modeling radiant heating and cooling systems: integration with a whole-building simulation program

9
Modeling radiant heating and cooling systems: integration with a whole-building simulation program R.K. Strand*, K.T. Baumgartner School of Architecture, University of Illinois at Urbana-Champaign, Urbana-Champaign, IL 61820, USA Received 12 May 2004; received in revised form 13 July 2004; accepted 30 July 2004 Abstract One of the primary goals of the radiant heating and cooling community for the past two decades has been the fair and accurate comparison of radiant space conditioning systems with conventional forced air systems. The desire of the community was to give architects and engineers the ability to compare the energy consumption of different system types while requiring the systems to produce equivalent thermal environments. Unfortunately, radiant systems, though relatively simple in concept, are relatively complex from a modeling standpoint. A rigorous model must include fundamental laws of thermodynamics and properly account for heat transfer due to radiation, convection, and conduction. In addition, the model must do this in a generic way within a program that is capable of modeling buildings of any size and type and their associated HVAC systems. Such challenges are not easily overcome and require careful attention to detail as well as a solid foundation from which to build. This paper summarizes a model which addresses these concerns and provides a radiant system model within a popular building energy simulation program. # 2004 Elsevier B.V. All rights reserved. Keywords: Simulation programs; HVAC systems; Heating and cooling systems 1. Introduction Determining the energy efficiency of a building before it is even built requires sophisticated science. While simula- tion has the potential to reduce and has been shown to be successful at reducing energy consumption, energy costs, and the impact of buildings on the environment, the process through which a simulation program has to go to simulate a building is inherently complex. Though computer technol- ogy certainly helps architects and engineers have better access to improved predictions of energy efficiency, the fact remains that simulation programs are, by their nature, complex and detail oriented. This is evident in the processes that take place within the building envelope as well as other parts of the building. The building envelope transmits heat to and from the inside and outside environments that are experienced by a building. Modeling a building properly requires that a simulation program adequately addresses fundamental heat transfer processes such as conduction through the solid materials in the building envelope, convection between the envelope and the surrounding air, and radiation between the envelope and the inside and outside environments. None of these processes are simple linear functions, and most are dependent not only on the material properties of the envelope but also the inside and outside environmental conditions. While some programs are capable of handling most heat transfer effects for general building types and materials, there are also challenges to modeling the way a building is heated or cooled. Most modern buildings have some form of HVAC system which has the function of maintaining thermally comfortable conditions inside the building regardless of the conditions on the outside of the building. One of the more common methods of maintaining thermal conditions within a building is through the use of conventional forced air systems. These systems aim to deliver conditioned air to a space to maintain a particular air temperature since air temperature is one of the main factors that determines thermal comfort. While in the case of most www.elsevier.com/locate/enbuild Energy and Buildings 37 (2005) 389–397 * Corresponding author. E-mail address: [email protected] (R.K. Strand). 0378-7788/$ – see front matter # 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.enbuild.2004.07.009

Upload: rk-strand

Post on 26-Jun-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

www.elsevier.com/locate/enbuild

Energy and Buildings 37 (2005) 389–397

Modeling radiant heating and cooling systems: integration with

a whole-building simulation program

R.K. Strand*, K.T. Baumgartner

School of Architecture, University of Illinois at Urbana-Champaign, Urbana-Champaign, IL 61820, USA

Received 12 May 2004; received in revised form 13 July 2004; accepted 30 July 2004

Abstract

One of the primary goals of the radiant heating and cooling community for the past two decades has been the fair and accurate comparison

of radiant space conditioning systems with conventional forced air systems. The desire of the community was to give architects and engineers

the ability to compare the energy consumption of different system types while requiring the systems to produce equivalent thermal

environments. Unfortunately, radiant systems, though relatively simple in concept, are relatively complex from a modeling standpoint. A

rigorous model must include fundamental laws of thermodynamics and properly account for heat transfer due to radiation, convection, and

conduction. In addition, the model must do this in a generic way within a program that is capable of modeling buildings of any size and type

and their associated HVAC systems. Such challenges are not easily overcome and require careful attention to detail as well as a solid

foundation from which to build. This paper summarizes a model which addresses these concerns and provides a radiant system model within a

popular building energy simulation program.

# 2004 Elsevier B.V. All rights reserved.

Keywords: Simulation programs; HVAC systems; Heating and cooling systems

1. Introduction

Determining the energy efficiency of a building before it

is even built requires sophisticated science. While simula-

tion has the potential to reduce and has been shown to be

successful at reducing energy consumption, energy costs,

and the impact of buildings on the environment, the process

through which a simulation program has to go to simulate a

building is inherently complex. Though computer technol-

ogy certainly helps architects and engineers have better

access to improved predictions of energy efficiency, the fact

remains that simulation programs are, by their nature,

complex and detail oriented.

This is evident in the processes that take place within the

building envelope as well as other parts of the building. The

building envelope transmits heat to and from the inside and

outside environments that are experienced by a building.

Modeling a building properly requires that a simulation

* Corresponding author.

