You are on page 1of 19

HVAC&R Research

ISSN: 1078-9669 (Print) 1938-5587 (Online) Journal homepage: https://www.tandfonline.com/loi/uhvc20

Dynamic modeling for vapor compression


systems—Part II: Simulation tutorial

Bryan P. Rasmussen & Bhaskar Shenoy

To cite this article: Bryan P. Rasmussen & Bhaskar Shenoy (2012) Dynamic modeling for vapor
compression systems—Part II: Simulation tutorial, HVAC&R Research, 18:5, 956-973

To link to this article: https://doi.org/10.1080/10789669.2011.582917

Published online: 27 Sep 2012.

Submit your article to this journal

Article views: 1192

View related articles

Citing articles: 1 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=uhvc21
Review Article
Dynamic modeling for vapor compression
systems—Part II: Simulation tutorial
Bryan P. Rasmussen∗ and Bhaskar Shenoy
Department of Mechanical Engineering, Texas A&M University, College Station, TX 77843-3123, USA

Corresponding author e-mail: brasmussen@tamu.edu

This two-part article provides an introduction to dynamic modeling for vapor compression systems. Part
II presents example physics-based models for each component with a discussion on common assumptions
and model variations. For two-phase heat exchangers, examples for both moving boundary and finite-
control volume approaches are given, along with their associated advantages and limitations. Particular
modeling challenges, such as model initialization, validation, and numerical simulation, are also addressed,
and sample simulation results are utilized to compare modeling paradigms and illustrate key issues. Rather
than advocating the use a particular software tool, this article provides a general tutorial on constructing
dynamic simulations of vapor compression systems, outlining potential challenges and possible solutions.

Introduction Defining model purpose


Advancements in computation, simulation, and Modeling is an exercise in abstraction. Thus, be-
control have served as enabling technologies for cre- fore initiating any modeling effort, it is critical to
ating accurate, real-time, physics-based models of define the desired purpose of the model so as to
vapor compression system (VCS) dynamics. While guide the necessary assumptions and specify the
some of the difficulties associated with VCS simula- criteria for evaluating the modeling effort. Dynamic
tion are common to simulation efforts in general, the models serve a multitude of purposes, and although
complexities of two-phase flows offer unique chal- developing highly detailed models may seem to be
lenges and rich transient behavior. Part II of this the best choice to ensure accuracy, it may impede
two-part article seeks to provide a tutorial regarding usefulness. The principle of parsimony must be em-
dynamic simulation of VCSs using physics-based ployed during the process to ensure that an appro-
models. Issues of modeling approach, validation, priate model is created.
integration, and computation will be addressed, as Dynamic models of VCSs are valuable for
well as potential difficulties. (1) system analysis, design, and optimization;

Received September 17, 2010; accepted February 23, 2011


Bryan P. Rasmussen, PhD, Member ASHRAE, is Associate Professor. Bhaskar Shenoy is Student.

956
HVAC&R Research, 18(5):956–973, 2012. Copyright 
C 2012 ASHRAE.

ISSN: 1078-9669 print / 1938-5587 online


DOI: 10.1080/10789669.2011.582917
HVAC&R RESEARCH 957

(2) model-based control design; (3) control tuning termine the evolution of system pressures with in-
and commissioning; and (4) fault detection and di- let/outlet mass flow rates as model inputs. Mass flow
agnosis. Each of these tasks may benefit from a devices, such as compressors and valves, are also
different modeling approach with varying levels of discussed briefly, as these models are significantly
detail. For example, while a linearized model of low simpler and utilize inlet and outlet pressures to de-
dynamic order is preferred for model-based multi- termine mass flow rates. Outlet fluid enthalpy is de-
variable control design, a detailed nonlinear simu- termined by each component model and supplied as
lation model may be preferred for tuning a single an input to subsequent component models. Receiver
proportional-integral-derivative (PID) controller or models are addressed separately. After individual
optimizing overall system design. Readers are en- component models are reviewed, fundamental theo-
couraged to keep this perspective as they evaluate retical and practical challenges associated with VCS
and select an appropriate modeling/simulation ap- model simulation and prediction is discussed.
proach.
Compressors
Overview of dynamic modeling
approaches The bulk of residential, commercial, and indus-
trial VCSs are driven by positive displacement com-
As discussed in part I of this article, dynamic pressors. Except for the thermal capacitance of the
models are often classified as physics-based and compressor shell, the dynamics of these compres-
data-based models. While data-based models are an sors typically evolve on much faster time scales
effective and fast method of generating dynamics than the heat exchanger dynamics, and they are thus
models, they are valid for only the system and sce- modeled with static relationships that dictate the re-
nario from which the data was obtained. Thus sim- frigerant mass flow rate and the outlet fluid enthalpy.
ulations of VCSs generally employ models based This approach is typical of the majority of publica-
on fundamental physics to ensure applicability to tions regarding VCS dynamics. The mass flow rate
a wide range of conditions and modularity in sys- (Equation 1) is characterized by the speed, ωk , the
tem design and configuration. For this reason, the volume, Vk , the inlet density ρk = ρ(Pk,in , h k,in )
following discussion is restricted to physics-based and a volumetric efficiency, typically modeled using
models; for an overview of data-based modeling a semi-empirical relationship based on the pressure
techniques and appropriate references, the reader is ratio and speed ηvol = f (Pk,out /Pk,in , ωk ). The state
referred to part I of this article. of the outlet fluid is determined assuming an adia-
batic efficiency (Equation 2), where the isentropic
Notation enthalpy is determined as
The Nomenclature section located at the end of h out,isen = h(Pk,out , sk,in ) and
this article defines the notation of variables used
throughout the article. To the extent possible, stan- sk,in = s(Pk,in , h k,in ),
dard notation is used. However, some exceptions are
ṁ k = ωk Vk ρk ηvol , (1)
necessary, as this article spans more than one field
of study, which results in some inevitable conflicts h k,out,isen − h k,in
= ηk . (2)
in standard notation. h k,out − h k,in

