s manual
provides an exception (in section 1.2.1.6) if the software has other initialization methods:
Each of the unsteady solution methods were initialized with the steady-state solution. The
slab heat loss reported is the final value after responding to the defined boundary conditions after
a year’s time of simulation, requiring significantly more time than the purely steady-state solution.
• There is relatively little difference in wall time among BESTEST cases for the same solution.
The ADE solution is the one exception as 30 minute time-steps were required in some cases
• Steady-state simulations are orders of magnitude faster since there is only a single calcula-
• ADE simulations are the fastest, though when the time-step is reduced to 30 minutes, they
• Explicit simulations, due to their increased number of time-steps, have simulation times
near those of the solutions that require an iterative solver (i.e., Implicit and Crank-Nicolson).
• Given the relatively small change in accuracy among each of the solution methods, the ADE
and ADI methods show the most promise for typical annual energy analysis practices.
90
Table 4.6 shows the specification of each of the harmonic test cases. For each harmonic test
Nday − 15 Nhour − 4
Tdb,a = (10 ◦ C) − (6 ◦ C) · cos · 2 · π − (2 ◦ C) · cos ·2·π (4.3)
365 24
where Nday is the N th day of the calendar year, and Nhour is the N th hour of the day (0 =
midnight).
Slab dimensions hint hext Ground depth Far-field width Soil conductivity
Case
[m x m] [W/m2 ·K] [W/m2 ·K] [m] [m] [W/m·K]
GC40a 12 x 12 ∞ ∞ 30 20 1.9
GC40b 12 x 12 100 100 15 15 1.9
GC40c 12 x 12 7.95 ∞ 15 8 1.9
GC45b 36 x 4 100 100 15 15 1.9
GC45c 36 x 4 7.95 ∞ 15 8 1.9
GC50b 80 x 80 100 100 15 15 1.9
GC55b 12 x 12 100 100 2 15 1.9
GC55c 12 x 12 7.95 ∞ 5 8 1.9
GC70b 12 x 12 7.95 11.95 15 15 1.9
GC80b 12 x 12 100 100 15 15 0.5
GC80c 12 x 12 7.95 ∞ 15 8 0.85
Other characteristics similar to all harmonic test cases include the following:
As specified by the IEA BESTEST User’s Manual (Part I of Neymark and Judkoff [90]) for
If possible for the software being tested, the simulation should be run long enough
that there is <= 0.1% variation between the annual floor slab conduction for the
last year of the simulation and the preceding year of the simulation.
Reducing the number of years required to initialize steady periodic conditions can significantly
decrease computation time. BESTEST specifies that all temperatures in the ground domain be
initialized to a spatially-constant temperature by default, but it allows room for other initialization
methods as well:
This section explores the advantages and disadvantages of different preconditioning (referred
to here as initialization) methods. There are three traditional initialization methods implemented
in Kiva:
(2) Kusuda: A one-dimensional analytical solution developed by [75] that provides tempera-
(3) Steady-state: A steady-state solution scheme initializes the temperatures with the first
in all dimensions.
Kiva’s other initialization method, called the accelerated method, simulates the ground do-
main prior to the first time-step, t = 0. This presimulation period is itself initialized using a
(1) A period of long time-step (on the order of days to months), implicit calculations. Since
the fully implicit method is unconditionally stable and physically realistic, these long time-
92
steps can be used to coarsely capture some of the thermal history driven by the annual
weather cycle.
(2) A subsequent period of normal time-step calculations prior to capture some of the shorter
The accelerated initialization method begins with a coarse timestep discretization and tran-
sitions to a finer discretization near the beginning of the simulation, similar to the way the spatial
Figure 4.9 shows the GC40a slab heat loss for the first year of simulation. Four traditional
initialization methods (Kusuda, steady-state, and two different constant temperatures) are com-
pared with an accelerated method and the “ideal” result, which is the steady periodic solution
after several years of simulation (i.e., the net result is within 0.1% of the previous year’s heat loss).
Figure 4.10 shows how the annual heat loss for each initialization method approaches the “ideal”
steady periodic value over seven years (the percent difference is shown in Table 4.7).
5000
Kusuda
4500 Constant temperature (10 ± C)
Steady-state
Slab heat loss [W]
4000 Accelerated
Ideal
3500 Constant temperature (30 ± C)
3000
2500
2000
1500
1000
0 1000 2000 3000 4000 5000 6000 7000 8000
Hour of year
Figure 4.9: BESTEST GC40a comparison of initialization methods: First year slab heat loss
30
26 10%
24
0%
22
-10%
20
Kusuda
Constant temperature (10 ± C) -20%
18 Steady-state
Accelerated
16 -30%
Ideal
Constant temperature (30 ± C)
14 -40%
1 2 3 4 5 6 7
Year
Figure 4.10: BESTEST GC40a comparison of initialization methods: Annual slab heat loss for
seven years after initialization
• The accelerated method is nearly identical to the “ideal” result, with only a slight deviation
• Of the four traditional methods, the steady-state initialization provides the best approxima-
tion of the “ideal” case for the first year, but the constant (10◦ C) and Kusuda initializations
respond faster to the dynamics and provide a better approximation in subsequent years.
• By the second year, the constant (10◦ C) method is the best approximation among the
selected initial temperature. Initializing to 30◦ C instead produces the worst approximation.
• Although the Kusuda method allows for spatially varying initial temperatures and may
does a poor job of approximating the initial conditions near the building.
The acceleration method exemplified above is one of many possible variants. Other variants
Constant Constant
Kusuda Steady-state Accelerated
temperature (10◦ C) temperature (30◦ C)
Year
1 22.1% 18.7% 7.49% 0.01% -32.3%
2 2.09% 2.07% 3.91% -0.19% -19.0%
3 1.02% 1.02% 2.69% -0.14% -14.5%
4 0.65% 0.65% 2.00% -0.10% -11.4%
5 0.46% 0.46% 1.54% -0.06% -8.96%
6 0.35% 0.36% 1.21% -0.02% -7.03%
7 0.28% 0.29% 0.96% 0.00% -5.50%
• Δtaccel : The time-step duration of the acceleration period (1 day, 1 week, 1 month, etc.)
• twarm-up : The total duration of the warm-up period (directly before primary simulation on
A wide range of potential accelerations were simulated to compare accuracy and relative
simulation time. This range represents all combinations of the following parameter values (for the
In Kiva, the presimulation duration is not input by the user but is, instead, calculated
internally as follows:
where Naccel is the number of acceleration time-steps used during the presimulation period and is
Figure 4.11 shows a comparison of the relative accuracy and wall time of all the initialization
methods tested. In this figure, the relative difference in annual heat loss is the percent difference
between the first year of simulation and the “ideal” steady periodic results. The additional sim-
ulation wall time represents the difference in time relative to the fastest simulation. The initial
results revealed that the warm-up period, twarm-up , has negligible impact on the accuracy of the
results and added considerable time to the simulation. Variations of the warm-up period are not
30%
annual slab heat loss
Relative difference in
20%
10%
0%
-10%
-20%
-30%
Figure 4.11: BESTEST GC40a comparison of first year slab heat loss for traditional and accelerated
initialization methods
• The acceleration method is consistently more accurate than the traditional initialization
methods.
• There is not an acceleration time-step that is clearly more accurate than the others, though
daily acceleration time-steps showed significant additional wall time without a consistent
improvement in accuracy.
• There is no clear trend between the duration of the presimulation period and accuracy. By
three months a majority of the thermal history has been captured by the presimulation.
• The most accurate presimulation initialization was found at weekly acceleration time-steps
for a three month period. This may not be the best set of presimulation parameters for all
Each of the Kiva solution methods was used to simulate the BESTEST harmonic test cases.
Each simulation leveraged an accelerated initialization followed by a full year of warm-up to sat-
isfy the output requirements for allowable variation in slab heat loss between consecutive years.
While the accelerated initialization method allows Kiva to simulate only two years, the secondary
The results for the harmonic BESTEST cases are shown in Figure 4.12 and in Table 4.8. It
is evident that each of the solution methods in Kiva produce results within the statistical variation
Table 4.8: BESTEST harmonic (unsteady) test case results: Annual slab heat loss [MWh]
GC40a GC40b GC40c GC45b GC45c GC50b GC55b GC55c GC70b GC80b GC80c
TRNSYS 23.0 22.1 18.6 32.8 27.0 277.9 35.1 20.8 17.4 6.0 9.2
FLUENT 22.8 21.9 18.6 32.5 26.9 278.0 34.9 20.7 17.4 5.9 9.1
MATLAB 23.6 22.5 18.9 33.5 27.4 281.4 35.5 21.0 17.6 6.2 9.3
Kiva: ADE 23.3 22.4 18.9 33.4 27.5 279.5 35.2 20.9 17.6 6.1 9.3
Kiva: Explicit 23.0 22.2 18.8 33.0 27.3 278.4 35.0 20.8 17.6 6.0 9.2
Kiva: ADI 22.5 21.9 18.6 32.5 27.1 276.2 34.7 20.6 17.5 6.0 9.2
Kiva: Implicit 23.3 22.4 18.9 33.4 27.5 279.5 35.2 20.9 17.6 6.1 9.3
Kiva: Crank-Nicolson 23.3 22.4 18.9 33.4 27.5 279.5 35.2 20.9 17.6 6.1 9.3
IEA BESTEST procedures specify that annual time-series results must be provided for each
of the GC40 cases (a, b, and c). In a plot of annual results (Figure 4.13), it can be difficult to
97
40
30
25
20
15
10
0
GC40a GC40b GC40c GC45b GC45c GC50b GC55b GC55c GC70b GC80b GC80c
TRNSYS MATLAB Kiva: Explicit Kiva: Implicit
FLUENT Kiva: ADE Kiva: ADI Kiva: Crank-Nicolson
Figure 4.12: BESTEST harmonic (unsteady) test case annual slab heat loss (results for GC50b are
divided by 10 to maintain a consistent scale)
notice some of the more subtle differences among the solution results at the time-step level. It is
more informative to compare the annual and daily time-scale characteristics of the solutions. As
the time-series solutions are driven by the sinusoidal variation in exterior temperature (Equation
4.3), the solution for slab heat loss takes a similar form:
• Φyear : The annual phase shift of the slab heat loss (relative to the phase of the annual
temperature variation)
• Φday : The daily phase shift of the slab heat loss (relative to the phase of the daily temper-
ature variation)
3400
3200
Slab heat loss [W]
3000
2800
2600
2400
2200
2000
0 1000 2000 3000 4000 5000 6000 7000 8000
Hour of year
TRNSYS MATLAB Kiva: Explicit Kiva: Implicit
FLUENT Kiva: ADE Kiva: ADI Kiva: Crank-Nicolson
Figure 4.13: BESTEST GC40a harmonic (unsteady) test case hourly results
The values of each of the five harmonic parameters for the three GC40 cases are shown in
Table 4.9. The annual slab heat loss never exceeds a 5% difference between solutions for a given
case. For each case any deviation from the median parameter values is indicated by the degree
of background shading (darker indicates more deviation). The only outliers occur for the daily
parameters, Q̇day and Φday , in the ADE solutions, which appears be less accurate at shorter time
scales.
Similar to the stability issues mentioned for the steady-state tests, the explicit solution
method required the same restrictions on time-step and cell size (1 minute time-steps and 20
mm minimum cell size) to ensure stability. The ADE solution method required shorter time-steps
for several cases involving infinite convective coefficients (the a and c cases). The stable time-steps
used in verification for the ADE solutions are listed in Table 4.10.