E-mail address: [email protected] (R.K. Strand).

0378-7788/$ – see front matter # 2004 Elsevier B.V. All rights reserved.

doi:10.1016/j.enbuild.2004.07.009

program adequately addresses fundamental heat transfer

processes such as conduction through the solid materials in

the building envelope, convection between the envelope and

the surrounding air, and radiation between the envelope and

the inside and outside environments. None of these

processes are simple linear functions, and most are

dependent not only on the material properties of the

envelope but also the inside and outside environmental

conditions. While some programs are capable of handling

most heat transfer effects for general building types and

materials, there are also challenges to modeling the way a

building is heated or cooled. Most modern buildings have

some form of HVAC system which has the function of

maintaining thermally comfortable conditions inside the

building regardless of the conditions on the outside of the

building. One of the more common methods of maintaining

thermal conditions within a building is through the use of

conventional forced air systems. These systems aim to

deliver conditioned air to a space to maintain a particular air

temperature since air temperature is one of the main factors

that determines thermal comfort. While in the case of most

Page 2: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397390

conventional forced air systems many of the system

interactions have been well-studied and can be modeled,

there are a wide variety of different system types and

applications. To be able to evaluate how well and efficiently

different system types will condition a building requires a

program that is capable of modeling various components as

well as various reasonable configurations of those compo-

nents.

Radiant heating and cooling systems pose additional

problems to simulation programs. The idea of using

radiation to heat and cool environments is not new and is

becoming more commonplace in Asia and Europe and to a

lesser extent in North America [8]. Radiant systems aim to

provide conditioning directly to space occupants by radiant

exchange between ‘‘conditioned’’ surfaces within the space

rather than indirectly via transferring conditioned air to the

space which then provides comfort to the occupants via

convection.

While the concept of a radiant system seems relatively

simple—directly transfer heat to or from people—there are

significant challenges to properly modeling a radiant system

and incorporating such a model into a whole-building energy

simulation program. A radiant system, by definition, is both

an HVAC system as well as a surface within the space. This

means that it is part of the calculation to determine how

much heating or cooling is needed as well as the system that

actually responds to those space conditions. Since some

radiant systems have thermal mass associated with them,

controls and transient behavior of the system are critical

components of a realistic model. Moreover, since radiant

systems condition in a significantly different way in

comparison to standard conventional forced air systems,

the system model must be able to evaluate thermal comfort

in the space and also account for the possibility of con-

densation when the system is used for cooling. This high

level of integration must then also account for a generic

simulation environment that is established to ensure the

ability to simulate buildings of many different shapes,

orientations, sizes, etc.

It is also important that the resulting simulation not be so

‘‘radiant-centric’’ that it cannot model more conventional

systems or combined load-sharing systems that are a hybrid

combination of radiant and forced air systems. One of the

most significant barriers to the specification of radiant

systems in buildings is the inability to fairly compare the

radiant and conventional systems to each other or to

combinations of the two. This is really the most pressing

need of the radiant community today—to be able to simulate

various system types and make energy and/or cost

comparisons based on reliable simulation data.

The remainder of this paper focuses on the incorporation

of radiant system models into a whole-building energy

analysis program sponsored by the US Department of

Energy. It will begin with a brief description of the

capabilities of the program that are important for the proper

modeling of radiant systems. It will then delve into the

mathematical background of the algorithm and controls of

two distinct radiant system types: flow controlled systems

and temperature controlled (constant flow) systems.

2. Modeling radiant systems

As was briefly mentioned in the preceding section, a

program that will accurately model radiant systems and

compare them to other system types based on equivalent

thermal (comfort) environments needs to be able to analyze

a building from fundamental heat transfer and thermo-

dynamic perspectives, the HVAC system, and the inter-

connections between the two. In addition, to be successful, it

needs to be general enough to handle any building type,

shape, or size and do so in a relatively efficient manner.

The program [3] chosen for implementing the radiant

system model has incorporated all of the features needed to

successful simulate buildings and conventional systems. In

addition, it has features that make it an ideal application for the

integration of a radiant system model. From a heat transfer

perspective, it uses a fundamental heat balance approach [7]

that applies the First Law of Thermodynamics to three key

interfaces: the inside surfaces for walls, roofs, floors, etc.; the

outside surfaces for walls, roofs, floors, etc.; and the air

contained within the various zones in the building. The inside

and outside heat balances are done for individual surfaces

within a building and incorporate fundamental approaches for

handling convection, long wavelength radiation, and short

wavelength radiation and their impact on each surface.

Moreover, heat transfer through each surface is modeled using

a solution to the transient conduction equation known as

conduction transfer functions (CTFs).