As needed, a lumped capacitance model of


Physics-based dynamic models the compressor shell thermal dynamics may be
included.
In this section, physics-based modeling ap-
proaches for each component of a typical VCS are Expansion valves
presented and discussed. A VCS model is con-
structed by appropriately integrating models for Expansion valves are a primary means of
each of the individual component models that com- metering refrigerant flow in a VCS. Orifice and
pose the cycle: condenser, expansion valve, evapo- capillary valves have a fixed area and are designed
rator, and compressor (Figure 1). The primary focus to meter the proper flow of refrigerant at a particular
is on two-phase heat exchanger dynamics that de- condition. More complex valves utilize mechanical
958 VOLUME 18, NUMBERS 5, OCTOBER 2012

Figure 1. System modeling: external inputs and interconnection variables (color figure available online).

or electronic means to regulate refrigerant flow Characterization of the orifice area depends on the
based on pressure or temperature. Thermostatic type of valve. For fixed orifice valves, the area is
expansion valves (TEVs) utilize a sensing bulb constant, Av = a0 ; whereas for electronic expan-
filled with a two-phase refrigerant placed at the sion valves, the orifice area is directly controlled,
evaporator outlet to measure superheat temperature. Av = f (u v ). In TEVs, the orifice area depends on
The bulb pressure acts on a diaphragm inside the internal geometry and is adjusted by fluctuating
valve to increase/decrease the flow area accordingly. bulb and evaporator pressures, Av = f (Pbulb , Pe ).
Expansion valves that alter the flow area to regulate Pressure in the sensing bulb is typically modeled
pressure, rather than superheat, are termed pressure by performing an energy balance (Equation 4),
regulation valves, or automatic expansion valves leading to a simple first-order dynamic model
(AEVs). These work on the same principle as the (Equation 5) with a time constant τ . By including
TEV, but instead of a sensing bulb, they use the separate energy balance equations for the bulb,
evaporator pressure directly. Electronic expansion pipe wall, refrigerant, etc., it is possible to develop
valves modify the flow area using an externally higher-order models of TEV behavior:
controlled mechanism, typically a stepper motor. 
In virtually all published models, these valves ṁ v = Av Cd ρv (Pv,in − Pv,out ), (3)
are assumed to be ideal throttling valves, and thus
isenthalpic, h v,in = h v,out . For fixed orifice expan- m bulb C p,bulb Ṫbulb = αo Ao (Tamb − Tbulb )
sion valves, the refrigerant flow rate is typically
modeled using a form of Bernoulli’s equation − αi Ai (Tbulb Ter o ), (4)
(Equation 3), where ρv = ρ(Pv,in , h v,in ). The
discharge coefficient is determined empirically or Tb (s) ko
from manufacturer information, Cd = f (Av , P). = . (5)
Ter o (s) τs + 1
HVAC&R RESEARCH 959

Receivers/accumulators ary” (MB) modes, where parameters are lumped in


regions defined by fluid phase, but the fluid phase
Receivers and accumulators are used in VCSs transition point is a dynamic variable, or “finite-
to store excess refrigerant and ensure that the fluid control volume” (FCV) models that discretize the
leaving the pressure vessel is in the appropriate fluid heat exchanger into many zones or cells and apply
phase (e.g., refrigerant leaving low-side accumula- conservation equations. Example models for each
tor and entering compressor is a vapor). Modeling of paradigm are presented here with simulation results.
these components is straightforward, applying con- For convenience, all heat exchanger models can
servation of mass and energy to the refrigerant and be placed in the nonlinear descriptor form given in
conservation of energy to the receiver shell (Estrada- Equation 9. In this form, x represents the dynamic
Flores et al. 2003). states with the associated time derivatives ẋ, and
the external inputs are represented by u. The vec-
ṁ r ec = ṁ r ec,in − ṁ r ec,out , (6) tor f (x, u) represents the mass and energy flows
d(m r ec u r ec ) that drive the dynamic response (e.g., Equation 10).
= ṁ r ec,in h r ec,in − ṁ r ec,out h r ec,out At equilibrium, ẋ = 0, and the resulting equation
dt f (x, u) = 0 represents the steady-state conserva-
− αi Ai (Tr ec − Tr ec,w ), (7) tion equations for refrigerant energy, refrigerant
(C p m)r ec,w Ṫr ec,w = αi Ai (Tr ec − Tr ec,w ) mass, and energy stored in the heat exchanger metal.
The matrix Z (x) is a transformation from variables
− αo Ao (Tr ec,w − Tamb ). (8) of intuitive importance (e.g., refrigerant mass stor-
age rate) to variables that facilitate simulation and
For simulation, these equations are expanded, recog- are easily measured (e.g. refrigerant pressure, tem-
nizing that the receiver volume is fixed, V f + Vg = perature):
Vr ec , by using relationships that define the refrig-
erant mass and energy in terms of their saturated Z (x)ẋ = f (x, u), (9)
liquid and vapor phases, m r ec = ρ f V f + ρg Vg and
f (x, u)
m r ec u r ec = ρ f V f u f + ρg Vg u g . In this way, all the ⎡ ⎤
time derivatives of saturated quantities can be de- ṁ in h in − ṁ out h out − αi Ai (Tr − Tw )
fined in terms of a single dynamic variable: the re- =⎣ ṁ in − ṁ out ⎦.
du
ceiver pressure, i.e., u̇ f = ( d Pf ) Ṗr ec . Depending on αi Ai (Tr − Tw ) − αo Ao (Tw − Tair )
the placement of the receiver on the high/low pres- (10)
sure side of the cycle, the exit enthalpy is known to
be either saturated liquid or vapor, respectively.
Computational difficulties may arise when com- MB models
bining pressure vessels with heat exchanger models, The MB approach seeks to capture the salient
as the intermediate mass flow rate is dictated by the dynamics of multiple fluid phase heat exchangers
small pressure differential between the two com- while minimzing the number of differential equa-
ponents. This tight coupling leads to stiff differen- tions required for simulation. This approach as-
tial equations (see subsequent section). Thus, some sumes a time-varying boundary between regions
researchers propose integrating the heat exchanger of different fluid states (i.e., subcooled liquid, two-
and receiver models, assuming isobaric conditions phase, or superheated vapor). Separate control vol-
and neglecting the intermediate pressure drop (El- umes are considered for each of the fluid regions,
dredge et al. 2008). and the necessary distributed parameters are lumped
for each of these regions (Figure 2). To derive gov-
Heat exchangers erning differential equations suitable for simula-
tion or analysis, the most common method is to
As discussed in part I of this article, the thermal begin with the governing partial differential equa-
dynamics of a VCS are typically slower than the tions (PDEs) for fluid flow in a tube (Grald and
mechanical dynamics, and the bulk of the model MacArthur 1992). Assumptions of isobaric phase
complexity lies in the two-phase flow heat exchang- change, negligible axial refrigerant conduction, and
ers. The majority of recent modeling and simulation unidirectional flow are applied, and the PDEs are
efforts are segregated into either “moving bound- integrated along the length of the heat exchanger
960 VOLUME 18, NUMBERS 5, OCTOBER 2012