99
Table 4.9: BESTEST GC40 harmonic parameters for each solution (shading indicates deviation
from the median value for each case)
¯
Q̇ Q̇year Φyear Q̇day Φday
Case/Solution
[W] [W] [days] [W] [hours]
GC40a:
TRNSYS 2,629 447 16.6 10.2 8.3
FLUENT 2,598 435 17.2 8.2 8.1
MATLAB 2,695 465 16.4 13.4 8.2
Kiva: ADE 2,664 453 16.4 13.7 2.3
Kiva: Explicit 2,631 443 16.9 10.6 9.4
Kiva: ADI 2,565 425 17.5 7.6 9.7
Kiva: Implicit 2,660 452 16.4 11.6 8.3
Kiva: Crank-Nicolson 2,660 453 16.4 12.6 8.9
GC40b:
TRNSYS 2,523 412 17.9 6.4 9.3
FLUENT 2,504 405 18.3 5.6 8.9
MATLAB 2,570 424 18.0 8.1 9.3
Kiva: ADE 2,554 416 18.1 14.2 7.7
Kiva: Explicit 2,532 409 18.4 6.8 10.4
Kiva: ADI 2,497 399 18.7 5.2 10.4
Kiva: Implicit 2,554 415 17.9 6.9 9.3
Kiva: Crank-Nicolson 2,554 415 17.9 7.6 10.0
GC40c:
TRNSYS 2,129 323 21.9 2.4 10.1
FLUENT 2,123 319 22.2 2.2 9.8
MATLAB 2,154 329 22.0 3.0 10.1
Kiva: ADE 2,156 325 22.5 1.1 8.8
Kiva: Explicit 2,141 321 22.4 2.4 11.4
Kiva: ADI 2,124 316 22.7 1.9 11.3
Kiva: Implicit 2,153 325 22.0 2.6 10.4
Kiva: Crank-Nicolson 2,154 325 22.0 3.0 11.1
Figure 4.14 and Table 4.11 show the total simulation time required for each solution method
and for each case. The following observations can be taken from these results:
100
• ADE simulations are again the fastest, though when the time-step is reduced, there is a
• Explicit simulations, due to their increased number of time-steps, have simulation times
near those of the solutions, which require a linear algebra solver (i.e., Implicit and Crank-
Nicolson).
• Given the relatively small change in accuracy between each of the solution methods, the
ADE and ADI methods remain the most promising for typical annual energy analysis
practices.
30000
Simulation wall time [s]
25000
20000
15000
10000
5000
0
GC40a GC40b GC40c GC45b GC45c GC50b GC55b GC55c GC70b GC80b GC80c
Kiva: ADE Kiva: Explicit Kiva: ADI Kiva: Implicit Kiva: Crank-Nicolson
Figure 4.14: BESTEST harmonic (unsteady) test case simulation wall times
The results of the BESTEST comparison cases indicate that there are no statistical outliers
with regard to total slab heat loss. Kiva’s capability to utilize several different finite difference
101
Table 4.11: BESTEST harmonic (unsteady) test case results: Simulation wall time [s]
GC40a GC40b GC40c GC45b GC45c GC50b GC55b GC55c GC70b GC80b GC80c
Kiva: ADE 3,488 1,618 4,599 1,610 4,283 1,963 1,286 3,637 1,630 1,607 3,180
Kiva: Explicit 15,017 12,233 10,528 13,243 10,411 17,063 7,527 8,242 12,376 12,201 10,387
Kiva: ADI 3,832 3,617 3,249 3,547 3,178 4,350 2,442 2,632 3,609 3,595 3,251
Kiva: Implicit 17,991 21,863 16,488 22,152 15,984 26,358 16,875 13,878 22,969 17,593 14,572
Kiva: Crank-Nicolson 17,771 20,344 15,169 20,051 15,184 24,728 15,018 13,314 21,623 17,567 13,712
solution methods makes it the first tool to directly compare multiple methods under the same
computational framework, thus ensuring that differences among the solutions are isolated from other
programming differences. Each method was evaluated regarding numerical accuracy, computation
In general, a fully implicit solution provides the most accurate results as it is not subject
to some of the stability issues associated with the explicit schemes or the sub-time-step approxi-
mations of the ADI scheme. The Crank-Nicolson scheme provides a better approximation of time
differencing than the fully implicit scheme though it is still subject to unrealistic oscillations.
Although the ADI solution tended to underapproximate the slab heat loss relative to the
other solutions, it required less than one quarter of the computation time of the other implicit
schemes. It is also superior to the ADE scheme in that it has less stringent stability criteria.
Finally, the investigation into initialization methods revealed that each of the traditional
methods (Kusuda, constant temperature, and steady-state) required seven or more years to arrive
at a steady harmonic response to the driving exterior temperature variation. Several variations of
accelerated initializations were tested and were all were found to provide a more accurate approxi-
mation of the first year heat loss than the traditional methods. Of these accelerated initializations,
those with a short (three month) initialization period and relatively long acceleration time-steps
(one week) provided a favorable trade-off between accuracy and the additional required simulation
time.
On the whole, the simulation time required to perform foundation heat transfer calculations
can be significantly reduced, relative to traditional approaches, without a significant loss in accuracy.
102
For the GC40a test case, a two-year simulation using the ADI solution method with accelerated
initialization required 56.5 minutes in contrast to the projected 18.25 hours required for a seven-
year simulation using the fully implicit method with a traditional initialization method (a nearly
Chapter 4 demonstrated how computation time is reduced through the selection of the Alter-
nating Direction Implicit (ADI) solution scheme and the accelerated initialization method, without
significantly affecting the veracity of Kiva’s calculations. Chapter 4 also established discretization
independence for the solutions within the context of the IEA BESTEST procedures.
Further reductions in computation time are possible, but they require a more delicate balance
with computational accuracy. Reducing the problem to two-dimensions is a common approach and
Many efforts have been taken to approximate three-dimensional effects using an equivalent
two-dimensional geometry. Walton [114] noted that improper application of two-dimensional calcu-
lations can result in up to 50% error, relative to three-dimensional calculations. Only a few authors
It is generally accepted that the heat transfer from building foundations is strongly two-
dimensional near the perimeter of the foundation. However, there is no consensus on how the
two-dimensional simulation results should be scaled to represent the heat transfer within the ac-
tual, three-dimensional context. This chapter aims to resolve the disparity between two- and
three-dimensional simulation results by exploring the implications of different two dimensional ap-
The sections below describe several established methods of scaling two-dimensional simulation
results. Each approximation method uses the same vertical dimensions as the three-dimensional
definition, but differs in how it defines the lateral dimension of the foundation floor, Δxf , in Figure
5.1.
z
Δxf
Many authors refer to Δxf /2 as the “half-width,” which is the dimension used for the foun-
5.1.1 Perimeter
Many authors provide results normalized by foundation perimeter. Scaling by the perimeter
is a logical first choice since the magnitude of the foundation heat transfer is stronger near the
where
105
• q̇cart,f is the two-dimensional heat flux [W/m2 ] through the foundation floor calculated in
• q̇cart,w is the two-dimensional heat flux [W/m2 ] through the foundation wall calculated in
the “Corner Correction Method.” This method introduces corner correction factors that relate the
heat transfer at the corners of the foundation to the heat transfer of the center portions of the
Corner correction factors were calculated from a regression based on 3,024 numerical simu-
lations of rectangular basements. The result is a table of 19 regression coefficients that correlate
the correction factors to six independent variables related to insulation configuration, soil thermal
Scaling by the perimeter does not strictly define the value of foundation floor width, Δxf ,
needed to create the two-dimensional domain. Both Bazjanac et al. [13] and Deru [39] defined the
foundation width as the narrow dimension of a rectangular foundation. However, instead of using
the relationship defined by Equation 5.1, Bazjanac et al. [13] weighted the floor results by the area
of discrete rectangular annuli corresponding to different locations along the floor width.
Bahnfleth [9] and Bahnfleth and Pedersen [11] simulated a series of three-dimensional square,
rectangular, and L-shaped slabs with a range of foundation areas and perimeter lengths. They
106
demonstrated that heat transfer does not vary linearly with the perimeter of the foundation, and
indicated that the area of the foundation also plays an important role. They concluded that the
average heat flux through rectangular and L-shaped slabs could be calculated as a function of the
where
The area-to-perimeter ratio also appears in the study of fluid flow—another diffusion driven
phenomenon. Fluid flowing through channels is often characterized by a single characteristic length,
A
Dh = 4 , (5.4)
P
where A is the cross-sectional area of the channel and P is the wetted perimeter.
The hydraulic diameter defines the effective diameter of a circular channel, allowing any
channel to use similar flow calculations. It is logical that that the same relationship applies to heat
Extrapolating Equation 5.3 beyond rectangular and L-shaped slabs implies that foundation
heat transfer can be calculated by scaling heat fluxes by foundation area. Therefore, heat trans-
fer from geometrically complicated shapes can be calculated from simpler shapes with the same
characteristic length.
107
method themselves, a number of authors have utilized the A/P characteristic length for scaling
two-dimensional results [95, 24, 25]. The A/P characteristic length leads to two possible shapes
that can be simulated in two dimensions: one for an axisymmetric cylindrical coordinate system
of radius
A
r=2 (5.5)
P
With Δxf = 2r, the total representative three-dimensional heat transfer for the circular
method is as follows:
where
• q̇cyl,f is the two-dimensional heat flux [W/m2 ] through the foundation floor calculated in a
• q̇cyl,w is the two-dimensional heat flux [W/m2 ] through the foundation wall calculated in a
A lw w
= = , (5.7)
P 2l + 2w 2 + 2w
l
and as l → ∞,
A w
→ , (5.8)
P 2
which means any foundation shape can be represented by an infinite rectangle of width
A
w=2 . (5.9)
P
With Δxf = w, the total representative three-dimensional heat transfer for the infinite
where
109
• q̇cart,f is the two-dimensional heat flux [W/m2 ] through the foundation floor calculated in
• q̇cart,w is the two-dimensional heat flux [W/m2 ] through the foundation wall calculated in
Chuangchid [24] and Chuangchid and Krarti [25] verified that both circular and infinite
method, transforms any three-dimensional foundation geometry into a rounded rectangle foundation
of the same perimeter and area. The rounded rectangle allows entire problems to be solved using
the superposition of heat transfer calculated using a Cartesian coordinate system (for the straight
sides) and a cylindrical coordinate system (for the rounded ends). From the area, A, and perimeter,
a = P − P 2 − 4πA /π, (5.11)
and
With Δxf = a for both coordinate systems, the total representative three-dimensional heat
transfer is as follows:
a
b
Figure 5.4: Dimensions of a rounded rectangle with equivalent area and perimeter
Walton verified his rounded rectangle approximations for uninsulated rectangular basements
and slabs under a steady-state solution, while results for X-shaped and H-shaped basements showed
The rounded rectangle approximation method uses two characteristic lengths, a and b, which
allow for an additional degree of freedom in determining the representative heat transfer. In this
case, the additional degree of freedom requires two independent two-dimensional simulations, one for
each coordinate system. However, because the simulations are independent, they can be calculated
on parallel processors without significantly more computation time than a single simulation.
Bahnfleth and Pedersen [11] and Walton [114] explain that two-dimensional results can better
approximate three-dimensional results when scaled by both area and perimeter of the shape, not just
perimeter. Bahnfleth and Pedersen [11] also introduced the characteristic length, A/P , which can
approximation method developed thus far accounts for concave foundation shapes. Even the corner-
correction method developed by Beausoleil-Morrison et al. [16] only provides adjustments for convex
rectangular corners.
111
Shape cavities are defined relative to the shape’s convex hull. The convex hull is the smallest
convex polygon that envelops the original polygon. The spatial difference between the original
shape (Figure 5.5(a)) and the convex hull (Figure 5.5(b)) defines the set of cavities (Figure 5.5(c)).
While heat transferred from convex edges is largely unaffected by other edges, the ground
within cavities is warmer because of the additive heat flow from neighboring concave edges. This
results in reduced heat transfer from the foundation. The influence of neighboring concave edges
on heat transfer can be observed in Figure 5.6 for three-dimensional simulation results. Heat flux
is higher through the ground near the concave slab corner, relative to heat fluxes further from the
corner. Consequently, heat flux is lower through the slab in proximity to the warmer ground.
The degree of heat transfer reduction within cavities depends on several factors beyond the
shape itself. These influential factors are illustrated by isolating the concave effects. The following
sections describe two separate types of interactive effects that occur between concave foundation
edges.
The first example of interaction between concave edges occurs when two parallel edges come
within close proximity of each other. This effect can be isolated by simulating heat transfer from
112
Figure 5.6: Plan view cross-section at grade height of heat flux near a concave slab corner (exterior
grade in the upper left)
two infinitely long rectangles separated by a distance, p (Figure 5.7). This arrangement can be
simulated in two-dimensions with two planes of symmetry. This exercise demonstrates how the
The extent of the impact of opposing parallel edges will depend on the actual perimeter of
opposing edges. In the example of infinite rectangles, heat transfer reduces to zero as the distance
between the edges closes. Actual foundation shapes are enclosed, which means that the effect
applies to less than half of the perimeter, at most (in the case of long U-shaped foundations).