Another key feature of the program is the integrated

nature of the simulation [13]. In the program, the impact of

the HVAC system on various spaces is calculated based on

the real capacity of the system. This means that when a

system is not sized large enough to meet all of the heating or

cooling needs for the spaces it is conditioning, the program

will simulate the deviation from the normal control band as

it would occur in a real building. This type of simulation

algorithm is particularly important when dealing with a

radiant system simulation since the radiant system is both a

surface within the building and an HVAC system. Many

more details about the program [4] and the initial

implementation of the radiant system model [10,11] and

thermal comfort models [6] in the program are available in

the literature. The main focus of this paper is two-fold: the

interconnection between the building and HVAC portions of

the simulation and how this connection is affected by the

type of controls used with the radiant system.

There are two main categories of radiant system controls

[1]: variable flow/fixed temperature systems and fixed flow/

variable temperature systems. The variable flow/fixed

temperature system has a central loop and pump that

supplies water at a particular temperature to all of the

Page 3: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397 391

Fig. 1. Variable flow low temperature radiant system controls.

Fig. 2. Variable temperature low temperature radiant system controls.

systems connected to the loop. Valves control how much

water flows through each portion of the system based on

some criteria relating back to conditions within the space.

The fixed flow/variable temperature system has a central

loop and pump but also has local pumps associated with each

portion of the system. The local pumps maintain a constant

flow rate through that portion of the system while valves

control the amount of recirculation in the local loops. A

central pump maintains flow so that there is enough fluid to

inject to each local loop to maintain the proper temperatures

at each local loop. While some systems have several local

loops or perhaps use slight variations on these main control

strategies, most systems can be categorized into one of these

two classes.

Regardless of which type of system (variable temperature

or variable flow) is being simulated, both systems are

essentially a heat exchanger between the primary/secondary

equipment loop and the building elements (floors, walls,

ceilings, etc.). The primary/secondary side of the model is

simply a fluid loop that receives conditioned water from a

boiler, chiller, heat pump, etc. and then sends unconditioned

water back to the primary equipment. The building element

portion of the model connects to the heat balance of the

surfaces via a modification of the conduction transfer function

(CTF) formulation that accepts heat sources and sinks [9]. This

link is critical because the two sides of the radiant system

are simulated in very different areas of the program code. The

next two sections discuss the details behind this integration

and the mathematical derivation that was necessary to make

the model generic enough to both fit in the program structure

and model as many systems as possible and to make the model

efficient by avoiding unnecessary iterations that can cause

excessive program runtimes.

2.1. Simulation control overview

The controls for variable flow low temperature radiant

systems within the simulation program are fairly simple

though there is some flexibility through the use of schedules.

The program user is allowed to define a setpoint temperature

as well as a throttling range through which the system varies

the flow rate of water (or current) to the system from zero to

the user defined maximum flow rate. The flow rate is varied

linearly with the flow reaching 50% of the maximum when

the controlling temperature reaches the setpoint tempera-

ture. Setpoint temperatures can be varied on an hourly basis

throughout the year if desired. The controlling temperature

can be the mean air temperature, the mean radiant tem-

perature, the operative temperature of the zone, the outdoor

dry-bulb temperature, or the outdoor wet-bulb temperature,

and this choice is also left to the user’s discretion. Since

flow rate is varied, there is no explicit control on the inlet

water temperature or mixing to achieve some inlet water

temperature in the variable flow system. However, the user

does have the ability to specify on an hourly basis through a

schedule the temperature of the water that would be supplied

to the variable flow radiant system. Graphical descriptions of

the controls for the low temperature radiant system model in

the program are shown in Fig. 1 for a variable flow radiant

system.

In the constant flow-variable temperature systems, the

controls are also considered piecewise linear functions, but

in this case the user selects both the control temperatures and

the water temperatures via schedules. This offers greater

flexibility for defining how the radiant system operates

though it may not model every situation. Fig. 2 shows how

the ‘‘desired’’ inlet water temperature is controlled based on

user schedules. The user has the ability to specify the high

and low water and control temperature schedules for heating

and cooling separately (a total of eight temperature

schedules). Note that this inlet temperature is a ‘‘desired’’

inlet temperature in that there is no guarantee that the system

will provide water to the system at that temperature. The

model includes a local loop that attempts to meet this

demand temperature through mixing and recirculation.

Page 4: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397392

Fig. 3. Variable temperature low temperature radiant system component

details.

The constant flow (variable temperature) low temperature

radiant system model is actually a combination of mixing

valves, a pump (constant speed, but the maximum flow can

be modified by a schedule), and the radiant system (surface,

panel, or group of surfaces/panels). This is connected to the

main loop through the standard inlet connections as shown

in Fig. 3. The system controls determine the desired inlet

temperature and system flow rate while loop controls

determine the flow rate and temperature of the loop. Note

that pump heat also factors into the model through a simple

constant speed pump model and user input.

Control of the variable flow system is considerably easier

because it simply accepts whatever temperature water is

available and varies the flow rate based on the controls

defined by Fig. 1. There is also no recirculation to account

for nor is there a local pump that adds heat to the loop.

Control of the variable temperature systems is more

challenging because of the desire to control temperature.

There are four possible conditions (separate for heating

and cooling) for the variable temperature radiant system.