Figure 2. MB evaporator model.

  
to remove the spatial dependence, resulting in Tw,int − Tw2
several ordinary differential equations (ODEs) (He (mC p )w L̇ 1 + λ1 Ṫw2
LT
et al. 1997; Rasmussen and Alleyne 2004).
As an example, for an evaporator, Equations 11 = qo2 − qi2 . (16)
and 12 are the conservation of refrigerant mass
for each region, Equations 13 and 14 are the con-
servation of refrigerant energy, and Equations 15 The final form of governing equations differs
and 16 are the conservation of heat exchanger wall slightly in the literature, depending on assumptions
energy. In these equations, q is the heat transfer regarding the lumped fluid properties, void frac-
to/from the wall, such as qi1 = αi1 Ai λ1 (Tw1 − Tr 1 ) tion, and wall temperature at the interface between
and qo2 = αo Ao λ2 (Tair − Tw2 ), and λ is the the regions, Tw,int . For example, the properties
normalized length of each fluid region, i.e., for the two-phase region typically employ a mean
λ2 = L 2 /L T : void fraction ρ1 h 1 = ρ f h f (1 − γ̄ ) + ρg h g (γ̄ ), the
properties in the single-phase region are generally
dρ1 lumped using an average value h 2 = 12 (h g + h o ),
Acs L 1 Ṗe + (ρ1 − ρg )Acs L̇ 1 = ṁ i − ṁ int , and the interface wall temperature is assumed to
d Pe
(11) simply be equal to the two-phase wall tempera-
ture Tw,int = Tw1 . These six equations can be al-
dρ2 dρ2 gebraically combined to eliminate the mass flow
Acs L 2 Ṗe + Acs L 2 ḣ o + (ρg − ρ2 )Acs L̇ 1
d Pe dh o rate at the fluid interface and then organized into
= ṁ int − ṁ o , (12) the Z (x)ẋ = f (x, u) form, as in Equation 17. The
  matrix Z (x) remains full rank and invertible as long
dρ1 h 1 as λ1,2 > 0, or in other words, as long as evaporator
− 1 Acs L 1 Ṗe + (ρ1 h 1 − ρg h g )Acs L̇ 1
d Pe superheat is maintained:
= ṁ i h i − ṁ int h g + qi1 , (13) ⎡ ⎤ ⎡ L̇ ⎤
  z 11 z 12 0 0 0 1
dρ2 h 2 dρ2 h 2 ⎢ z 21 z 22 z 23 0 0 ⎥⎢ Ṗ ⎥
− 1 Acs L 2 Ṗe + ⎥⎢ ⎥
e
Acs L 2 ḣ o ⎢
d Pe dh o ⎢ z 31 z 32 z 33 0 ⎢
0 ⎥ ⎢ ḣ o ⎥ ⎥
⎣ 0 0 0 z 44 0 ⎦ ⎣ Ṫw1 ⎦
+(ρg h g − ρ2 h 2 )Acs L̇ 1
z 51 0 0 0 z 55 Ṫw2
= ṁ int h g − ṁ o h o + qi2 , (14) ⎡ ⎤
   ṁ i (h i − h g ) + qi1
Tw1 − Tw,int ⎢ ṁ o (h g − h o ) + qi2 ⎥
(mC p )w L̇ 1 + λ1 Ṫw1 ⎢ ⎥
LT =⎢ ṁ i − ṁ o ⎥. (17)
⎣ q −q ⎦
= qo1 − qi1 , (15) o1 i1
qo2 − qi2
HVAC&R RESEARCH 961

While this is a distinctly nonlinear model, con- lumped parameter approaches that rely on effective
structing a linear approximation for the purposes of parameters and only describe the dominant physical
dynamic analysis and control design is straightfor- mechanisms for heat transfer, fluid flow, etc. How-
ward. First the values of the state variables are de- ever, the resulting model contains many differential
termined at the desired steady-state operating point, equations, which increases computation time, with
xo = [L 1 (0) Pe (0) h o (0) Tw1 (0) Tw2 (0)], as the possibility of additional numerical challenges
are the values of the external inputs, u o = (e.g., stiff dynamics, differential algebraic equations
[ṁ i (0) ṁ o (0) h i (0) αo (0) Tair (0)]. The lin- [DAEs], iterative solvers, etc.). If a pressure gradi-
earized model is then formulated in terms of the ent is assumed, then a staggered grid approach for
deviation variables δx = x − x0 , and δu = u − u 0 modeling the mass and energy flow for each control
as: region is typical (Eborn 2001), where pressure dif-

ferentials dictate mass flow rates, and mass flow rate
−1 ∂ f (x 0 , u 0 ) differentials drive the pressure and energy dynamics
δ ẋ = Z (x0 ) δx
∂x (Figure 4).

The governing equations for each region include
−1 ∂ f (x 0 , u 0 )
+ Z (x0 ) δu. (18) (1) the conservation of refrigerant mass (Equation
∂u 19), where the intermediate mass flow rates are
Additional details regarding this linearization pro- determined using a manipulated form of the
cess, and a complete description of the resulting Darcy-Weisbach equation (Equation 20), where the
model, can be found in several places in the litera- coefficient Cd is a function of the tube diameter,
ture (He et al. 1998; Rasmussen and Alleyne 2004). length, friction factor, etc.; (2) the conservation
The linearized model is a local approximation, and of refrigerant energy (Equation 21), where the
the approximation errors grow as the model oper- storage rate of specific energy u̇ e, j is expanded
ates away from the operating point of linearization. in terms of pressure and and enthalpy; and (3)
Although nonlinearities in the heat exchanger pa- the conservation of heat exchanger wall energy
rameters (e.g., heat transfer correlations) and fluid (Equation 22), where qi, j = αi, j Ai, j (Tw, j − Tr, j )
properties do result in some error, the dynamics re- and qo, j = αo, j Ao, j (Tair, j − Tw, j ).
main qualitative the same (Figure 3).
ṁ e, j = ṁ j−1 − ṁ j , (19)
Fixed-control volumes models 
Finite-difference or FCV approaches offer the ṁ j = Cd A ρ(P j − P j+1 ), (20)
capability to model the fluid behavior with ex-  
du e, j du e, j
treme detail, capturing the thermo-physical gradi- U̇j = ṁ e, j u e, j + m e, j Ṗ j + ḣ j
d Pe dh j
ents and distributed parameters. This, reportedly,
lends itself to higher accuracy prediction than the = ṁ j−1 h j−1 − ṁ j h j + qi, j , (21)