Parallel concave edges are simulated for a combination of area-to-perimeter ratios, soil con-
ductivity, and insulation levels. All simulations are steady-state with an indoor-outdoor tempera-
ture difference of 20◦ C. Unless stated otherwise, the foundations are slabs-on-grade with exterior
horizontal insulation extending roughly one meter from the perimeter. The results presented in
113
A/P p A/P
Figures 5.8–5.10 demonstrate the sensitivity of opposing parallel edges for a single example foun-
dation.
100%
Percent of convex heat loss
80%
60%
40%
A=P [m]
1.5
20% 4.6
7.6
10.7
13.7
0%
1 2 3 4 5
Distance between parallel boundaries [m]
Figure 5.8: Reduction in heat loss between two parallel concave edges over a range of area-to-
perimeter ratios (ksoil = 1.0 W/m·K, R-10 insulation)
114
As the area-to-perimeter ratio increases, the heat loss reduction relative to a convex (i.e.,
with no opposing edge) infinite rectangle decreases. Therefore the interaction effect has a higher
impact for larger foundations. Smaller foundations see a less gradual reduction and are relatively
unaffected until the two edges are very close, at which point heat transfer drops steeply.
100%
Percent of convex heat loss
80%
60%
40%
ksoil [W/m ¢K]
0.25
0.50
20% 0.75
1.00
1.25
1.50
0%
1 2 3 4 5
Distance between parallel boundaries [m]
Figure 5.9: Reduction in heat loss between two parallel concave edges over a range of soil conduc-
tivity (A/P = 4.6 m, R-10 insulation)
As the soil conductivity increases, the interaction between the two edges becomes stronger.
Higher soil conductivity allows heat to flow further from the foundation, causing a greater impact
With increased levels of insulation, the path of heat flow is driven further from foundation
perimeter, where there is a higher potential for interaction with the opposing edge. As a result,
The shape of the curves in Figures 5.8–5.10 represents reduced heat loss as the two edges
grow closer together. The curves show a noticeable change where the distance between the two
edges is less than the width of the insulation. At long distances there is little effect, but within
115
100%
60%
40%
R-value
R-0
20% R-5
R-10
R-15
R-20
0%
1 2 3 4 5
Distance between parallel boundaries [m]
Figure 5.10: Reduction in heat loss between two parallel concave edges over a range of insulation
R-values [hr·ft2 ·R/Btu] (ksoil = 1.0 W/m·K, A/P = 4.6 m)
the last several meters the heat transfer reduction drops drastically toward zero. The heat transfer
reduction is related to the magnitude of heat flux from the foundation, which is strong near the
perimeter and drops quickly to zero beyond a few meters. Though the distance where the magnitude
drops to zero is difficult to quantify precisely, it is similar to the concept of a boundary layer in
fluid mechanics. The boundary of heat flow near the perimeter of the foundation progressively
The profile of this boundary layer can be calculated for any vertical cross-section extending
from the perimeter of the foundation. The boundary layer is defined by the vertical heat flux along
the grade plane. Figure 5.11 illustrates the boundary layer profile for a slab with exterior horizontal
insulation.
The integral of the boundary layer profile from the foundation perimeter yields the function,
δ (x). This function is normalized to represent the fraction of total heat transfer within any distance,
116
z
. x
qz(x)
Figure 5.11: Illustration of a heat flux boundary layer profile for a slab with exterior horizontal
insulation
x, of the perimeter. In the limit as x approaches infinity, the value of δ (x) approaches unity:
Figure 5.12 shows examples of the normalized boundary layer profile, q̇z (x), and the inte-
The boundary layer interaction of the two opposing edges causes reduced heat transfer pro-
portional to the value of the integrated boundary layer function at a given distance. For any shape
cavity, the degree of interaction is related to the integrated boundary layer profile. The next section
Depending on the angle of a corner, the boundary layer will be more or less constrained.
Sharper angles will cause a higher level of interaction within a cavity. As Kiva is only capable of
117
100%
80%
q_ z (x)
60%
Percent of total boundary heat loss
40%
20%
0%
100%
Z x
80% ±(x) = q_ z (x)dx
0
60%
40%
20%
0%
0 1 2 3 4 5
Distance from perimeter, x [m]
Figure 5.12: Example boundary layer profile for a slab with exterior horizontal insulation (ksoil =
1.0 W/m·K, A/P = 4.6 m, R-10 insulation)
simulating rectilinear foundation shapes, the effects in concave corners can only be demonstrated
The effect of corners can be isolated by defining shapes with the same area and perimeter,
but with a different number of corners. The range of shapes with corners is somewhat limited
as complex shapes with many corners create several instances of near-field mesh refinements that
can drastically increase the computational requirements. For this reason, only four shapes are
selected to demonstrate the effect of concave corners on foundation heat transfer. These shapes are
Heat transfer from slabs of each of these shapes was simulated under the same range of area-
to-perimeter ratios, soil conductivities, and insulation levels that were simulated for the parallel
118
13d
d
3d
d
6d
d
d
4d A = 13d2
P = 28d
Figure 5.13: Shapes with equal area and perimeter but different number of concave corners
concave edge example. The proportion of the shapes remains constant even with different area-
to-perimeter ratios. For a given area-to-perimeter ratio, A/P , the shape thickness, d, is defined
as
28
d= [A/P ] . (5.15)
13
Figures 5.14–5.16 show the sensitivity of heat loss reduction to changes in area-to-perimeter
ratio, soil conductivity, and insulation R-value. Results are presented for each concave shape
The jump from zero corners (the rectangle shape at 100%) to a single corner (the L shape) is
much larger than the difference between one corner and two corners (the T shape). By four corners
(the + shape) the impact of additional corners continues to drop. Although the fraction of the
perimeter affected by the corners increases proportionally to the number of corners, the heat flux
Unlike the opposing parallel edges, the reduction effect for corners increases as the area-
to-perimeter ratio decreases. For lower area-to-perimeter ratios, the interaction effect for corners
119
100%
96%
94%
92%
90%
0 2 4 6 8 10 12 14
A=P [m]
Figure 5.14: Reduction in heat loss due to corners over a range of area-to-perimeter ratios (ksoil =
1.0 W/m·K, R-10 insulation)
100%
Percent of convex heat loss
98%
96%
94%
92%
90%
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
2
ksoil [W/m ¢K]
Figure 5.15: Reduction in heat loss due to corners over a range of soil conductivity (A/P = 4.6 m,
R-10 insulation)
120
100%
96%
94%
92%
90%
0 5 10 15 20
2
Insulation R-value [hr ¢ft ¢R/Btu]
Figure 5.16: Reduction in heat loss due to corners over a range of insulation R-values (ksoil = 1.0
W/m·K, A/P = 4.6 m)
between shorter edges impacts a larger portion of the perimeter causing a larger heat transfer
reduction. For larger area-to-perimeter ratios, the corners represent a smaller portion of the total
perimeter, and therefore the heat transfer reduction is less impactful. In the limit as the area-to-
The impact of increasing soil conductivity and insulation R-value for corners is similar to the
impact for parallel edges as shown in Figures 5.15 and 5.16, respectively.
The boundary layer profile is amplified near the corner due to interactions between the two
intersecting concave edges. Figure 5.17 illustrates example cross-sections for both corner bisector
The boundary layer heat flux profile along a corner bisector, q̇z,b (x), can be estimated from
where
121
Figure 5.17: Illustration of corner bisector and unaffected edge boundary profile cross-sections
• the sin (θ/2) accounts for the change in dimensions of the foundation elements along the
bisector,
• the 1 + cos (θ/2) accounts for increased heat flux in the corner due to interacting boundary
layer profiles.
Kiva can only simulate right angles, and cannot confirm this relationship for other angles. Fig-
ure 5.18 validates the estimated bisector boundary layer profile and the calculated two-dimensional,
unaffected boundary layer profile against three-dimensional simulation results. The discrete points
in Figure 5.18 represent the three-dimensional heat flux calculated along each cross-section. The
curves represent the generated unaffected boundary layer profile, and the calculated boundary layer
profile using the relationship in Equation 5.16. There is near-perfect agreement between the three
80%
60%
40%
20%
0%
0 5 10 15 20
Distance from perimeter [m]
The heat transfer reduction near concave corners is related to differences between the bisector
integrated boundary layer profile and the integrated boundary layer profile of an unaffected edge.
Many shapes exhibit both opposing parallel edges and corners. The interactive effects com-
bine as illustrated in Figure 5.19 for an H-shaped slab. The cavities of the H-shape are adjusted
by varying the dimensions of the web (the connector) and flanges (the parallel sides), while main-
taining constant area and perimeter. Figure 5.19 demonstrates the progression of heat transfer
reduction at discrete intervals between convex extremes where the shape has no flanges or no web.
The flange and web thickness, d, is set to 10 meters, and the area-to-perimeter ratio is
13d2
A/P = = 4.64 m. (5.18)
28d
123
R-20, k = 1.50 W/m ¢K R-20, k = 0.25 W/m ¢K
Unins., k = 1.50 W/m ¢K Unins., k = 0.25 W/m ¢K
100%
95%
85%
80%
75%
70%
65%
60%
55%
The four curves in Figure 5.19 represent bounding values of soil conductivity and insulation level
On the left-hand side of Figure 5.19, there is relatively little heat transfer reduction where
the flanges are not long enough to form an unaffected boundary layer. For the cases with narrow
boundary layers (e.g., uninsulated with low soil conductivity), the relative heat transfer settles at
a constant value for shorter flanges, while shapes with wide boundary layers require longer flanges
before an unaffected boundary layer is formed. In the case of a high insulation value and high soil
conductivity, the unaffected boundary layer is not established before the interaction between the
The established approximation methods described in Section 5.1 fail to account for heat
transfer reduction in cavities and thus predict the same heat loss for each shape on a given curve in
Figure 5.19. In the extreme case with no web, there is a potential 40% error in the approximations
124
relative to the three-dimensional solution. It is not coincidence that the opposing parallel edges in
The integrated boundary layer profile can be used to estimate heat transfer reduction for a
foundation shape with cavities. The profile can be calculated easily within the existing computa-
tional framework, and it contains all of the information about the domain which influences heat
transfer within shape cavities. There is no need to develop regressions for all of the variables which
impact the heat transfer reduction within cavities because the profile is specific to the combination
of foundation design characteristics and soil thermal properties defined for a given model.
Bahnfleth and Pedersen [11] demonstrated that as the area-to-perimeter ratio of a foundation
increases, the average heat flux from the foundation decreases. The concept behind the boundary
layer adjustment method is to adjust the dimensions of parallel edges and concave corners, based
on the integrated boundary layer profile, to produce a new adjusted area-to-perimeter ratio. The
adjusted area-to-perimeter ratio defines the dimensions for a representative two-dimensional domain
where
The value of Pv for each concave feature of the foundation shape depends on the integrated
boundary layer profile function. If thermal properties are linear, then the boundary layer profile
125
simulation with an arbitrary non-zero temperature difference. The values of vertical heat flux are
integrated along the grade plane to generate a mapping of the percent total heat flux to the distance
from the perimeter at each discrete cell. The map is as coarse as the spatial discretization in the
lateral direction. The value of the integrated boundary layer profile, F , at any distance, x, is eval-
uated using a linear interpolation between mapped values. The inverse of the integrated boundary
layer profile function, δ −1 (F ), also uses the same map to provide the distance, x, corresponding to
For opposing parallel edges, the edge perimeters are adjusted as follows:
Pv = 2a [1 − δ (p)] . (5.20)
where p is the distance between the edges, and a is the common length of the parallel edges as
p
a a
For concave corners, the perimeter is adjusted based on the integrated boundary layer profile
of the bisector described in Equation 5.17, the angle of the corner, θ, and the lengths, A and B, of
A chamfer at a distance, xc , from the corner along the bisector, as shown in Figure 5.21,
θ b
The value xc is selected such that the total heat flux beyond the chamfer is the same as the
total heat flux from an unaffected edge. The total relative heat flux along the bisector is
1 + cos (θ/2)
lim δb (x) = . (5.21)
x→∞ sin (θ/2)
1 + cos (θ/2)
δb (xc ) = − 1. (5.22)
sin (θ/2)
δ −1 (1 − tan (θ/4))
xc = . (5.24)
sin (θ/2)
127
The distance, d, from the angle to the points where the chamfer intersects the edges is
calculated as
xc
d= . (5.25)
cos (θ/2)
The final dimensions of the edges used depend on the lengths of the edges relative to d:
and
The verification testing in Chapter 4 was based solely on the IEA BESTEST test case
specifications for slab-on-grade foundations. These test cases are constrained to idealized conditions
and do not cover the breadth of modeling capability in Kiva. The two-dimensional approximation
methods discussed in this chapter are tested in a more realistic context under a wider range of
parameters.