First, if the loop has adequate temperature and flow to meet

system requests, then the model sets the radiant system inlet

temperature and controls to the desired values based on the

controls and simulates. This is the best condition and

recirculation and bypass amounts are adjusted accordingly

based on radiant system outlet temperatures. Second, if the

loop temperature is adequate but the loop flow rate is less

than the radiant system flow rate, we may or may not be able

to meet the desired inlet temperature since recirculation

might lower the temperature below the desired temperature.

In this second case, the model first simulates the radiant

system with the desired conditions and then resimulates it to

solve for the actual inlet temperature (see later in this

section) if it cannot achieve the desired inlet temperature.

Third, if the loop flow is greater than the radiant flow but the

temperature of the loop is not adequate, then there is no

amount of mixing that will solve this problem. All of the

radiant flow comes from the loop and the loop temperature

(after pump heat addition) becomes the radiant system inlet

regardless of the temperature controls. Finally, if both the

temperature and the flow of the loop are inadequate, then the

model simply solves for the actual radiant system inlet

temperature and does not try to meet the controls (merely

tries to get as close as physically possible given the loop

conditions).

Note that Fig. 3 shows that multiple surfaces can be

attached to a particular local loop. This option is available

for both the variable flow and variable temperature hydronic

radiant systems in the program. This is intended to account

for potential situations where a local loop serving a

particular zone has multiple active surfaces within that

zone. For example, both the floor and ceiling of a space may

serve as part of the radiant system, several different ceiling

surfaces may also be linked together, etc.

2.2. Modeling the connection between the heat balance

and HVAC system

As was mentioned earlier, the radiant system itself is a

heat exchanger with a ‘‘stationary’’ fluid (i.e., the building

element) assumed to be at a single temperature on one side

and a water loop on the other. The stationary fluid side

relates to the heat balance portion of the program and must

be integrated with the surface heat balances. The water loop

side is part of the primary plant simulation. A heat exchanger

formulation merges the two sides as described below.

As shown by [9], the standard conduction transfer

function formulation of heat conduction through a surface

can be adapted to include the effect of an internal heat

source/sink. This leads to in the following equation which is

used for the inside surface heat balance:

q00i;t ¼XM

m¼1

XmTi;t�mþ1 �XM

m¼1

YmTo;t�mþ1

þXk

m¼1

Fmq00i;t�m þXM

m¼1

Wmqsource;t�mþ1 (1)

A similar equation can be established for the outside surface

heat balance:

q00o;t ¼XM

m¼1

YmTi;t�mþ1 �XM

m¼1

ZmTo;t�mþ1

þXk

m¼1

Fmq00o;t�m þXM

m¼1

Vmqsource;t�mþ1 (2)

Additionally, we can characterize the temperature at the

source/sink location with a similar equation:

Ts;t ¼XM

m¼1

xmTi;t�mþ1 �XM

m¼1

ymTo;t�mþ1

þXk

m¼1

fmTs;t�m þXM

m¼1

wmqsource;t�mþ1 (3)

In this series of equations, X, x, Y, y, Z, z, F, and f are

conduction transfer function (CTF) coefficients that are

calculated once at the beginning of a simulation for each

unique surface construction. The V, W, and w terms are heat

Page 5: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397 393

source transfer function (QTF) coefficients that allow the

evaluation of the response of a building element to an

embedded heat source or sink. Coefficients, temperatures,

and fluxes at time ‘‘t’’ are for the current moment in time

while terms at other times are history terms—values

from past time steps which are known as a result of the

simulation.

Eqs. (1)–(3) serve as a starting point for the actual heat

balance solution. These equations alone, however, are not

enough information to solve for what is happening in the

radiant system at a given point in time. The heat source/

sink term is not known directly for a radiant system since

this relates back to the inlet water temperature, the water

flow rate, and the temperature of the radiant surface. The

actual heat transferred between the building element and

the water loop is related to the temperature of the building

element at the source location as well as the water system

inlet and outlet temperatures. In the model, it is assumed

that the inlet temperature to the slab (defined by a user

schedule and/or controls and the primary equipment

simulation) and the mass flow rate (determined by the

control algorithm and/or user input) are known and that the

remaining parameters must be calculated. However, the

heat balance equations require the heat transferred to the

building element from the water loop in order to calculate

the heat transferred from the element to the building

environment.

Even though systems defined by this model can vary

somewhat, the same characteristic link between the system

variables exist. For modeling purposes, the overall water/

slab system can be thought of as a heat exchanger. While

in principle there are two alternative heat exchanger

methodologies, it is more convenient to use the effective-

ness-NTU method in this case. Several assumptions will be

incorporated into the heat exchanger analysis. It is assumed

that the building element that contains the water loop is

stationary and that its temperature along the length of the

tubing is constant. The latter part of this assumption stems

from assumptions made in both the one- and two-

dimensional heat source transfer function derivations. In

either case, the source was added at a single node that

was characterized by a single temperature. For consistency,

this assumption must be made again in the heat exchanger

analysis. Another assumption for the model is that the

fluid in the tubing is water. Additionally, it is assumed

that the thermal properties of the water do not vary

significantly over the length of the tubing. Finally, the

temperature at the inside surface of the water tubing is

assumed to be equal to the temperature at the source

location. In other words, it is assumed that the water tubing

itself has no appreciable effect on the heat transfer process

being modeled.