Figure 3. Comparison of nonlinear and linearized MB models for step changes in expansion valve and compressor speed. Signals
shown are evaporator pressure, evaporator superheat, and evaporator air outlet temperature (color figure available online).
962 VOLUME 18, NUMBERS 5, OCTOBER 2012

Figure 4. FCV evaporator model.


Ė w, j = mC p w Ṫw, j = qo, j − qi, j . (22) If isobaric conditions are assumed for the heat
exchanger, P j = Pe , ∀ j, and the conservation of
momentum is neglected, the conservation equations
When organized into the standard Z (x)ẋ = f (x, u)
may be simplified to eliminate the intermediate mass
form, j
flow variables as ṁ j = ṁ in − k=1 ṁ e,k . The con-
⎡ ⎤ ⎡ ⎤ servation of refrigerant energy (Equation 21) now
U̇1 Ṗ1 becomes
⎢ .. ⎥ ⎢ .. ⎥
⎢ . ⎥ ⎢ . ⎥  
⎢ ⎥ ⎢ ⎥ du e, j du e, j
⎢ U̇n ⎥ ⎢ Ṗn ⎥ U̇j = ṁ e, j u e, j + m e, j Ṗe + ḣ j
⎢ ⎥ ⎡ ⎤ ⎢ ⎥ d Pe dh j
⎢ ṁ e,1 ⎥ ⎢ ḣ 1 ⎥
⎢ ⎥ Z 11 Z 12 0 ⎢ ⎥  
⎢ .. ⎥ ⎢ .. ⎥
⎢ . ⎥ = ⎣ Z 21 Z 22 0 ⎦ ⎢ . ⎥ 
j−1

⎢ ⎥ ⎢ ⎥ = ṁ in − ṁ e,k h j−1 − h j
⎢ ṁ e,n ⎥ 0 0 Z 33 ⎢ ḣ n ⎥
⎢ ⎥    ⎢ ⎥ k=1
⎢ Ė w,1 ⎥ ⎢ Ṫw,1 ⎥
⎢ ⎥ pseudo−state ⎢ ⎥
⎢ .. ⎥ transformation matrix ⎢ . ⎥ + ṁ e, j h j + qi, j , (24)
⎣ . ⎦ ⎣ .. ⎦
Ė w,n Ṫw,n where the rate of mass storage in each re-
     
mass and energy dynamic state gion is also expanded as ṁ e, j = ρ̇e, j V j =
storage rates dρ dρ
variables
( d Pe,ej Ṗe + dhe,jj ḣ j )V j . After these substitutions, the
⎡ ⎤
ṁ i h i − ṁ 1 h 1 + qi,1 conservation equations in Equation 25 are trans-
⎢ .. ⎥ formed into Equation 26:
⎢ . ⎥
⎢ ⎥
⎢ ṁ n−1 h n−1 − ṁ o h o + qi,n ⎥ ⎡ ⎤ ⎡ ⎤
⎢ ⎥ U̇1 ṁ i h i − ṁ 1 h 1 + qi,1
⎢ ṁ i − ṁ 1 ⎥
⎢ ⎥ ⎢ .. ⎥ ⎢ .. ⎥

=⎢ .
.. ⎥. (23) ⎢ . ⎥ ⎢ . ⎥
⎥ ⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ U̇n ⎥ ⎢ ṁ h − ṁ h + q ⎥
⎢ ṁ n−1 − ṁ o ⎥ ⎢ ⎥ ⎢ n−1 n−1 o o i,n

⎢ ⎥ ⎢ ṁ e ⎥ = ⎢ ṁ − ṁ ⎥,(25)
⎢ qo,1 − qi,1 ⎥ ⎢ ⎥ ⎢
in out

⎢ ⎥ ⎢ Ė w,1 ⎥ ⎢ q − q ⎥
⎣ .. ⎦ ⎢ ⎥ ⎢
o,1 i,1

. ⎢ .. ⎥ ⎣ .. ⎦
qo,n − qi,n ⎣ . ⎦ .
   Ė w,n qo,n − qi,n
steady state mass and      
energy balances mass and energy steady state mass and
storage rates energy balances
HVAC&R RESEARCH 963
⎡ ⎤
Ṗe air temperature predictions converge for a modest
⎢ ⎥ ḣ 1 level of discretization, the prediction of refriger-
⎢ ⎥
⎡ ⎤ ⎢ ⎥ .. ant outlet temperature (superheat) clearly requires
Z 11 Z 12 0 ⎢ ⎥ . n > 20 before the simulations show quantative
⎢ ⎥
⎣ Z 21 Z 22 0 ⎦ ⎢ ⎥
⎢ ḣ n ⎥ convergence.
0 0 Z 33 ⎢ Ṫw,1 ⎥
⎢ While n > 20 may be required for numerical
   ⎢ . ⎥ ⎥
pseudo−state ⎣ .. ⎦ consistency, the dominant dynamics appear to be
transformation matrix of much lower order. This indicates some level of
Ṫw,n “over-modeling,” as data-driven models do not re-
  