All simulations performed to test approximation methods use the set of conditions described
in Table 5.1. All three-dimensional simulations are modeled using planes of symmetry where they
Several other parameters are varied between tests to highlight potential sensitivities in results.
5.4.1 Climates
Approximation methods are tested using realistic weather data representing a range of cli-
mates. The selection of weather locations is based on the range of outdoor dry-bulb temperatures
128
Table 5.1: Approximation method testing conditions
Description Value
Foundation wall thickness 0.3 m
Slab thickness 0.2 m
Interior temperature 22.2 ◦ C
Soil conductivity 0.864 W/m·K
Soil density 1,510 kg/m3
Soil specific heat 1,260 J/kg·K
Surface absorptivity 0.8
Deep-ground boundary depth 40 m
Deep-ground boundary condition Zero-flux
Far-field boundary width 40 m
Minimum cell dimension 0.006 m
Cell growth coefficient 1.20
Time-step duration 60 minutes
Numerical method ADI
Initialization acceleration time-step duration 1 week
Initialization acceleration period duration 12 weeks
Initialization warm-up duration 1 year
in Typical Meteorological Year (TMY3) data [117]. Figure 5.22 illustrates the ranges of dry-bulb
temperatures selected for the representative hot (Phoenix, Arizona), mixed (Albuquerque, NM),
The approximation method test covers a total of 25 different foundation designs. These
designs represent the three most common foundation types. Insulation configurations are shown
in Figures 5.23, 5.24, and 5.25 for slabs, crawlspaces, and basements, respectively. For each of the
insulated configurations, there are two designs at different insulation levels: R-5 (0.88 m2 ·K/W)
All basements are modeled with an 8-foot (2.4 m) foundation depth, crawlspaces are modeled
with a 4-foot (1.2 m) foundation depth, and slabs are modeled 8 inches (0.2 m) above grade. The
foundation wall footer is 5-feet-8-inches (1.7 m) deep for slabs and crawlspaces, and 9-feet-8-inches
40
30
Dry-bulb temperature [ ± C]
20
10
−10
Hot Mixed Cold
−20
Time of year
Figure 5.22: Annual outdoor dry-bulb temperatures from weather data files for selected climates.
Hot = Phoenix, Arizona; Mixed = Albuquerque, New Mexico; Cold = Anchorage, Alaska
The foundation shapes selected for testing two-dimensional approximation methods represent
The characteristic length, defined by the area-to-perimeter ratio, A/P , describes the relative
size of the foundation. The range of characteristic lengths used in the tests are as follows:
Compactness is a nondimensional characteristic that ranges between zero and one, defined
Table 5.2 demonstrates the compactness of several shapes. Each of the shapes shown in Table
5.2 can be simulated in Kiva with the exception of the equilateral triangle, because it is neither
rectilinear nor described by a two-dimensional coordinate system. The equilateral cross and square
shapes are included in the approximation tests. Two variations of the H-shape are also included
to represent different distances between the opposing parallel edges: a narrow-H with 3 meters (10
feet) between flanges, and a wide-H with 122 meters (400 feet) between flanges.
The final set of shapes used in the approximation method tests includes four shapes simu-
lated in three-dimensional Cartesian coordinates (square, cross, narrow-H, and wide-H), a circle in
Cartesian coordinates.
131
The shapes simulated in three-dimensional Cartesian coordinates are shown in Figure 5.26
Initial simulation results for the mixed climate revealed notable differences in foundation heat
flux among shapes with the same characteristic length. Figure 5.27 shows deviation from the mean
heat flux for uninsulated foundation walls and floors, across all shapes.
where q̇f and q̇w are the heat fluxes through the floor and wall, respectively. Although the
crawlspace and foundation heat flux deviates more through floors than through walls, it
132
is also a lower magnitude. Depending on the dimensions of the foundation, this deviation
• Heat flux is less sensitive to foundation shape for shallower foundations. The area-to-
perimeter ratio, which Bahnfleth and Pedersen [11] demonstrated characterizes heat trans-
fer through slab foundations, does not extend as effectively to crawlspace and basement
foundations.
• For crawlspaces and basements, there are two independent trends worth noting:
◦ Shapes with higher compactness have proportionally higher heat flux than the infinite
◦ As explained for Figure 5.19, shapes with concavities have lower heat fluxes than the
Circle 1
Square π
4 ≈ 0.785
Equilateral triangle π
√
3 3
≈ 0.605
5π
Equilateral cross 36 ≈ 0.436
7π
Equal-spaced H 64 ≈ 0.344
Infinite
0
rectangle
When comparing different insulation designs, engineers are often more concerned with rel-
ative differences among designs rather than absolute heat transfer from foundations. Figure 5.28
illustrates the reduction in heat flux for different insulation designs relative to the uninsulated
foundations portrayed in Figure 5.27. Floor heat flux reductions are not shown for crawlspaces and
basements because their magnitudes are substantially lower than heat flux through walls. Similar
heat flux reductions are demonstrated among all shapes, with a slight disparity for shapes with
higher compactness.
134
Figure 5.26: Shapes simulated in three dimensions for approximation method testing. Each plot is
200 m × 200 m. A = Area, C = Compactness.
The four two-dimensional approximation methods tested are listed in Table 5.3.
Characteristic Coordinate
Method
Length(s) System(s)
Circle A/P Cylindrical
Infinite rectangle A/P Cartesian
Cartesian and
Rounded rectangle a, b
Cylindrical
Boundary layer adjustment [A/P ]adj Cartesian
135
Slab Crawlspace Basement
Wall Floor
20%
A=P = 1.5 m
10%
0%
-10%
Deviation from mean heat flux
-20%
20%
A=P = 4.6 m
10%
0%
-10%
-20%
20%
A=P = 13.7 m
10%
0%
-10%
-20%
Circle Semi-Infinite Rectangle Cross
Square Wide-H Narrow-H
Figure 5.27: Comparison of heat flux from a range of uninsulated foundation shapes
The approximation methods are compared to three-dimensional simulation results for a com-
bination of:
• 3 climates,
• 25 foundation designs,
• 4 foundation shapes.
136
Square Narrow-H Circle
Cross Wide-H Semi-Infinite Rectangle
100%
A=P = 1.5 m
80%
60%
40%
20%
0%
100%
A=P = 4.6 m
80%
60%
40%
20%
0%
100%
A=P = 13.7 m
80%
60%
40%
20%
0%
Figure 5.28: Heat flux reduction relative to uninsulated foundations in mixed climate for a range
of foundation shapes
The results presented in Figures 5.29–5.31 represent a total of 4,500 simulations. Each plot
total heat transfer. The results demonstrate generally good agreement among each of the approxi-
mation methods with the exception of the circle approximation, which tended to over-predict three
dimensional heat transfer. Mean absolute deviation (MAD), and mean bias deviation (MBD) are
shown for each approximation method. Although the differences between the infinite rectangle,
rounded rectangle, and boundary layer adjustment methods are not clearly visible in the plots,
137
these metrics show that the former two methods have MAD values of 4–5%, while the boundary
The boundary layer adjustment method provides a more accurate representation of three-
dimensional heat transfer than previously established approximation methods. This is achieved
160 160
Circle Infinite Rectangle
140 140
Three-dimensional annual total heat transfer [MWh]
120 120
100 100
80 80
60 60
40 40
20 MAD = 8.9% 20 MAD = 4.8%
MBD = 8.8% MBD = 3.1%
0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
160 160
Rounded Rectangle Boundary Layer
140 140 Adjustment
120 120
100 100
80 80
60 60
40 40
20 MAD = 4.6% 20 MAD = 2.5%
MBD = 4.5% MBD = -0.8%
0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
Figure 5.29: Comparison of annual total heat transfer between three-dimensions and two-
dimensional approximations for mixed climate simulations
More detailed results are presented for mixed, cold, and hot climates in Appendices B, C,
and D, respectively. In these figures, the deviations of each approximation method are compared
for each three-dimensional simulation. Observations from these figures that are not apparent in
300 300
200 200
600 600
Rounded Rectangle Boundary Layer
500 500 Adjustment
400 400
300 300
200 200
Figure 5.30: Comparison of annual total heat transfer between three-dimensions and two-
dimensional approximations for cold climate simulations
• Methods other than the boundary layer adjustment method tend to over-predict concave
• The approximation methods perform better for shapes with larger area-to-perimeter ratios,
While annual total heat transfer is often the most important metric for energy design of
building foundations, variations in heat transfer throughout the year also impact peak heating and
cooling loads in a building used to calculated the appropriate capacity of HVAC equipment. Figure
5.32 demonstrates the time varying accuracy of the boundary layer adjustment method for a cross-
139
0 0
Circle Infinite Rectangle
−20 −20
−60 −60
−80 −80
−100 −100
−120 −120
0 0
Rounded Rectangle Boundary Layer
−20 −20
Adjustment
−40 −40
−60 −60
−80 −80
−100 −100
−120 −120
Figure 5.31: Comparison of annual total heat transfer between three-dimensions and two-
dimensional approximations for hot climate simulations
shaped basement with exterior, half-wall R-20 insulation. The three-dimensional results and the
Appendix E contains figures comparing annual minimum and maximum heat transfer rates
for all the tested combinations. The trends of agreement are similar to the comparisons of annual
total heat transfer. The boundary layer adjustment method provides the best approximation of
−1
l t r r r r
u ary ruaryarch Apri May June July gus mbetobe mbe mbe
n
Ja Feb M Au p te O c N o v e D e c e
Se
Time of year
Figure 5.32: Example comparison of hourly three-dimensional heat transfer and approximated
two-dimensional heat transfer
Table 5.4 presents the average simulation wall time for each of the two-dimensional approx-
imation methods. The methods that use only a Cartesian coordinate system require less time
Table 5.4: Average simulation wall time for two-dimensional approximation methods
Average
Method
Simulation Time [s]
Circle 40.9
Infinite rectangle 37.4
Rounded rectangle 40.2
Boundary layer adjustment 34.8
Within a two-dimensional context, the largest driver of simulation time is the foundation
design. Figure 5.33 demonstrates the variability of simulation time across foundation designs for
141
the boundary layer adjustment method. Simulations with more foundation elements require more
cells to define the near-field discretization region and therefore take longer to simulate.
A=P [m]
1.5 4.6 4.6
13.7 13.7 13.7
40
Simulation wall time [s]
30
20
10
Uninsulated
Uninsulated
The average simulation wall time of the boundary layer adjustment method is approximately
35 seconds. This is considerably faster than three-dimensional alternatives, but it still represents
a considerably large portion of the total simulation time for a whole-building energy simulation.
Further reductions in simulation time can be achieved by relaxing the refinement of the spatial
discretization. All simulations performed in approximation method tests used a minimum cell
dimension, Δxmin , of 6 millimeters, and a cell growth factor, fg , of 1.2. Figure 5.34 illustrates
the reduction in simulation wall time and corresponding impact on approximation accuracy for
142
the foundation design with the longest simulation time, a cross-shaped basement with exterior,
1.5%
1.0%
0.5%
0.0%
-0.5%
fg = 1.2 ¢xmin = 6 mm
-1.0%
fg = 1.5 ¢xmin = 37 mm
fg = 1.7 ¢xmin = 69 mm
-1.5%
fg = 2.0 ¢xmin = 100 mm
-2.0%
0 10 20 30 40 50
Simulation wall time [s]
Figure 5.34: Deviation from three-dimensional annual heat transfer vs. simulation wall time for
different combinations of discretization parameters
Increasing the cell growth factor results in larger positive deviations in approximation accu-
racy. Increasing the minimum cell dimension results in larger negative deviations in approximation
accuracy. These counteracting results cannot be relied upon to generate faster, more accurate sim-
ulations. However, it is possible to relax either or both of the discretization parameters to reduce
simulation wall time within 10 seconds without introducing deviations greater than 2%.
The accuracy and computational performance of the boundary layer adjustment method
demonstrate how the Kiva framework is applied to improve foundation heat transfer calculations.
Testing approximations against a verified, three-dimensional solution bolsters the algorithmic con-
fidence behind the calculations. Each simplification, whether to the spatial discretization, the
143
numerical scheme, the initialization method, or the number of dimensions, is justified through
comparative analysis.