Using these assumptions and the effectiveness-NTU heat

exchanger algorithm, several equations can be defined which

establish the relationship between the heat source and the

water temperatures. First, a heat balance on the water loop

results in:

q ¼ ðmcpÞwaterðTwi � TwoÞ (4)

where q is the energy transferred between the water

loop and the building element, m the mass flow rate of

the water, cp the specific heat of the water, Twi the

inlet water temperature, and Two the outlet water tempera-

ture.

The maximum amount of heat transfer that is possible

according to the Second Law of Thermodynamics is:

qmax ¼ ðmcpÞwaterðTwi � TsÞ (5)

where qmax is the maximum amount of energy transfer that is

possible and Ts the temperature at the source location.

The effectiveness of the heat exchanger, e, is defined as

the ratio of the actual energy transfer to the maximum

possible, or:

e� q

qmax(6)

For a heat exchanger where one fluid is stationary, the

effectiveness can be related to NTU, the number of transfer

units, by the following equation [5]:

e ¼ 1 � e�NTU (7)

where NTU is defined by:

NTU� UA

ðmcpÞwater

(8)

Since the water tubes were assumed to have no effect on the

heat transfer process, the only term present in the overall

heat transfer coefficient, UA, is a convection term. Thus, the

equation for UA is:

UA ¼ hðpDLÞ (9)

where h is the convection coefficient, D the interior tube

diameter, and L the total length of the tube.

The convection coefficient can be obtained from internal

flow correlations that relate the Nusselt dimensionless

number to other flow properties. For laminar flow in a tube of

constant surface temperature, the Nusselt number is defined

by:

NuD ¼ hD

k(10)

where k is the thermal conductivity of the water.

For turbulent internal flow, the Colburn equation can be

used to define the Nusselt number:

NuD ¼ hD

k¼ 0:023 Re

4=5D Pr1=3 (11)

where Pr is the Prandtl number of water and ReD the

Reynolds number which is defined by:

ReD ¼ 4m

pmD(12)

Page 6: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397394

The parameter m is the absolute viscosity of water. For

internal pipe flow, the flow is assumed to be turbulent for

ReD � 2300.

Knowledge of the flow conditions allows Eqs. (6)–(12) to

be calculated. This fixes the value for e which can then be

used to relate Eqs. (4) and (5) to each other. The controls and

the plant define the water mass flow rate and the inlet water

temperature, leaving two equations (Eqs. (4) and (5)) and

three unknowns. The third equation that can be used in

conjunction with Eqs. (4) and (5) is Eq. (3), which is the CTF/

QTF equation for the temperature at the source location.

Unfortunately, it has two additional unknowns—the inside

and outside surface temperatures. This requires the addition

of modified forms of Eqs. (1) and (2) which adds two more

equations without the introduction of any further unknowns.

Eq. (1) is the standard conduction transfer function

formula for the inside surface of a building element with an

embedded source/sink of heat. In the actual simulation

program, the surface flux on the left hand side of the

equation is replaced with a surface heat balance:

surface heat balance ¼XM

m¼1

XmTi;t�mþ1 �XM

m¼1

YmTo;t�mþ1

þXk

m¼1

Fmq00i;t�m þXM

m¼1

Wmqsource;t�mþ1

(13)

The surface heat balance includes terms for incident solar

energy, radiation heat transfer from internal heat sources

such as lights and electrical equipment, radiation between

surfaces using Hottel’s Gray Interchange concept, and con-

vection to the surrounding air. A similar equation can be

derived from Eq. (2) for the outside surface heat balance and

includes thermal forces such as solar radiation, convection to

the outside air, long wavelength radiant exchange with the

external environment, etc. It should be noted that the internal

radiation heat balance is dependent on conditions at the

other surfaces within the space. Thus, an iterative loop is

required to provide a more accurate estimate of the radiative

exchange within the building interior. Thus, the terms of the

heat balance on the left hand side of the equation have been

set with the only unknown quantity being Ti, the inside

surface temperature at the current time step. On the right

hand side of Eq. (13), most of the terms are already defined

since they depend on known values from previous time steps

(temperature, flux, and source histories). The only terms

which are not defined are the inside surface temperature (Ti),

outside surface temperature (To), and internal heat source/

sink (qsource) of the current time step. Similar arguments can

be made for reformulation of the outside heat balance

equation and the base form of the heat source/sink location

temperature equation (Eq. (3))—the only undefined terms

are the inside surface temperature (Ti), outside surface

temperature (To), and internal heat source/sink (qsource) of

the current time step.