dynamic state quire 50+ dynamic states to reproduce the required
variables transient response (Rasmussen et al. 2005). In this
⎡ ⎤
ṁ in (h in − h 1 ) + qi,1 form, these models capture subtle parameter varia-
⎢ .. ⎥ tions but are of high dynamic order, making them
⎢ . ⎥ most useful for control evaluation or tuning, rather
⎢ ⎥
⎢ ṁ in (h n−1 − h out ) + qi,n ⎥ than dynamic analysis or model-based control de-
⎢ ⎥
=⎢ ṁ in − ṁ out ⎥. (26) sign. If such design tasks are attempted, numerical
⎢ qo,1 − qi,1 ⎥
⎢ ⎥ model reduction techniques are generally applied
⎢ . ⎥
⎣ .. ⎦ to obtain a simple low-order approximate model.
qo,n − qi,n Although computation time was reported to be a
   serious concern in some earlier publications (e.g.,
modified steady state mass and Wang and Touber 1991), computer and software ad-
energy balances
vances are rapidly eliminating this as a significant
issue. Current capabilities allow real-time (or bet-
This particular model has 2n + 1 dynamic states, ter) simulation of VCS models with multiple FCV
where n is the number of control volumes or cells heat exchangers.
(conservation of refrigerant/heat exchanger energy When the FCV and MB models are initialized
for each region, and one equation for the conser- with the same physical parameters, the resulting
vation of refrigerant mass). If conservation of mo- simulations exhibit only small differences (Fig-
mentum is included, the model includes 3n dynamic ure 6). The MB model responds slightly faster than
equations, with the additional algebraic constraints the FCV model to step changes in expansion valve
that define the mass flow rate between regions. or compressor, indicating that some adjustment of
Publications typically report using n > 20 regions the “effective” parameters used by the MB model
to achieve grid (or mesh) independence. Figure 5 may be appropriate (e.g., the user may opt to use
presents simulation results for FCV models with a weighted average of a particular parameter ver-
increasing numbers of regions. While pressure and sus strict average when lumping the parameters for

Figure 5. Comparison of FCV with differing levels of discretization for step changes in expansion valve and compressor speed.
Signals shown are evaporator pressure, evaporator superheat, and evaporator air outlet temperature (color figure available online).
964 VOLUME 18, NUMBERS 5, OCTOBER 2012

Figure 6. Comparison of FCV and MB models for step changes in expansion valve and compressor speed. Signals shown are
evaporator pressure, evaporator superheat, and evaporator air outlet temperature (color figure available online).

two-phase or superheated fluid regions). The models solvers, thus requiring the user to resolve the issue
also give slightly different steady-state predictions of numerical stiffness prior to simulation.
for evaporator outlet conditions (superheat). This The primary source of multi-time scale behavior
discrepancy demonstrates the fundamental differ- is the inclusion of the conservation of momentum
ence between the modeling paradigms, as the FCV in the dynamic equations. Given that the pressure in
model avoids extensive lumping of parameters and the system propagates much faster than the evolu-
captures parameter and temperature gradients that tion of the thermal dynamics, this results in a natu-
result in a more precise prediction of outlet condi- ral conflict. A common approach of addressing this
tions. However, it is noted that these prediction er- issue is to assume isobaric conditions in the heat ex-
rors are typically within the differences seen when changers (thus neglecting the associated momentum
validating the models against experimental data and, change). Zhang et al. (2009) presented a direct com-
therefore, may not be a critical factor in selecting parison of the most common assumptions regarding
which model to use. how the momentum equation is handled, revealing
that including momentum effects is not critical for
Modeling challenges and innovations large transient simulations but that discernible dif-
Stiff dynamics and DAEs ferences are present for instaneous step changes in
VCSs exhibit dynamics that evolve on several mass flow rate. However, as most compressors and
time scales. At the extremes are the dynamics of valves do not respond instantly, this may not be a
the building or cabin cooling load whose dynamic critical effect to capture.
response is on the order of hours or days, whereas Beyond momentum dynamics, most two-phase
the propagation of pressure waves through the re- heat exchangers exhibit multi-time scale behavior.
frigerant is completed in milliseconds. The user An analysis of the eigenvalues of the linearized heat
typically isolates a time scale of primary interest exchanger dynamics reveals dynamic modes whose
and approximates slower dynamics as constant, and time scale varies by several orders of magnitude
faster dynamics as instantaneous, resulting in a set (Rasmussen and Alleyne 2004). This leads naturally
of DAEs. However, isolating dynamics at particular to the search for a reduced-order model that elimi-
time scales is not always possible, and the resulting nates dynamics outside of the time scale of interest,
system retains the multi-time scale behavior and is serving both to simplify the model computation-
referred to as a “stiff ” set of ODEs. Most simulation ally for simulation tasks and analytically for control
software packages include variable time-step nu- design tasks. The particular method of model re-
merical integration algorithms for stiff systems and duction and the physical interpretation varies in the
algorithms for solving simple DAEs (i.e., DAEs of literature. He et al. (1997) used both pure numerical
index 1). However, compiling simulations for real- reduction as well as a physically motivated tech-
time or embedded computing requires fixed step nique for removing non-essential modes (He et al.
HVAC&R RESEARCH 965