Foundation heat transfer, which would take several days to simulate in three-dimensions on
a supercomputer, is approximated within five seconds on a laptop with less than 5% deviation
in results. The boundary layer adjustment method, with a suitably coarse spatial discretization,
would account for a suitably small fraction of total simulation time in a whole-building energy
simulation context.
Chapter 6
6.1 Conclusions
fully continuous range of insulation design options. The boundary layer adjustment method allows
for accurate simulations of almost any foundation shape. Automated discretization of domain
dramatically decreases the complexity of the problem, requiring the user to only input information
about the foundation system, and not details of numerical methods used to solve the problem.
Three different components of the foundation heat transfer calculation were tested within the
Kiva framework to assess their impact on both computational performance and accuracy. These
components include the numerical scheme, the initialization method, and the two-dimensional
approximation of three-dimensional heat transfer. Of the numerical schemes tested, the Alternating
Direction Implicit (ADI) scheme demonstrates the best balance between accuracy, performance, and
numerical stability. The new accelerated initialization approach significantly reduces the required
years of presimulation. The boundary layer adjustment method improves accuracy over other
This method accounts for reduced heat transfer from concave foundation shapes, which has not
been adequately addressed to date. Within the Kiva framework, three-dimensional heat transfer
145
that can require several days to simulate is approximated in two-dimensions in a matter of seconds
of foundation heat transfer providing algorithmic confidence without the computational burden
The combined recommendations presented in Table 6.1 result in reliably efficient and accurate
foundation heat transfer calculations that can be performed in the context of whole-building energy
analysis.
Table 6.1: Recommended methods and numerical parameters for foundation heat transfer
Description Value
Minimum cell dimension 0.02 m
Cell growth coefficient 1.50
Time-step duration 60 minutes
Numerical method ADI
Initialization method Accelerated
Initialization acceleration time-step duration 1 week
Initialization acceleration period duration 12 weeks
Initialization warm-up duration 1 year
Two-dimensional approximation Boundary layer adjustment method
The flexible framework and open source development positions Kiva for extension into future
studies and further improvements to calculating heat transfer from building foundations.
Kiva must be integrated into a whole-building energy simulation engine before its advantages
are fully utilized by energy modeling professionals. The Kiva framework was designed with integra-
tion in mind, by exposing access to important run-time variables that determine the heat balance
between the foundation system and the rest of the building. Once integrated, Kiva will provide
146
convective heat flux from foundation surfaces to thermal zones in a whole-building energy model,
which, in turn, provides the thermal zone temperatures that determine heat transfer through the
In addition to interior zone temperatures, the whole-building simulation engine will provide
information from HVAC systems that interact with the foundation and the surrounding ground.
Kiva will use heat source values to simulate radiant slabs, ground source heat pumps, and snow-melt
systems.
Once Kiva is integrated into a whole-building simulation engine, it can be used to help in-
form changes to foundation design guidelines. Current guidelines for foundation insulation design
include the Builder’s Foundation Handbook [20], which includes insulation recommendations based
on analysis performed in 1988 [77] and ASHRAE Standard 90.1 [8], which defines levels of insu-
lation based on analysis performed in 1991 [12]. The underlying methodology behind both sets
of guidelines are woefully out of date. Both the recommendations in the Builder’s Foundation
Handbook and the standard insulation levels in ASHRAE Standard 90.1 can be updated based on
There are several capabilities mentioned in the literature review chapter (Chapter 2) that
were not incorporated into Kiva by the time of this thesis. Some suggested capabilities are listed
• Coupled heat and moisture transfer calculations with dynamic thermal properties
Two types of sensitivity analyses can be performed using the Kiva framework. The first
type of analysis tests computational performance and accuracy impacts of certain computational
routines. There are several routines that are not included in Kiva because of their perceived impact
on computational performance. The goal of these analyses is to quantify the trade-off between
computational performance and accuracy of their respective routines. Potential routines include:
• evapotranspiration
The other type of analysis involves exploration of propagation of uncertainty in model inputs.
Because there are many relevant inputs to a foundation heat transfer model that are difficult
uncertainty as it propagates into results and, ultimately, recommended foundation insulation levels.
Two important inputs to investigate are the thermal properties of the soil, and the deep ground
boundary condition, which is strongly influenced by the presence of a high water table.
The source code of Kiva has not been optimized for memory utilization or CPU performance.
There are likely several areas in the source code that can be improved to increase Kiva’s speed
and accessibility. There are several tools available to profile the execution of Kiva’s source code to
Iterative solvers have also improved drastically over the past several decades. Many authors
cite Algebraic Multigrid (AMG) methods with computational efficiencies comparable to the Al-
ternating Direction Implicit (ADI) method. AMG methods do not require a structured grid, and
can be solved using parallel processes [19]. With the increasing trend of parallel architecture in
computing, AMG methods should be evaluated as a potential replacement for the ADI method.
Bibliography
[1] M.H. Adjali, M. Davies, and J. Littler. Three-Dimensional Earth-Contact Heat Flows: A
Comparison of Simulated and Measured Data for a Buried Structure. Renewable Energy, 15:
356–359, April 1998.
[2] M.H. Adjali, M. Davies, and J. Littler. Earth-Contact Heat Flows: Review and Application
of Design Guidance Predictions. Building Services Engineering Research and Technology, 19
(3):111–121, 1998.
[3] M.H. Adjali, M. Davies, and J. Littler. A Numerical Simulation of Measured Transient
Temperatures in the Walls, Floor and Surrounding Soil of a Buried Structure. International
Journal of Heat and Mass Transfer, 9(4):405–422, 1999.
[4] M.H. Adjali, M. Davies, S.W. Rees, and J. Littler. Temperatures in and under a Slab-on-
Ground Floor: Two- and Three-Dimensional Numerical Simulations and Comparison with
Experimental Data. Building and Environment, 35:655–662, October 2000.
[5] M.H. Adjali, M. Davies, and S.W. Rees. A Comparative Study of Design Guide Calculations
and Measured Heat Loss through the Ground. Building and Environment, 39(11):1301–1311,
November 2004.
[6] A. Al-Anzi and M. Krarti. Local/global analysis of transient heat transfer from building
foundations. Building and Environment, 39(5):495–504, May 2004.
[8] ASHRAE. ANSI/ASHRAE/IES Standard 90.1-2013: Energy Standard for Buildings Except
Low-Rise Residential Buildings, 2013.
[9] W. Bahnfleth. Three-Dimensional Modeling of Heat Transfer from Slab Floors. Ph.d. thesis,
University of Illinois at Urbana Champaign, 1989.
[10] W. Bahnfleth and J. Amber. Algorithms for Slab-on-Grade Heat Transfer Calculations. Tech-
nical report, US Army Corps of Engineers Construction Engineering Research Laboratory,
Champaign, IL, 1990.
[12] D. Baylon and M. Kennedy. Calculating the Impact of Ground Contact on Residential Heat
Loss. In Proceedings of Thermal Performance of the Exterior Envelopes of Buildings X,
Clearwater, Florida, 2007. ASHRAE.
[14] I. Beausoleil-Morrison and M. Krarti. Predicting Foundation Heat Losses: Neural Networks
Versus the BASESIMP Correlations. In Proceedings of Building Simulation 1997, Prague,
Czech Republic, 1997. IBPSA.
[18] C. Booten, N. Kruis, and C. Christensen. Identifying and Resolving Issues in EnergyPlus
and DOE-2 Window Heat Transfer Calculations. Technical Report August, NREL, Golden,
Colorado, 2012.
[19] W. Briggs, V. Hensen, and S. McCormick. A Multigrid Tutorial. Society of Industrial and
Applied Mathematics, Philadelphia, Pennsylvania, 2 edition, 2000.
[20] J. Carmody, J. Christian, and K. Labs. Builder’s Foundation Handbook. Oak Ridge National
Laboratory, Oak Ridge, Tennessee, 1991.
[21] M. Chang, L. Chow, and W. Chang. Improved Alternating-Direction Implicit Method for
Solving Transient Three-Dimensional Heat Diffusion Problems. Numerical Heat Transfer:
Part B: Fundamentals, 19:69–84, 1991.
[22] S. Choi and M. Krarti. Heat transfer for slab-on-grade floor with stepped ground. Energy
Conversion and Management, 39(7):691–701, 1998.
[23] S. Choi and M. Krarti. Thermally optimal insulation distribution for underground structures.
Energy and Buildings, 32(3):251–265, September 2000.
[24] P. Chuangchid. Heat Gain from Slab-on-Grade Floor of Refrigerated Structures. Ph.d. thesis,
University of Colorado at Boulder, 2000.
[26] J. Claesson and C.E. Hagentoft. Heat loss to the ground from a building–I. General theory.
Building and Environment, 26(2):195–208, 1991.
151
[27] D. Claridge. Design Methods for Earth-Contact Heat Transfer. In Karl Böer, editor, Advances
in Solar Energy: An Annual Review of Research and Development, chapter 7, pages 305–351.
American Solar Energy Society, Inc., Boulder, CO, 4 edition, 1988.
[28] J. Clark and M. Korybalski. Algebraic Methods for the Calculation of Radiation Exchange
in an Enclosure. Wärme - und Stoffübertragung, 7(1):31–44, 1974.
[29] E. Clements. Three Dimensional Foundation Heat Transfer Modules for Whole-Building
Energy Analysis. Master’s thesis, Pennsylvania State University, 2004.
[30] C. Cogil. Modeling of Basement Heat Transfer and Parametric Study of Basement Insulation
for Low Energy Housing. Master’s thesis, Pennsylvania State University, 1998.
[31] M. Crowley. Modeler Report for BESTEST Cases GC10a-GC80c, MATLAB 7.0.4.365 (R14)
Service Pack 2. Technical report, Dublin Institute of Technology, Dublin, Ireland, 2007.
[32] J. Cullin, L. Xing, E. Lee, J. Spitler, and D. Fisher. Feasibility of Foundation Heat Ex-
changers for Residential Ground Source Heat Pump Systems in the United States. ASHRAE
Transactions, 118(1):1039–1048, 2012.
[34] M. Davies, S. Zoras, and M.H. Adjali. A Potentially Fast, Flexible and Accurate Earth-
Contact Heat Transfer Simulation Method. In Proceedings of Building Simulation 1999,
number 1987, Kyoto, Japan, 1999. IBPSA.
[35] D. De Vries. Simultaneous Transfer of Heat and Moisture in Porous Media. American
Geophysical Union Transactions, 39(5):909–916, 1958.
[36] D. De Vries. Thermal Properties of Soils. In W Van Wijk, editor, Physics of Plant
Environment, chapter 7, pages 210–235. North-Holland Publishing Co., Amsterdam, 1966.
[37] A. Delsante. The effect of water table depth on steady-state heat transfer through a slab-on-
ground floor. Building and Environment, 28(3):369–372, 1993.
[38] A. Delsante, A. Stokes, and P. Walsh. Application of Fourier Transforms to Periodic Heat
Flow into the Ground Under a Building. International Journal of Heat and Mass Transfer,
26(1):121–132, 1983.
[39] M. Deru. A Model for Ground-Coupled Heat and Moisture Transfer from Buildings A Model
for Ground-Coupled Heat and Moisture Transfer from Buildings. Technical Report June,
National Renewable Energy Laboratory, Golden, Colorado, 2003.
[40] M. Deru, R. Judkoff, and J. Neymark. Whole Building Energy Simulation with a Three-
Dimensional Ground-Coupled Heat Transfer Model. ASHRAE Transactions, 109(1):557–565,
2003.
[41] G.H. dos Santos and N. Mendes. Multidimensional Effects of Ground Heat Transfer on the
Dynamics of Building Thermal Performance. ASHRAE Transactions, 110(2):345–354, 2004.
[42] G.H. dos Santos and N. Mendes. Simultaneous heat and moisture transfer in soils combined
with building simulation. Energy and Buildings, 38(4):303–314, April 2006.
152
[43] J. Duffie and W. Beckman. Solar Engineering of Thermal Processes. John Wiley & Sons Ltd,
Hoboken, New Jersey, third edit edition, 2006.
[44] D. Fisher and S. Rees. Modeling Ground Source Heat Pump Systems in a Building Energy
Simulation Program (EnergyPlus). In Proceedings of Building Simulation 2005, pages 311–
318, Montreal, Canada, 2005. IBPSA.
[45] P. Fritzson. Principles of Object-Oriented Modeling and Simulation with Modelica 2.1.
Wiley-IEEE Press, 2003. ISBN 0471471631.