One can summarize the situation by hiding some of the

complexity for the purpose of deriving equations to solve the

system of equations at hand. The three basic equations for

the temperature at the inside and outside surface as well as at

the location of the heat source/sink can be reduced to:

Tinside ¼ Ca þ CbToutside þ Ccq00 (14)

Toutside ¼ Cd þ CeTinside þ Cfq00 (15)

Tsource ¼ Cg þ Chq00 þ CiTinside þ CjToutside (16)

where Tinside is the current temperature at the inside surface,

Toutside the current temperature at the outside surface, Tsource

the current temperature within the radiant system at the

location of the source/sink, q00 the heat source/sink from the

radiant system water loop, Ca is all of the other terms in the

inside heat balance (solar, LW exchange, conduction history

terms, etc.), Cb the current cross CTF term, Cc the QTF

inside term for the current heat source/sink, Cd is all of the

other terms in the outside heat balance (solar, LW exchange,

conduction history terms, etc.), Ce the current cross CTF

term, Cf the QTF outside term for the current heat source/

sink, Cg the summation of all temperature and source history

terms at the source/sink location, Ch the QTF term at the

source/sink location for the current heat source/sink, Ci the

CTF inside term for the current inside surface temperature,

Cj the CTF outside term for the current outside surface

temperature.

Eqs. (14) and (15) above can each be solved to remove the

other surface temperature. Then, substituting these new

equations for Tinside and Toutside as a function of C and q00 into

the equation for Tsource and simplifying, one obtains the

following equation:

Tsource ¼ Ck þ C1q00 (17)

Ck ¼ Cg þCiðCa þ CbCdÞ þ CjðCd þ CeCaÞ

1 � CeCb(18)

Cl ¼ Ch þCiðCc þ CbCfÞ þ CjðCf þ CeCcÞ

1 � CeCb(19)

Combining this with the heat exchanger analysis as shown

above, we eventually arrive at the following equation to

relate the flux to the slab to the water inlet temperature and

mass flow rate:

q00 ¼ Twater;in � Ck

ðCl=AÞ þ ð1=eðmcpÞwaterÞ(20)

which includes all of the inside and outside heat balance

terms (‘‘hidden’’ in the Ck and Cl coefficients). Once the flux

to the slab is known, the remaining terms of interest (outlet

water temperature, inside and outside surface temperatures,

etc.) can be calculated using the relatively simpler equations

shown above.

The above development is valid for both the variable flow

radiant system and for variable temperature radiant systems

Page 7: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397 395

where the inlet temperature is known (based on controls).

However, due to loop conditions and the presence of

recirculation, it is not always possible to know the inlet

temperature to the variable temperature radiant system

without actually simulating it. For cases where the main loop

temperature and flow rates are adequate, one can assume an

inlet temperature and then determine the recirculation

needed after the fact. Yet, when the main loop flow or

temperature is not adequate to meet the needs of the local

loop as determined by the system controls, one is faced with

a ‘‘chicken-and-egg’’ situation. One needs to know the inlet

temperature in order to solve for the system output, but one

needs the system output in order to solve for the inlet

temperature. As a result, the model must either iterate using

a guess for inlet temperature as a starting point or perform

more mathematics to arrive at the inlet temperature as a

function of system characteristics. This implementation

chose to avoid iteration and solved for the inlet temperature

as shown below.

In addition to resolving the iteration concerns, one must

also account for the fact that there could be several potential

parts to a particular local loop as shown in Fig. 3. Each

of these individual surfaces may have a different outlet

temperature. Thus, all of the surfaces that are part of an

individual loop must be solved at the same time to obtain

both the combined outlet temperature for the local loop

and thus the inlet temperature based on the amount of

recirculation needed.

The previous equation combines with the following

equation which is valid for an individual surface in the

current local loop of a radiant system:

qj ¼ mjcpðTwater � Twater;out;jÞ (21)

where qj is the heat transfer to the jth surface in the radiant

system, mj the mass flow rate only to this surface, and

Twater,out,j the outlet temperature for the jth surface. Combin-

ing the previous two equations results in:

Twater;in � Twater;out;j ¼ejðTwater;in � Ck;jÞ

1 þ ððejmjcpCl;jÞ=AÞ (22)

where Ck,j and Cl,j are evaluated for each individual surface

as shown in Eqs. (18) and (19). If for each surface in the

radiant system, we let:

Cm;j ¼ej

1 þ ððejmjcpCl;jÞ=AÞ (23)

then the previous equations becomes the slightly less com-

plex:

Twater;in � Twater;out;j ¼ Cm;jðTwater;in � Ck;jÞ (24)

Rearranging to obtain the outlet temperature for the jth

surface:

Twater;out;j ¼ ð1 � Cm;jÞTwater;in þ Cm;jCk;j (25)

The overall outlet temperature from the radiant system is just

a simple mixing of all of the surface outlet temperatures

based on flow fraction and results in Twater,out as shown in:

Twater;out ¼X

j

½ðFlowFractionjÞTwater;out;j (26)