1998), an approach later duplicated by Leducq et al. tion options were used). The same simulations were
(2003). Rasmussen and Alleyne (2004) explored al- also compiled into stand-alone executable files re-
ternative choices of state variables to identify those sulting in order-of-magnitude improvements in ex-
that virtually decoupled the modes according to time ecution speed. The real-time factor was computed
scale and then applied singular pertubation tech- by dividing the total simulation time (initialization
niques for model reduction. Regardless of the ap- + execution) by the length of the simulated time
proach, these studies indicate that some commonly (500 s).
modeled dynamics are not essential for accurate
transient simulation. Fluid properties and heat transfer correlations
Modelers of centrifugal compressors should Refrigerant properties pose a particular chal-
also be aware of the potential challenges spe- lenge for dynamic VCS simulation, as standard
cific to the “surge-and-stall” dynamics inherent methods using equations of state (EOS) for de-
to this particular component. Models of this phe- termining fluid properties can be computationally
nomena are readily available in the literature (see expensive and not suitable for real-time dynamic
Moore and Greitzer, 1986) and numerous subse- simulations that require updating of fluid proper-
quent publications). These unstable dynamics pose a ties at each simulation time step. Likewise, using
particular challenge for simulation but also typi- software interfaces to external databases or com-
cally contribute to the “stiff systems” problem, as panion software programs can significantly increase
these dynamics are much faster than heat exchanger simulation computation time. A typical solution
dynamics. Integrating standard VCS models with to this dilemma is to utilize high-accuracy fluid
centrifugal compressors is likely to be an area of property database software, e.g., NIST’s REFPROP
research in the near future, specifically for aircraft (Lemmon et al. 2007), to generate tables of fluid
cooling systems. properties for a predetermined range of conditions.
This table is then used for fast interpolation of
Real-time simulation fluid properties during simulation. Likewise, tables
Computation speed for VCS simulations is for the partial derivatives of fluid properties may
largely dependent on the chosen heat exchanger be generated by computing the numerical gradient
modeling paradigm. As described above, solving or from fundamental thermodynamic relationships
the nonlinear differential equation Z (x)ẋ = f (x, u) (e.g., Appendix B of Tummescheit 2002). Care must
requires a matrix inversion, where the matrix size be taken to appropriately handle discontinuities in
depends on the number of dynamic states. MB the property tables, such as the transition point be-
models typically have five to nine states, depending tween fluid phases, or significant numerical inac-
on the formulation, and FCV models have 2n + 1 curacies will occur during interpolation. While this
dynamic states, where n is the number of control method of fluid property calculation sacrifices some
volumes. Linearized model approximations are level of accuracy, increasing the grid density of the
capable of much faster simulation, as the matrix interpolation tables can reduce the numerical errors
inversion is only computed once during the intiliza- to insignificant levels.
tion of the simulation to compute the appropriate Correlations for heat transfer, void fraction, and
state space matrices (Equation 18). Compiled code fluid flow are also the source of potential difficul-
also offers a significant improvement in execution ties. For maximum accuracy, the values for heat
time, although this eliminates the option of modify- transfer coefficients, pressure drop, etc. must be up-
ing the simulation or viewing the predicted outputs dated during the simulation. Some approximations
during execution. may be necessary to enable fast explicit calculation
For comparison, Table 1 summarizes the compu- of values and to avoid circular algebraic dependen-
tation time required to produce the 500-second sim- cies that would require calling a numerical equation
ulations shown in Figures 3 and 5. The simulations solver at each time step. For example, a heat trans-
were conducted on a standard desktop computer (3- fer coefficient correlation may require heat flux as a
GHz processor) using a fourth-order Runge-Kutta parameter. However, the instantaneous heat flux is
solver with a fixed step size of 0.01 s. The “standard determined using the heat transfer coefficient. With
simulation” results were created by simulating the only minor loss in accuracy, the heat flux from the
systems using MATLab/Simulink with the default previous time step may be used as an approxima-
simulation settings (i.e., no “accelerated” simula- tion of the current heat flux for calculating the heat
Table 1. Comparison of simulation execution time.

Standard simulation Compiled simulation


Model type Dynamic states Initialization (s) Execution (s) Real-time factor Compilation (s) Execution (s) Real-time factor
Linearized MB 5 4.1 5.4 0.02 16.8 < 0.1 0.03
Nonlinear MB 5 10.2 21.2 0.06 28.7 0.6 0.06
Nonlinear FCV—1 Region 3 8.3 22.8 0.06 30.5 1.7 0.06

966
Nonlinear FCV—10 Region 21 12.8 158.7 0.34 31.9 11.1 0.09
Nonlinear FCV—20 Region 41 13.9 774.5 1.58 32.6 23.8 0.11
Nonlinear FCV—30 Region 61 14.4 2166.1 4.36 33.8 37.1 0.14
Nonlinear FCV—40 Region 81 13.2 4772.5 9.57 41.5 53.2 0.19
Nonlinear FCV—50 Region 101 14.2 9153.4 18.34 43.0 72.4 0.23
Nonlinear FCV—100 Region 201 19.0 70,594 141.23 47.5 275.2 0.65
HVAC&R RESEARCH 967

transfer coefficient and avoiding an algebraic depen- (Ljung 1999). Parametric-based approaches explore
dency. Although this is an obvious approximation, the sensitivity of the model predictions to parameter
numerical inaccuracies are typically well within the variations and techniques for improving model fit by
uncertainty levels of the correlation. More impor- parameter tuning (Butterfield and Thomas 1986a,
tant is to ensure continuity in the computation to 1986; Butterfield 1990). Robust model validation
avoid false transient responses due to discontinu- is specifically designed for control-oriented models
ous changes in model parameters during a simula- and seeks to determine the level of dynamic model
tion. For example, numerical problems in solving uncertainty required to reproduce the experimental
for heat transfer correlation may lead to a sudden data (Poolla et al. 1994; Dullerud and Smith 1996;
change in the value of the coefficient. This, in turn, Smith and Dullerud 1996).
would induce a transient response not unlike a sud- As seen from the dicussion in previous sections,
den blockage or fouling of the heat exchanger. Thus, despite the large number of publications regarding
care must be taken to ensure continuous calculation VCS dynamic models, relatively few provide de-
of the correlation from one simulation time step to tailed experimental validation. The most common
the next. deficiencies are (1) no validation, (2) only steady-
state validation of a dynamic model, (3) compari-
Model validation and tuning son with other simulation models, (4) comparison of
Central to any model-based research effort is only a few model outputs or not the critical outputs,
establishing the validity of the model. The prob- or (5) qualitative comparison with no quantitative
lem of model validation is separate and distinct metrics. Depending on the intended model use, val-
from the issues associated with model purposive- idation of the model for small transient deviations
ness (e.g., whether the model meets the required pur- may or may not be appropriate; this includes fre-
pose), model plausibility (e.g., whether the model quency domain validation, which focuses on linear
conceptually agree with the process physics), and model behavior.
model verification (e.g., whether the software em- Perhaps the lack of detailed model validation in
bodiment of the model is accurate) (Bohlin 1991). the literature stems from the challenge of proper ini-
Model validation is a question of assessing whether tializing and tuning of the model. Most VCS mod-
the model outputs agree with experimental data els contain measurable physical parameters (areas,
(Murray-Smith 1998). From a philosophical per- volumes) as well as unmeasurable or uncertain pa-
spective, it is impossible to truly validate a model, rameters (void fraction, heat transfer correlations).
as this would require an infinite amount of experi- Additionally, many model parameters are lumped
mental data (Smith et al. 1997); model invalidation (effective surface areas, effective volumes, etc.) and
is a more proper term, as the majority of techniques must be adjusted given model assumptions. For ex-
are used to determine when a model is not capa- ample, whereas the effective internal surface area of
ble of reproducing the measurements (i.e., model a heat exchanger includes only the area that directly
falseness). participates in heat transfer, the effective heat ex-
The techniques for external model validation can changer volume should be adjusted to include head-
be considered qualitative (e.g., visual inspection of ers and entrance exit pipe lengths because all of
model/data agreement) and quantitative (e.g., nu- these participate in the increase/decrease in system
merical metric of model validity). Common quanti- pressure. The following figures illustrate some of
tative methods for dynamic models include (1) sta- the key tuning parameters and their relative effects
tistical, (2) residual-based, (3) parametric, and (4) on predicted temperatures and pressures.
robustness methods. Statistical methods can be ei- Figure 7 presents MB model simulations where
ther descriptive statistics that evaluate the mean, the external heat transfer coefficient has been in-
variances, and correlations of the data or infer- creased. While the pressure and superheat response
ential statistics that deal with hypothesis testing remain relatively unchanged, the outlet air temper-
(Ljung 1999). Residual-based approaches rely on ature is lowered by several degrees and exhibits a
numerical performance measures by calculating the transient response that is more responsive to valve
root mean square of the prediction error, relative and compressor changes. Figure 8 shows the effect
prediction error, or other variations (Murray-Smith of increasing the effective cross-sectional area of
1998) and generally assume that model residuals the heat exchanger or, in other words, increasing
are caused by noise that is uncorrelated to the input the total internal volume. Assuming a constant void
968 VOLUME 18, NUMBERS 5, OCTOBER 2012