[46] K. Fukuyo. Conditional Stability of Larkin Methods with Non-Uniform Grids. Theoretical
and Applied Mechanics, 37(2):139–159, 2010. ISSN 1450-5584. doi: 10.2298/TAM1002139F.
[47] R. Gilpin and B. Wong. ”Heat-Valve” Effects in the Ground Thermal Regime. Journal of
Heat Transfer, 98(4):537–542, 1976.
[48] C.E. Hagentoft. Heat Loss to the Ground from a Building: Slab on the Ground and Cellar.
Ph.d. thesis, Lund Institute of Technology, Sweden, 1988.
[49] C.E. Hagentoft. Heat Losses and Temperature in the Ground under a Building with and with-
out Ground Water Flow—I. Infinite Ground Water Flow Rate. Building and Environment,
31(1):3–11, January 1996.
[50] C.E. Hagentoft. Steady-State Heat Loss for an Edge-Insulated Slab : Part I. Building and
Environment, 37(1):19–25, January 2002.
[51] C.E. Hagentoft and J. Claesson. Heat Loss to the Ground from a BuildingII. Slab on the
Ground. Building and Environment, 26(4):395–403, 1991.
[52] C.E. Hagentoft and J. Claesson. Influence of Rain Water Percolation on Ground Heat
Losses and Temperature for Basement Foundation. In Proceedings of the 3rd International
Conference on Research in Building Physics and Building Engineering, Montreal, Canada,
2006.
[53] H. Hottel and A. Sarofim. Radiative Transfer. McGraw Hill Book Co., 1967.
[54] Y. Huang, L. Shen, J. Bull, and L. Goldberg. Whole-house simulation of foundation heat
flows using the DOE-2.1 C program. ASHRAE Transactions, 94(2):936–958, 1988.
[56] H. Janssen. The Influence of Soil Moisture Transfer On Building Heat Loss Via the Ground.
Ph.d. thesis, Katholieke Universiteit Leuven, July 2002.
[57] H. Janssen, J. Carmeliet, and H. Hens. The Influence of Soil Moisture Transfer on Building
Heat Loss via the Ground. Building and Environment, 39(7):825–836, July 2004.
[58] N. Jones and D. Greenberg. Fast Computation of Incident Solar Radiation from Preliminary
to Final Building Design. In Proceedings of Building Simulation 2011, pages 14–16, Sydney,
Australia, 2011. IBPSA.
153
[59] R. Judkoff and J. Neymark. Home Energy Rating System Building Energy Simulation Test
( HERS BESTEST ) Volume 1 Tier 1 and Tier 2 Tests User s Manual. Technical Report
November, National Renewable Energy Laboratory, Golden, Colorado, 1995.
[60] R. Judkoff and J. Neymark. Model Validation and Testing : The Methodological Foundation
of ASHRAE Standard 140. ASHRAE Transactions, 112(2):367–376, 2006.
[61] B.D. Kay and E. Perfect. State of the art: Heat and mass transfer in freezing soils. In
Proceedings of 5th International Symposium on Ground Freezing, Rotterdam, Netherlands,
1988. Balkema, Rotterdam.
[63] M. Krarti. Time-Varying Heat Transfer From Slab-on-Grade Floors with Vertical Insulation.
Building and Environment, 29(1):55–61, January 1994.
[64] M. Krarti. Comparison of a Neural Network Model with a Regression Model for Foundation
Heat Loss Calculation. In Proceedings of Thermal Performance of the Exterior Envelopes of
Buildings VI, pages 79–85, Clearwater, Florida, 1995. ASHRAE.
[65] M. Krarti. Effect of Spatial Variation of Soil Thermal Properties on Slab-on-Ground Heat
Transfer. Building and Environment, 31(1):51–57, January 1996.
[66] M. Krarti. Building Foundation Heat Transfer. In Yogi Goswami and Karl Böer, editors,
Advances in Solar Energy: An Annual Review of Research and Development, chapter 6, pages
241–315. American Solar Energy Society, Inc., Boulder, CO, 13 edition, 1999.
[68] M. Krarti and S. Choi. Simplified Method for Foundation Heat Loss Calculation. ASHRAE
Transactions, 102(1):140–152, 1996.
[69] M. Krarti and S. Choi. A Simulation Method for Fluctuating Temperatures in Crawlspace
Foundations. Energy and Buildings, 26(2):183–188, 1997.
[70] M. Krarti, D. Claridge, and J. Kreider. ITPE Technique Applications to Time-Varying Two-
Dimesional Ground-Coupling Problems. International Journal of Heat and Mass Transfer, 31
(9):1899–1911, 1988.
[71] M. Krarti, D. Claridge, and J. Kreider. The ITPE Technique Applied to Steady-State Ground-
Coupling Problems. International Journal of Heat and Mass Transfer, 31(9):1885–1898, 1988.
[72] M. Krarti, J. Kreider, and D. Claridge. ITPE Technique Applications to Time-Varying Three-
Dimensional Ground-Coupling Problems. Journal of Heat Transfer, 112:849–856, 1990.
[73] M. Krarti, P. Chuangchid, and P. Ihm. Foundation Heat Transfer Module for EnergyPlus
Program. In Proceedings of Building Simulation 2001, pages 931–938, Rio de Janeiro, Brazil,
2001. IBPSA.
[75] T. Kusuda and P.R. Achenbach. Earth temperature and thermal diffusivity at selected sta-
tions in the United States. ASHRAE Transactions, 73:61–75, 1965.
[76] T. Kusuda and J.W. Bean. Simplified Methods for Determining Seasonal Heat Loss from
Uninsulated Slab-on-Grade Floors. ASHRAE Transactions, 90(1):611–632, 1984.
[79] J. Latta and G. Boileau. Heat Losses from House Basements. Canadian Building, XIX(10):
39–42, 1969.
[80] Lawrence Berkeley National Laboratory and James J. Hirsch & Associates.
DOE-2.2: Building Energy Use and Cost Analysis Program. Volume 2: Dictionary. James
J. Hirsch & Associates, Camarillo, California, 2009. URL http://www.doe2.com/.
[81] N. Li, F. Chen, B. Xu, and G. Swoboda. Theoretical Modeling Framework for an Unsaturated
Freezing Soil. Cold Regions Science and Technology, 54:19–35, 2008.
[82] X. Liu, S. Rees, and J. Spitler. Modeling Snow Melting on Heated Pavement Surfaces. Part
I: Model Development. Applied Thermal Engineering, 27(5-6):1115–1124, April 2007.
[83] C. Ludeman. Passive Project Foundation and Slab Insulation, 2009. URL http://www.
100khouse.com/2009/06/11/passive-project-foundation-slab-insulation/.
[84] X. Ma, B. Cheng, J. Mao, W. Liu, and D. Zi. Finite Element Modelling of Coupled Heat and
Moisture Transfer in Typical Earth-Sheltered Building Envelope. In Proceedings of Building
Simulation 2009, pages 1850–1856, Glasgow, Scotland, 2009. IBPSA.
[85] J. MacArthur, G. Meixel, and L. Shen. Application of Numerical Methods for Predicting
Energy Transport in Earth Contact Systems. Applied Energy, 13:121–156, 1983.
[87] G. Mitalas. Calculation of Below-Grade Residential Heat Loss: Low-Rise Residential Build-
ing. ASHRAE Transactions, 93(1):743–783, 1987.
[88] F. Moukalled and Y. Saleh. Heat and Mass Transfer in Moist Soil, Part I. Formulation and
Testing. Numerical Heat Transfer: Part B: Fundamentals, 49(5):467–486, October 2006.
[89] A. Nakhi. Modeler Report for BESTEST Cases GC10a-GC80c, FLUENT Version 6.0.20.
Technical report, Public Authority for Applied Education and Training, Kuwait, 2007.
155
[90] J. Neymark and R. Judkoff. International Energy Agency Building Energy Simulation Test
and Diagnostic Method ( IEA BESTEST ): In-Depth Diagnostic Cases for Ground Coupled
Heat Transfer Related to Slab-on-Grade Construction. Technical Report September, National
Renewable Energy Laboratory, Golden, Colorado, 2008.
[91] J. Rantala and V. Leivo. Heat Loss into Ground from a Slab-on-Ground Structure in a Floor
Heating System. International Journal of Energy Research, 30(12):929–938, October 2006.
[92] S. Rees, J. Spitler, and X. Xiao. Transient Analysis of Snow-Melting System Performance.
ASHRAE Transactions, 108(2):406–423, 2002.
[93] S.W. Rees and H. Thomas. Two-Dimensional Heat Transfer beneath a Modern Commercial
Building: Comparison of Numerical Prediction with Field Measurement. Building Services
Engineering Research and Technology, 18(3):169–174, 1997.
[94] S.W. Rees, R. Lloyd, and H. Thomas. A Numerical Simulation of Measured Transient Heat
Transfer through a Concrete Ground Floor Slab and Underlying Substrata. International
Journal of Numerical Methods for Heat & Fluid Flow, 5(8):669–683, 1995.
[95] S.W. Rees, H. Thomas, and Z. Zhou. Ground Heat Transfer: Some Further Insights into
the Influence of Three-Dimensional Effects. Building Services Engineering Research and
Technology, 21(4):233–239, 2000.
[96] S.W. Rees, Z. Zhou, and H. Thomas. The Influence of Soil Moisture Content Variations
on Heat Losses from Earth-Contact Structures : An Initial Assessment. Building and
Environment, 36:157–165, 2001.
[97] S.W. Rees, H. Thomas, and Z. Zhou. A Numerical and Experimental Investigation of Three-
Dimensional Ground Heat Transfer. Building Services Engineering Research and Technology,
27(3):195–208, 2006.
[98] S.W. Rees, Z. Zhou, and H. Thomas. Ground Heat Transfer: A Numerical Simulation of a
Full-Scale Experiment. Building and Environment, 42(3):1478–1488, March 2007.
[99] B. Rock. Sensitivity Study of Slab-on-Grade Transient Heat Transfer Model Parameters.
ASHRAE Transactions, 110(1):177–184, 2004.
[100] V. Saul’yev. Integration of Equations of Parabolic Type by the Method of Nets. Pergamon
Press, New York, 1964.
[101] L. Shen. An Investigation of Transient, Two-Dimensional Coupled Heat and Moisture Flow
in Soils. Ph.d. thesis, University of Minnesota, 1986.
[102] L. Shen and J. Ramsey. An Investigation of Transient, Two-Dimensional Coupled Heat and
Moisture Flow in the Soil Surrounding a Basement Wall. International Journal of Heat and
Mass Transfer, 31(7):1517–1527, 1988.
[104] Solar Energy Laboratory. TRNSYS 17: a TRaNsient SYstem Simulation program. Solar En-
ergy Laboratory, University of Wisconsin-Madison, Madison, Wisconsin, 2010.
[105] R. Sterling, S. Gupta, L. Shen, and L. Goldberg. Assessment of Soil Thermal Conductivity
for Use in Building Design and Analysis. Technical report, ASHRAE, Atlanta, GA, 1993.
[106] The Scalable Software Infrastructure Project. Lis User Guide. The Scalable Software Infras-
tructure Project, Fukuoka, Japan, 2014.
[107] H. Thomas and S.W. Rees. The Thermal Performance of Ground Floor Slabs–a Full Scale
in-situ experiment. Building and Environment, 34:139–164, March 1999.
[108] H. Thomas and S.W. Rees. Measured and Simulated Heat Transfer to Foundation Soils.
Géotechnique, 59(4):365–375, 2009.
[109] J. Thornton. Modeler Report for BESTEST Cases GC10a-GC80c, TRNSYS Version 16.1.
Technical report, Thermal Energy Systems Specialists, Madison, Wisconsin, 2007.
[110] H. Trethowen and A. Delsante. A Four-Year Site Measurement of Heat Flow in Slab-on-
Ground Floors with Wet Soils. In Proceedings of Thermal Performance of the Exterior
Envelopes of Buildings VII, Clearwater, Florida, 1998. ASHRAE.
[111] United States Department Of Energy (DOE). EnergyPlus: Input Output Reference. The
Board of Trustees of the University of Illinois and the Regents of the University of California
through the Ernest Orlando Lawrence Berkeley National Laboratory, Berkeley, California,
version 8. edition, 2014.
[112] B. Usowicz, J. Lipiec, and A. Ferrero. Prediction of soil thermal conductivity based on
penetration resistance and water content or air-filled porosity. International Journal of Heat
and Mass Transfer, 49(25-26):5010–5017, December 2006.