An energy balance on the mixing valve-pump group results

in the following equation that relates the radiant system inlet

temperature (Twater,in) to the loop inlet water temperature

(Tloop,in), the radiant system outlet temperature (Twater,out),

and the pump heat addition

Twater;in ¼ mloop

msystemTloop;in þ

mrecirc

msystemTwater;out

þ PumpHeat

msystemcp(27)

Plugging in the definition of Twater,out based on the summa-

tion of Twater,out,j equations results in:

Twater;in ¼ LoopTerm þ RecircTerm

TwiCoeff(28)

where

LoopTerm ¼mloop

msystemTloop;in þ

PumpHeat

msystemcp(29)

mrecircX

RecircTerm ¼msystem j

FlowFractionjTwater;out;j (30)

mrecirc

� �X

TwiCoeff ¼ 1 �

msystem j

½FlowFractionjð1 � Cm;jÞ

(31)

Once the actual water inlet temperature is calculated with

this equation, it is then possible to calculate individual outlet

temperatures for each surface, the overall outlet tempera-

ture, and finally all of the necessary flow and loop quantities.

This procedure avoids iteration but is somewhat complex to

follow. However, this second mathematical process is only

needed for select cases of the variable temperature radiant

system when the inlet temperature is not known explicitly.

With the proper establishment of input data, such cases can

be avoided, though one cannot simply assume this for a

comprehensive energy simulation program.

The proceeding development shows several important

points. First, it is possible to model the connection between

the heat balance and the HVAC portions of the simulation in

a fundamental, generic way. Second, it is possible to handle

this connection without resorting to excessive iterations

despite the intimate link between the two sides of the radiant

system. In the simulation program, iteration does take place,

but the radiant system model as described here does not

result in additional iteration loops that could lead to long

execution times. Finally, the implementation works within

the current structure of the program—meaning that it does

not require special rules for this particular system and that it

automatically has access to any improvement made to either

the heat balance or HVAC portion of the simulation.

Page 8: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397396

2.3. Avoiding undesirable operating conditions

Both the variable flow and variable temperature systems

have checks within their algorithms to avoid operating the

radiant system under what might be considered undesirable

conditions. One check is aimed at avoiding heat additions by

the system when cooling is required and heat removal by the

system when heating is required. While this likely will never

happen in most systems, such conditions might occur during

transitional seasons or changes in controls from heating to

cooling or cooling to heating. Thus, the system compares the

water inlet temperature with the current temperature of the

radiant system where the source/sink term will be applied.

If sending water to the radiant system will produce the

opposite of what is desired from a heat addition/removal

standpoint, the algorithm will interrupt flow to the radiant

system.

Another check is related to the potential for condensation

in a radiant cooling system. If moisture levels within the

space being cooled are high enough either due to indoor or

outdoor conditions (or both), the surface temperature of the

radiant system could drop below the dew-point temperature

of the space. If this occurs, condensation will take place.

This can be a potential problem in radiant cooling systems,

especially in many of the climates found in North America.

Currently, the program compares the dew-point temperature

of the space with the predicted surface temperature of the

radiant system based on what the normal controls dictate

should be happening within the radiant system. If the pre-

dicted surface temperature drops below the dew-point

temperature plus a small temperature tolerance, the

algorithm will stop flow to the radiant system.

2.4. Connecting to primary systems

One of the advantages to having access to the HVAC

portion of the program through the primary plant water loop

is that it allows the ability to try out different equipment and

various different ideas. One idea that has already been

addressed in the literature [12] is the possibility of hooking

the radiant system up to a water loop that is served only by an

evaporative condenser. Recent additions of ground and pond

loops make it possible to investigate the energy efficiency of

using temperate water in a radiant system and determine

whether or not such a system can produce thermally

comfortable conditions inside the building. The beauty of

such a simulation model is such alternative systems can be

tested for a variety of conditions and environments—even

ones that would not be thought of as climates where such

technology would be successful.

3. Concluding remarks

The model described in this paper is already available to

the general public and has been available for some time. It

incorporates all of the algorithm, control, modeling,

and other features described in the preceding sections.

While this is a significant advance toward the ultimate

goal of modeling all features of radiant systems and having

the ability to compare them fairly and accurately to

conventional forced air systems, there are still features that

needed to be added and are the focus of upcoming research

work on the program. The two issues that will likely be

addressed next are the incorporation of a more sophisti-

cated condensation limit for variable temperature radiant

systems and the modeling of ‘‘load-sharing’’ hybrid

systems. Currently, the condensation check is a fairly

blunt strategy that shuts the system down when it perceives

that condensation may occur. This is based on a comparison

between the dew-point temperature and the predicted

surface temperature if the radiant system is running. There

is a 1 8C tolerance that is presently used within the

program. This was chosen so that the system will shut off

when the surface temperature would be close but still

higher than the dew-point temperature. This limit may be

somewhat strict in that some systems that have the radiant

system connected to an evaporative condenser may

actually trigger the condensation limit. Further study is

needed here to determine if this tolerance is, in fact,

appropriate and whether this needs to be a parameter that

the user should be allowed to change. In addition, the

variable temperature systems do not need to interrupt flow

when condensation is predicted to occur. Since they have

the ability to increase the amount of recirculation, it would

be more appropriate to increase the water inlet temperature

first in an attempt to provide some amount of cooling

without causing condensation.