Figure 7. Comparison of MB simulations for variations in external heat transfer coefficient. Transients are induced by step changes
in expansion valve and compressor speed (color figure available online).

fraction, this also equates to more refrigerant mass A final comment on model tuning: VCS models
in the heat exchanger. As expected given the large exhibit a significant sensitivity to refrigerant mass
volume to pressurize and the greater mass to af- flow rate prediction. Because the heat exchanger
fect, the transient responses are somewhat slower, model dynamics are driven by mass flow rate dif-
and the superheat response is noticeably reduced in ferentials (inlet versus outlet), accurate valve and
magnitude. Figure 9 presents an example of a pa- compressor models are critical. This is somewhat
rameter that affects primarily the transient response of a paradox, as the bulk of the dynamic complex-
rather than the steady-state values. For void frac- ities lie with the heat exchanger models, but it is
tion correlations, the slip ratio defines the velocity the quantitative accuracy of the mass flow models
of the vapor phase to liquid phase (homogeneous that largely determines the quality of the valida-
flow refers to a slip ratio of one). Increasing the slip tion and comprise much of the parameter tuning
ratio decreases the mean void fraction and increases effort. Given the large number of parameters to be
the amount of refrigerant mass. Thus, for higher tuned, it is therefore essential to cross-validate the
slip ratios, the transient response of all signals is models, using different datasets to tune and validate
slower. the model.

Figure 8. Comparison of MB simulations for variations in heat exchanger cross-sectional area (i.e., heat exchanger internal volume).
Transients are induced by step changes in expansion valve and compressor speed (color figure available online).
HVAC&R RESEARCH 969

Figure 9. Comparison of MB simulations for variations in void fraction slip ratio. Transients are induced by step changes in expansion
valve and compressor speed (color figure available online).

Simulation of hard transients efforts examine cycling from a purely experimen-


VCS startup and shutdown scenarios are among tal perspective, recent innovations in both SMB and
those termed “hard” or “large” transients in the lit- FCV models have resulted in experimentally val-
erature. The simulation challenge posed by these idated models for startup and shutdown transients
scenarios is the violation of modeling assumptions (Bendapudi et al. 2008; Hermes and Melo 2008;
and the resulting numerical singularities. For exam- Li and Alleyne 2010). Typically, these studies fo-
ple, for the MB modeling paradigm, the loss of the cus on matching the gross pressure or cooling ca-
evaporator superheat region results in a singularity pacity trends, rather than more specific transients
of the Z (x) matrix, which prevents inversion and responses like superheat. They also tend to focus
the calculation of time derivatives. Before this hap- on short-term shutdown, as longer off periods tend
pens, however, the model outputs become infinitely to accentuate the effect of unmodeled phenomena,
large, as the equations have terms in the denomi- such as refrigerant pooling and the resulting slug
nators that approach zero (Equations 16 and 17). flow.
Overcoming this challenge led to the development While the FCV models are intended to handle
of switched MB (SMB) models that handle the dis- hard transients, the MB models require specific ad-
appearance or introduction of a fluid phase region. ditions to handle the loss of superheat or subcooling.
Similar challenges are encountered when attempt- The published models differ in terms of the switch-
ing to simulate the dynamic effect of some system ing criteria and the assumptions regarding the heat
faults, such as fan blockage, valve failure, or refrig- exchanger wall temperature at the interface between
erant leak. Of secondary concern is the heightened fluid phases. Pettit et al. (1998) used a simple switch-
role of phenomenon typically ignored during mod- ing strategy based on comparing refrigerant exit en-
eling for pseudo steady-state operation (e.g., pool thalpies to saturation enthalpies to handle all cases
boiling or free convection heat transfer when fans of switching on/off superheat or subcool, and they
are off, back flow or leakage when the compressor added small numercial terms to avoid singularities
is off, and gravity-driven refrigerant pooling during when L = 0. Zhang and Zhang (2006) and McKin-
system shutdown). ley and Alleyne (2008) handled the disappearance
By assuming appropriate initial conditions, in- of superheat or subcool regions based on the length
cluding a distribution of refrigerant that meets the associated region, the former simply requiring
model assumptions, the startup scenario can be han- L S H > 0 for persistance of the superheat region,
dled using standard FCV and MB models (e.g., Ben- and the latter using L S H < L min and d L S H/dt < 0
dapudi and Braun 2002). However, cycling (startup as the conditions for switching the superheat re-
and shutdown) of a VCS requires models capable gion off. Zhang and Zhang (2006) did not spec-
of handling the hard transient effects. While earlier ify criteria for handling the cases where superheat
970 VOLUME 18, NUMBERS 5, OCTOBER 2012

Figure 10. Comparison of FCV, MB, and SMB models for the loss of evaporator superheat induced by step changes in expansion
valve. Signals shown are evaporator pressure, evaporator superheat, and evaporator air outlet temperature (color figure available
online).