[113] G. Walton. Thermal Analysis Research Program Reference Manual. U.S. Dept. of Com-
merce, National Bureau of Standards, Washington D.C., 1983.
[114] G. Walton. Estimating 3-D heat loss from rectangular basements and slabs using 2-D calcu-
lations. ASHRAE Transactions, 93(1):791–797, 1987.
[115] F. Wang. Mathematical Modeling and Computer Simulation of Insulation Systems in Below
Grade Applications. In Proceedings of Thermal Performance of the Exterior Envelopes of
Buildings I, pages 456–471, Clearwater, Florida, 1979. ASHRAE.
[116] E. Wentzel. Thermal Modeling of Walls, Foundations and Whole Buildings Using Dynamic
Thermal Networks. Ph.d. thesis, Chalmers University of Technology, 2005.
[117] S. Wilcox and W. Marion. Users Manual for TMY3 Data Sets. Technical report, National
Renewable Energy Laboratory, Golden, Colorado, 2008.
[118] F. Winkelmann. Underground Surfaces: How To Get A Better Underground Surface Heat
Transfer Calculation In DOE-2.1E. DOE-2 Articles from the Building Energy Simulation
User News, 23(5):5–14, 2002.
157
[119] X. Xie, Y. Jiang, and J. Xia. A New Approach to Compute Heat Transfer of Ground-Coupled
Envelope in Building Thermal Simulation Software. Energy and Buildings, 40(4):476–485,
2008.
[120] M. Yazdanian and J. Klems. Measurement of the exterior convective film coefficient for
windows in low-rise buildings. ASHRAE Transactions, 100(1), 1994. URL http://btech.
lbl.gov/papers/34717.pdf.
[121] C. Zhang and G. Ding. A novel thermal response factor method for the dynamic load cal-
culation of buildings. Journal of Asian Architecture and Building Engineering, 1(1):75–79,
2002.
Appendix A
The discretization equations for each of the finite difference methods implemented in Kiva
Heat transfer in the ground domain is solved using the diffusion equation
∂T
CH,T = ∇ [k∇T ] + q̇ (A.1)
∂t
where:
• t is time [s]
The spatial gradient operator creates different equations depending on the coordinate system.
There are four types of boundary conditions encountered in Kiva. The formulation of the
boundary conditions are framed generically for a direction w, which is perpendicular to the surface.
Interior flux boundaries in Kiva fall along the foundation floor, the foundation wall, and the
where
Exterior flux boundaries are very similar to the interior boundaries except that they include
a separate long-wave radiation term which accounts for the difference between the radiative tem-
perature, T∞ F ∗ 1/4 , and the outdoor ambient temperature, T∞ . This boundary is applied to the
where F ∗ is a correction factor to account for the difference between the sky temperature and the
Zero-flux boundaries occur at far-field, deep-ground and symmetry planes as well as the top
Constant temperature boundaries are found either at deep-ground or at the wall-top (to
T (w) = Tc (A.8)
The following sections describe the discretized equations for each of the finite difference
methods implemented in Kiva. The equations presented here correspond to the continuous partial
differential equation for a two-dimensional cylindrical coordinate system (Equation A.75). For-
mulations for Cartesian two- and three-dimensional coordinate systems can be easily extended by
The only difference between the different finite difference schemes is the time-step associated
with the temperature terms. A generic finite difference scheme with temperatures T ∗ will provide
n+1
Ti,j − Ti,j
n ∗
Ti+1,j ∗
− Ti,j ∗ − T∗
Ti,j
1 Δrm Δrp i−1,j
CH,T = kr,p + kr,m
Δt ri Δrm + Δrp Δrp Δrm + Δrp Δrm
∗
Ti+1,j ∗
− Ti,j ∗ − T∗
Ti,j
2 i−1,j (A.9)
+ kr,p − kr,m
Δrm + Δrp Δrp Δrm
∗
Ti,j+1 ∗
− Ti,j ∗ − T∗
Ti,j
2 i,j−1
+ kz,p − kz,m + q̇.
Δzm + Δzp Δzp Δzm
where
Δri + Δri+1
Δrp = , (A.10)
2
Δri + Δri−1
Δrm = , (A.11)
2
Δzi + Δzi+1
Δzp = , (A.12)
2
Δzi + Δzi−1
Δzm = , (A.13)
2
−1
Δri Δri+1
kr,p = + , (A.14)
2Δrp ki,j 2Δrp ki+1,j
−1
Δri Δri−1
kr,m = + , (A.15)
2Δrm ki,j 2Δrm ki−1,j
−1
Δzj Δzj+1
kz,p = + , (A.16)
2Δzp ki,j 2Δzp ki,j+1
−1
Δzj Δzj−1
kz,m = + , (A.17)
2Δzm ki,j 2Δzm ki,j−1
where
The nontemperature terms can then be collected into constant coefficients to simplify the
Δrm kr,p
cr,p = θ , (A.19)
ri [Δrm + Δrp ] Δrp
162
Δrp kr,m
cr,m = θ , (A.20)
ri [Δrm + Δrp ] Δrm
2kr,p
cr,p = θ , (A.21)
[Δrm + Δrp ] Δrp
2kr,m
cr,m = −θ , (A.22)
[Δrm + Δrp ] Δrm
2kz,p
cz,p = θ , (A.23)
[Δzm + Δzp ] Δzp
2kz,m
cz,m = −θ . (A.24)
[Δzm + Δzp ] Δzm
For Cartesian coordinates, the radial terms (cr,p , cr,m , cr,p , and cr,m ) will be omitted (set to
zero), and the x and y terms (cx,p , cx,m , cy,p , and cy,m ) can be defined in a similar fashion as the z
terms.
With the definitions of the constant coefficients, Equation A.9 can be re-written as:
∗ ∗ ∗
n+1
Ti,j − Ti,j
n
= cr,p Ti+1,j ∗
− Ti,j + cr,m Ti,j ∗
− Ti−1,j + cr,p Ti+1,j ∗
− Ti,j
∗ ∗
∗ ∗
∗ ∗
(A.25)
+ cr,m Ti,j − Ti−1,j + cz,p Ti,j+1 − Ti,j + cz,m Ti,j − Ti,j−1 + θq̇.
The spatial derivative in the boundary condition relationships can be descretized as follows:
∂T T ∗ − Tl∗
≈ l−1 (A.26)
∂w Δwm
where l is the spatial index in the direction normal to the boundary, and
Δwm = Δwl−1 /2. (A.27)
∗ − T∗
Tl−1
−kw,m l
= hc [T∞ − Tl∗ ] + hr T∞ F ∗ 1/4 − Tl∗ + q̇sw , (A.29)
Δwm
∗ − T∗
Tl−1 l
= 0, (A.30)
Δwm
Tl∗ = Tc . (A.31)
163
In the explicit scheme, all the temperatures on the right-hand side of Equation A.9 represent
n n n
n+1
Ti,j − Ti,j
n
= cr,p Ti+1,j − Ti,j
n
+ cr,m Ti,j − Ti−1,j
n
+ cr,p Ti+1,j − Ti,j
n
n n n (A.32)
+ cr,m Ti,j − Ti−1,j
n
+ cz,p Ti,j+1 − Ti,j
n
+ cz,m Ti,j − Ti,j−1
n
+ θq̇.
This equation can be solved explicitly for the temperature at the new time-step, n + 1:
n+1
Ti,j n
= Ti,j 1 + cr,m + cr,m + cz,m − cr,p − cr,p − cz,p + Ti−1,j
n
−cr,m − cr,m
n
n n
(A.33)
+ Ti+1,j cr,p + cr,p + Ti,j−1 [−cz,m ] + Ti,j+1 [cz,p ] + θq̇.
In the fully implicit scheme, all the temperatures on the right-hand side of Equation A.9
n+1
Ti,j − Ti,j
n
= cr,p Ti+1,j
n+1 n+1
− Ti,j + cr,m Ti,j
n+1 n+1
− Ti−1,j n+1
+ cr,p Ti+1,j n+1
− Ti,j
(A.41)
n+1 n+1 n+1
+ cr,m Ti,j − Ti−1,j + cz,p Ti,j+1 − Ti,j n+1 n+1 n+1
+ cz,m Ti,j − Ti,j−1 + θq̇.
This equation can be rearranged with all the unknown temperatures at the new time-step,
n + 1, on the left-hand side of the equation with constant coefficients and all the constant terms
n+1
Ti,j 1 + cr,p + cr,p + cz,p − cr,m − cr,m − cz,m + Ti−1,jn+1
cr,m + cr,m
n+1
(A.42)
+ Ti+1,j n+1
−cr,p − cr,p + Ti,j−1 n+1
[cz,m ] + Ti,j−1 n
[−cz,p ] = Ti,j + θq̇
which can be arranged with the unknown temperatures on the left-hand side as
kw,m n+1 kw,m
Tln+1 + [hc + hr ] − Tl−1 = [hc + hr ] T∞ − q̇sw . (A.44)
Δwm Δwm
which can be arranged with the unknown temperatures on the left-hand side as
kw,m
n+1 kw,m
Tln+1 + [hc + hr ] − Tl−1 = hc + hr F ∗ 1/4 T∞ − q̇sw . (A.46)
Δwm Δwm
which can be arranged with the unknown temperatures on the left-hand side as
Tln+1 − Tl−1
n+1
= 0. (A.48)
Tln+1 = Tc . (A.49)
In the Crank-Nicolson scheme, the right-hand side of the equation is the average of the explicit
and fully implicit schemes: In the fully implicit scheme, all the temperatures on the right-hand side
This equation can be rearranged with all the unknown temperatures at the new time-step,
n + 1, on the left-hand side of the equation with constant coefficients and all the constant terms
1 n+1 1
n+1
Ti,j 1+ cr,p + cr,p + cz,p − cr,m − cr,m − cz,m + Ti−1,j cr,m + cr,m
2 2
n+1 1
n+1 1 n+1 1
+ Ti+1,j −cr,p − cr,p + Ti,j−1 [cz,m ] + Ti,j−1 [−cz,p ]
2 2 2 (A.51)
1 1
n
= Ti,j 1+ cr,m + cr,m + cz,m − cr,p − cr,p − cz,p + Ti−1,j n
−cr,m − cr,m
2 2
1 1 1
n
+ Ti+1,j c + cr,p + Ti,j−1
n n
[−cz,m ] + Ti,j+1 [cz,p ] + θq̇.
2 r,p 2 2
The boundary conditions for the Crank-Nicolson scheme are identical to those of the fully
implicit scheme.
166
In the ADE scheme, two discretized equations are solved independently. In the upward sweep,
all the lower indexed temperature differences on the right-hand side of Equation A.9 represent the
new time-step, n + 1 and all the higher indexed temperature differences represent the previous
n n
n+1
Ui,j − Ui,j
n
= cr,p Ui+1,j − Ui,j
n
+ cr,m Ui,j
n+1 n+1
− Ui−1,j + cr,p Ui+1,j − Ui,j
n
n (A.52)
n+1 n+1
+ cr,m Ui,j − Ui−1,j + cz,p Ui,j+1 − Ui,j + cz,m Ui,j − Ui,j−1 + θq̇.
n n+1 n+1
This equation can be solved explicitly for the temperature at the new time-step, n + 1:
n+1
Ui,j n
= Ui,j 1 − cr,p − cr,p − cz,p − Ui−1,j
n+1 n
cr,m + cr,m + Ui+1,j n+1
cr,p + cr,p − Ui,j−1 [cz,m ]
−1 (A.53)
+ Ui,j+1 [cz,p ] + θq̇ 1 − cr,m − cr,m − cz,m
n
In the downward sweep, all the higher indexed temperature differences on the right-hand
side of Equation A.9 represent the new time-step, n + 1 and all the lower indexed temperature
differences represent the previous time-step, n. The temperatures calculated in the downward
n
n+1
Vi,j − Vi,j
n
= cr,p Vi+1,j
n+1 n+1
− Vi,j + cr,m Vi,j − Vi−1,j
n n+1
+ cr,p Vi+1,j n+1
− Vi,j
n n (A.54)
n n+1
+ cr,m Vi,j − Vi−1,j + cz,p Vi,j+1 − Vi,j n+1
+ cz,m Vi,j − Vi,j−1 + θq̇.