Concerns about condensation often result in the

specification of a ‘‘load-sharing’’ or hybrid system. In such

a system, a radiant system is combined with a conventional

forced air system to provide the conditioning needs of the

space. Generally, the radiant system provides most or all of

the sensible cooling while the air system handles the latent

cooling. While it currently is possible to model a space being

served by both a radiant and a conventional forced air system

within the program, their interaction is serial rather than

parallel. In the current structure of the program, one system

will run to capacity in an effort to meet the current space

conditioning needs of the zone. Then, additional systems

will attempt to meet any remaining load by operating in

series. True ‘‘load-sharing’’ systems are controlled differ-

ently and thus this type of system is the subject of a current

research project.

Despite the continuing work on the radiant model within

the simulation program, the current model has many

capabilities and has been the subject of recent validations

studies [2]. Many current systems can already be modeled

and compared to conventional forced air systems, and the

comparisons can be made on the more accurate basis of

equivalent thermal comfort rather than simply equal air

temperatures.

Page 9: Modeling radiant heating and cooling systems: integration with a whole-building simulation program

R.K. Strand, K.T. Baumgartner / Energy and Buildings 37 (2005) 389–397 397

Acknowledgment

The authors wish to thank the US Department of Energy’s

National Energy Technology Laboratory (NETL) for

funding the work under Grant DE-FC26-01NT41250 which

led to this paper.

References

[1] Invensys, An introduction to invensys radiant ceiling technologies,

July 2002.

[2] C. Chantrasrisalai, V. Ghatti, D.E. Fisher, D.G. Schaetzle, Experi-

mental validation of the EnergyPlus low-temperature radiant simula-

tion, ASHRAE Transactions: Part 2 109 (2003), ASHRAE, Atlanta,

Georgia.

[3] D.B. Crawley, L.K. Lawrie, F.C. Winkelmann, W.F. Buhl, A.E. Erdem,

C.O. Pedersen, R.J. Liesen, D.E. Fisher, The next-generation in

building energy simulation—a glimpse of the future, in: Proceedings

of the Building Simulation’97, vol. II, Prague, Czech Republic,

IBPSA, September 1997, pp. 395–402.

[4] Crawley, B. Drury, L.K. Lawrie, F.C. Winkelmann, W.F. Buhl, C.O.

Pedersen, R.K. Strand, R.J. Liesen, D.E. Fisher, M.J. Witte, R.H.

Henninger, J. Glazer, D. Shirey, EnergyPlus: new, capable, and linked,

in: Proceedings of the 2002 International Green Building Conference,

November 2002, Austin, Texas, US Green Building Council, Washing-

ton, DC, 2002.

[5] F.P. Incropera, D.P. DeWitt, Introduction to Heat Transfer, Wiley, New

York, NY, 1985.

[6] J. Lee, R.K. Strand, An analysis of the effect of the building envelope

on thermal comfort using the EnergyPlus program, in: Proceedings of

the 2001 ACS A (Association of Collegiate Schools of Architecture)

Technology Conference, Austin, TX, July 2001, ACSA, Washington,

DC, 2001.

[7] C.O. Pedersen, R.J. Liesen, R.K. Strand, D.E. Fisher, L. Dong, P.G.

Ellis, Loads toolkitRP-987, ASHRAE, Atlanta, GA, 2001.

[8] K.W. Roth, D. Westphalen, J. Dieckmann, S.D. Hamilton, W. Goet-

zler, Energy consumption characteristics of commercial building

HVAC systems, vol. III: Energy savings potential, TIAX LLC Report

for US Department of Energy Building Technologies Program,

Washington, DC, 2002.

[9] R.K. Strand, Heat source transfer functions and their application to low

temperature radiant heating systems, Ph.D. Dissertation, Department

of Mechanical and Industrial Engineering, University of Illinois at

Urbana-Champaign, 1995.

[10] R.K. Strand, C.O. Pedersen, Implementation of a radiant heating and

cooling model into an integrated building energy analysis program,

ASHRAE Transactions: Part 1 103 (1997).

[11] R.K. Strand, C.O. Pedersen, Modeling radiant systems in an integrated

heat balance based energy simulation program, ASHRAE Transac-

tions: Part 2 108 (2002).

[12] R.K. Strand, Investigation of a condenser-linked radiant cooling

system using a heat balance based energy simulation program, ASH-

RAE Transactions: Part 2 109 (2003).

[13] R.D. Taylor, C.O. Pedersen, D. Fisher, R. Liesen, L. Lawrie, Impact of

simultaneous simulation of building and mechanical systems in heat

balance based energy analysis programs on system response and

control, in: Proceedings of the Building Simulation’91 Conference,

Nice, France, International Building Performance Simulation Asso-

ciation (IBPSA), 1991.