n
This equation can be solved explicitly for the temperature at the new time-step, n + 1:
n+1
Vi,j n
= Vi,j 1 + cr,m + cr,m + F − Vi−1,j
n n+1
cr,m + cr,m + Vi+1,j cr,p + cr,p − Vi,j−1
n
[F ]
(A.55)
n+1 −1
+ Vi,j+1 [cz,p ] + θq̇ 1 + cr,p + cr,p + cz,p
The ADE boundary conditions follow the same pattern of higher and lower indices. The
equations below provide only the cases where indices are increasing in a direction normal to the
For interior flux boundaries, the discretized equation for the upward sweep becomes
n+1
Ul−1 − Uln+1
−kw,m = [hc + hr ] T∞ − Uln+1 + q̇sw , (A.57)
Δwm
kw,m n+1
U + [hc + hr ] T∞ − q̇sw
Δwm l−1
Uln+1 = . (A.58)
kw,m
+ [hc + hr ]
Δwm
The discretized equation for the downward sweep becomes
n − V n+1
Vl−1 l
−kw,m = [hc + hr ] T∞ − Vln+1 + q̇sw , (A.59)
Δwm
kw,m n
V + [hc + hr ] T∞ − q̇sw
Δwm l−1
Vln+1 = . (A.60)
kw,m
+ [hc + hr ]
Δwm
For exterior flux boundaries, the discretized equation for the upward sweep becomes
n+1
Ul−1 − Uln+1
−kw,m = hc T∞ − Uln+1 + hr T∞ F ∗ 1/4 − Uln+1 + q̇sw , (A.61)
Δwm
kw,m n+1
Ul−1 + hc + hr F ∗ 1/4 T∞ − q̇sw
Δwm
Uln+1 = . (A.62)
kw,m
+ [hc + hr ]
Δwm
The discretized equation for the downward sweep becomes
n − V n+1
Vl−1
−kw,m l
= hc T∞ − Vln+1 + hr T∞ F ∗ 1/4 − Vln+1 + q̇sw , (A.63)
Δwm
168
kw,m n
Vl−1 + hc + hr F ∗ 1/4 T∞ − q̇sw
Δwm
Vln+1 = . (A.64)
kw,m
+ [hc + hr ]
Δwm
For zero-flux boundaries, the discretized equation for the upward sweep becomes
n+1
Ul−1 − Uln+1
= 0, (A.65)
Δwm
For constant temperature boundaries, the discretized equation for both sweeps becomes
Tln+1 = Tc . (A.69)
two dimensions, there are two sub-time-steps within each time-step, and the value of θ in Equation
In the first sub-time-step, all the temperatures in the r terms represent the new sub-time-step,
n + 1/2, and all the temperatures in the z represent the previous sub-time-step, n:
= [2 − f ] cr,p Ti+1,j − Ti,j + cr,m Ti,j
n+1/2 n+1/2 n+1/2 n+1/2 n+1/2
Ti,j − Ti,j
n
− Ti−1,j
n+1/2 n+1/2 n+1/2 n+1/2 (A.70)
+ cr,p Ti+1,j − Ti,j + cr,m Ti,j − Ti−1,j
n n
+ f cz,p Ti,j+1 − Ti,j
n
+ cz,m Ti,j − Ti,j−1
n
+ θq̇.
This equation can be rearranged with all the unknown temperatures at the new sub-time-
step, n + 1/2, on the left-hand side of the equation with constant coefficients and all the constant
n+1/2
Ti,j 1 + [2 − f ] cr,p + cr,p − cr,m − cr,m
(A.71)
+ Ti−1,j [2 − f ] cr,m + cr,m + Ti+1,j [2 − f ] −cr,p − cr,p
n+1/2 n+1/2
n
= Ti,j [1 + f [cz,m − cz,p ]] − Ti,j−1
n n
f [cz,m ] + Ti,j−1 f [−cz,p ] + θq̇
In the second sub-time-step, all the temperatures in the z terms represent the new sub-time-
step, n + 1, and all the temperatures in the r represent the previous sub-time-step, n + 1/2:
= f cr,p Ti+1,j − Ti,j + cr,m Ti,j
n+1 n+1/2 n+1/2 n+1/2 n+1/2 n+1/2
Ti,j − Ti,j − Ti−1,j
n+1/2 n+1/2 n+1/2 n+1/2 (A.72)
+ cr,p Ti+1,j − Ti,j + cr,m Ti,j − Ti−1,j
n+1
+ [2 − f ] cz,p Ti,j+1 n+1
− Ti,j n+1
+ cz,m Ti,j n+1
− Ti,j−1 + θq̇.
This equation can be rearranged with all the unknown temperatures at the new sub-time-
step, n + 1, on the left-hand side of the equation with constant coefficients and all the constant
n+1
Ti,j n+1
[1 + [2 − f ] [cz,p − cz,m ]] + Ti,j−1 n+1
[2 − f ] [cz,m ] + Ti,j−1 [2 − f ] [−cz,p ]
1 + f cr,m + cr,m − cr,p − cr,p
n+1/2 (A.73)
= Ti,j
n+1/2 n+1/2
− Ti−1,j f cr,m + cr,m + Ti+1,j f cr,p + cr,p + θq̇
The boundary conditions for the ADI scheme are identical to those of either the explicit
scheme or the fully implicit scheme depending on the orientation of the boundary relative to the
implicit direction, r or z, for the sub-time-step. When the implicit direction is perpendicular to
the boundary the conditions are fully implicit, and when the implicit direction is parallel to the
The steady-state finite difference formulation, used for some of the initialization methods in
cr,p [Ti+1,j − Ti,j ] + cr,m [Ti,j − Ti−1,j ] + cr,p [Ti+1,j − Ti,j ] (A.76)
+ cr,m [Ti,j − Ti−1,j ] + cz,p [Ti,j+1 − Ti,j ] + cz,m [Ti,j − Ti,j−1 ] + θq̇ = 0.
Each of the constant coefficients contains θ, which can be divided out of the equation, but it
is kept in the equation for consistency with the other finite difference scheme descriptions.
This equation can be rearranged with all the unknown temperatures on the left-hand side of
the equation with constant coefficients and all the constant terms on the right-hand side:
Ti,j cr,m + cr,m + cz,m − cr,p − cr,p − cz,p + Ti−1,j −cr,m − cr,m
(A.77)
+ Ti+1,j cr,p + cr,p + Ti,j−1 [−cz,m ] + Ti,j+1 [cz,p ] = −θq̇
The boundary conditions for the steady-state solution are identical to those of the fully
implicit scheme.
171
The system of equations generated for the set of discretized cells is generalized as
• n is equation index,
algebra techniques need be applied. For implicit schemes, where there are potentially multiple
unknowns in an equation, the linear system is solved either directly, using Gaussian elimination, or
iteratively. The ADI method uses a simplified form of Gaussian elimination specific to tridiagonal
matrices (each equation has up to three consecutive non-zero coefficients centered at n, n).
Appendix B
The following results were simulated according to the two-dimensional approximation method
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure B.1: Percent difference in annual heat transfer between three-dimensional square foundation
173
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure B.2: Percent difference in annual heat transfer between three-dimensional cross-shaped
174
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure B.3: Percent difference in annual heat transfer between three-dimensional narrow-H foun-
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure B.4: Percent difference in annual heat transfer between three-dimensional wide-H foundation
176
Appendix C
The following results were simulated according to the two-dimensional approximation method
A=P = 1.5 m
10%
0%
-10%
Deviation from mean heat flux
-20%
20%
A=P = 4.6 m
10%
0%
-10%
-20%
20%
A=P = 13.7 m
10%
0%
-10%
-20%
Circle Semi-Infinite Rectangle Cross
Square Wide-H Narrow-H
Figure C.1: Comparison of heat flux from a range of uninsulated foundation shapes in the cold
climate
Reduction in heat flux relative to uninsulated foundations
40%
40%
40%
0%
20%
0%
20%
0%
20%
60%
60%
60%
80%
80%
80%
100%
100%
100%
R-5 ext. hor.
Figure C.2: Heat flux reduction relative to uninsulated foundations in mixed climate for a range of
179
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure C.3: Percent difference in annual heat transfer between three-dimensional square foundation
180
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure C.4: Percent difference in annual heat transfer between three-dimensional cross-shaped
181
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Boundary Layer
R-20 half-wall ext.
Basement
Figure C.5: Percent difference in annual heat transfer between three-dimensional narrow-H foun-
182
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure C.6: Percent difference in annual heat transfer between three-dimensional wide-H foundation
Appendix D
The following results were simulated according to the two-dimensional approximation method
A=P = 1.5 m
10%
0%
-10%
Deviation from mean heat flux
-20%
20%
A=P = 4.6 m
10%
0%
-10%
-20%
20%
A=P = 13.7 m
10%
0%
-10%
-20%
Circle Semi-Infinite Rectangle Cross
Square Wide-H Narrow-H
Figure D.1: Comparison of heat flux from a range of uninsulated foundation shapes in the hot
climate
Reduction in heat flux relative to uninsulated foundations
40%
40%
40%
0%
20%
0%
20%
0%
20%
60%
60%
60%
80%
80%
80%
100%
100%
100%
R-5 ext. hor.
Figure D.2: Heat flux reduction relative to uninsulated foundations in mixed climate for a range
186
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure D.3: Percent difference in annual heat transfer between three-dimensional square foundation
187
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure D.4: Percent difference in annual heat transfer between three-dimensional cross-shaped
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Boundary Layer
R-20 half-wall ext.
Basement
Figure D.5: Percent difference in annual heat transfer between three-dimensional narrow-H foun-
Deviation from three-dimensional annual heat transfer
-5%
0%
5%
-5%
20%
0%
5%
-5%
20%
0%
5%
20%
-20%
-15%
-10%
-20%
-15%
-10%
-20%
-15%
-10%
10%
15%
10%
15%
10%
15%
Uninsulated
Circle
R-5 int. whole-slab
Slab
R-20 ext. hor.
Infinite Rectangle
Uninsulated
Rounded Rectangle
R-20 int. vert. & hor.
Uninsulated
Boundary Layer
R-20 half-wall ext.
Basement
Figure D.6: Percent difference in annual heat transfer between three-dimensional wide-H foundation
190
Appendix E
The following figures compare simulated annual minimum and maximum heat transfer rates
results. Results were simulated according to the two-dimensional approximation method test de-
5 5
Circle Infinite Rectangle
0 0
Three-dimensional annual minimum heat transfer [kW]
−5 −5
−10 −10
−15 −15
−20 −20
5 5
Rounded Rectangle Boundary Layer
0 0 Adjustment
−5 −5
−10 −10
−15 −15
−20 −20
Figure E.1: Comparison of annual minimum heat transfer between three-dimensions and two-
dimensional approximations for mixed climate simulations
193
60 60
Circle Infinite Rectangle
50 50
Three-dimensional annual maximum heat transfer [kW]
40 40
30 30
20 20
60 60
Rounded Rectangle Boundary Layer
50 50 Adjustment
40 40
30 30
20 20
Figure E.2: Comparison of annual maximum heat transfer between three-dimensions and two-
dimensional approximations for mixed climate simulations
194
30 30
Circle Infinite Rectangle
25 25
Three-dimensional annual minimum heat transfer [kW]
20 20
15 15
10 10
30 30
Rounded Rectangle Boundary Layer
25 25 Adjustment
20 20
15 15
10 10
Figure E.3: Comparison of annual minimum heat transfer between three-dimensions and two-
dimensional approximations for cold climate simulations
195
120 120
Circle Infinite Rectangle
100 100
Three-dimensional annual maximum heat transfer [kW]
80 80
60 60
40 40
120 120
Rounded Rectangle Boundary Layer
100 100 Adjustment
80 80
60 60
40 40
Figure E.4: Comparison of annual maximum heat transfer between three-dimensions and two-
dimensional approximations for cold climate simulations
196
0 0
Circle Infinite Rectangle
−10 −10
Three-dimensional annual minimum heat transfer [kW]
−20 −20
−30 −30
−40 −40
0 0
Rounded Rectangle Boundary Layer
−10 −10 Adjustment
−20 −20
−30 −30
−40 −40
Figure E.5: Comparison of annual minimum heat transfer between three-dimensions and two-
dimensional approximations for hot climate simulations
197
25 25
Circle Infinite Rectangle
20 20
Three-dimensional annual maximum heat transfer [kW]
15 15
10 10
5 5
25 25
Rounded Rectangle Boundary Layer
20 20 Adjustment
15 15
10 10
5 5
Figure E.6: Comparison of annual maximum heat transfer between three-dimensions and two-
dimensional approximations for hot climate simulations