reappears, but McKinley and Alleyne (2008) used The following figures illustrate the behavior of
a criteria based on void fraction. The assumptions these various models under hard transients. The
regarding interface wall temperatures also vary. Pet- simulation shown in Figure 10 assumes that the
tit et al. (1998) made the standard assumption that expansion valve is opened wide for a period of
the wall temperature at the interface is equal to time, resulting in a large increase in refrigerant flow
the two-phase wall temperature (Case 1), whereas into the evaporator. Without changing the airflow
McKinley and Alleyne (2008) used an approach rate or the compressor speed, this results in a loss
suggested in Jensen and Tummescheit (2002) that of evaporator superheat. The MB model fails, with
the interface temperature depends on whether the the prediction values exploding toward infinite due
two-phase region is growing or shrinking, Tw,int = to the singularity in the Z (x) matrix that makes
T1 if ddtL 1 < 0 and Tw,int = T2 if ddtL 1 > 0 (Case 2). it not invertable. However the FCV and SMB
Zhang and Zhang (2006) assumed a weighted av- models both handle the transient without numerical
erage Tw,int = LLtotal
2
T1 + LLtotal
1
T2 , which eliminated failures. Although examination of the internal
the numerical singularity associated with dividing model variables, such as wall temperature, differ
by the length of a disappearing region (Case 3). distinctly depending on the assumptions regarding

Figure 11. Comparison of expansion valve strategies during compressor shutdown using FCV models. Signals shown are evaporator
pressure, evaporator superheat, and evaporator air outlet temperature (color figure available online).
HVAC&R RESEARCH 971

the interface wall temperature, the prinicpal output q = heat


signals are virtually indistinguishable. s = specific entropy
Figure 11 presents simulation results for startup u = specific internal energy
and shutdown transients for a simple air-cooling α = heat transfer coefficient
evaporator with an electronic expansion valve and a γ̄ = mean void fraction
reciprocating compressor. The simulation considers η = efficiency
two scenarios: (1) the expansion valve closes simul- ρ = density
taneously with the compressor and (2) the expansion
valve is open during compressor shutoff. Note that Mathematical notation
for the former case, complete vaporization of the re-
frigerant in the evaporator is achieved. Additionally, A, B, C, D = state space matrices
note that for the latter case there is a slight disconti- Z = matrix
nuity, as the last region transitions from two phases f (·) = continuous functions
to single phase. This discontinuity is decreased with G(s) = transfer function
finer discretization of the heat exchanger. s = Laplace transform variable
d
dt
= continuous derivative
Conclusion  = differential
u = input signals
Recent developments in techniques and software y = output signals
significantly expand the range of tools for dynamic x = dynamic states
modeling of VCSs. Users should carefully select e = external noise
the appropriate modeling paradigm based on the in- a, b, c, k = empirical parameters
tended use of the model. Because of the uncertain τ = time constant
nature of some VCS model parameters, experimen- Td = time delay
tal validation is integral to every modeling effort. θ = scheduling variable
While existing models can capture salient VCS dy- ∀ = “for all”
namics in the majority of scenarios, continuing ef-
forts in the areas of model reduction, numerical Physical variables
solvers, and automated model tuning and valida-
tion are necessary to increase the capabilities of A = area
these models for analysis, design, control design, Cd = discharge coefficient
and fault detection. L = length
V = volume
Acknowledgments m = mass
t = time
The authors gratefully acknowledge the support λ = normalized length
provided by NSF grant CMMI-0644363 and the as- ω = rotational velocity
sistance of anonymous reviewers in improving the Cp = specific heat
manuscript. Any opinions, findings, and conclu-
sions or recommendations expressed in this material
Subscripts
are those of the author(s) and do not necessarily re-
flect the views of the National Science Foundation. 1, 2, . . . j, . . . n = first, second, etc. region
air = air
Nomenclature amb = ambient
bulb = bulb
Thermodynamic variables c = condenser
E = energy cs = cross-sectional
P = pressure e = evaporator
T = temperature f = liquid
U = internal energy g = vapor
h = specific enthalpy i = in
ṁ = mass flow rate int = interface
972 VOLUME 18, NUMBERS 5, OCTOBER 2012

isen = isentropic HVAC systems. ASME Journal of Dynamic Systems Mea-


k = compressor surement & Control 119(2):183–91.
o = out Hermes, C.J.L., and C. Melo. 2008. A first-principles simula-
tion model for the start-up and cycling transients of house-
r = refrigerant
hold refrigerators. International Journal of Refrigeration
rec = receiver 31(8):1341–57.
v = valve Jensen, J.M., and H. Tummescheit. 2002. Moving boundary mod-
vol = volumetric els for dynamic simulations of two-phase flows. Proceedings
w = wall of the 2nd International Modelica Conference, Oberpfaffen-
hofen, Germany.
Leducq, D., J. Guilpart, and G. Trystram. 2003. Low order dy-
namic model of a vapor compression cycle for process control
References design. Journal of Food Process Engineering 26(1):67–91.
Lemmon, E.W., M.L. Huber, and M.O. McLinden 2007. NIST
Bendapudi, S., and J.E. Braun. 2002. Development and valida- Standard Reference Database 23, reference fluid thermo-
tion of a mechanistic, dynamic model for a vapor compres- dynamic and transport properties—REFPROP, Version 8.0.
sion centrifugal chiller. Report #4036-4. ASHRAE, Atlanta, Gaithersburg: National Institute of Standards and Technol-
GA. ogy, Standard Reference Data Program.
Bendapudi, S., J.E. Braun, and E.A. Groll. 2008. A compari- Li, B., and A.G. Alleyne. 2010. A dynamic model of a vapor
son of moving-boundary and finite-volume formulations for compression cycle with shut-down and start-up operations.
transients in centrifugal chillers. International Journal of Re- International Journal of Refrigeration 33(3):538–52.
frigeration 31(8):1437–52. Ljung, L. 1999. System Identification: Theory for the User. Upper
Bohlin, T. 1991. Interactive System Identification: Prospects and Saddle River, NJ: Prentice Hall PTR.
Pitfalls. New York, Springer-Verlag. McKinley, T.L., and A.G. Alleyne. 2008. An advanced nonlinear
Butterfield, M.H. 1990. A method of quantitative validation switched heat exchanger model for vapor compression cycles
based on model distortion. Transactions of the Institute of using the moving-boundary method. International Journal of
Measurement and Control 12(4):167–73. Refrigeration 31(7):1253–64.
Butterfield, M.H., and P.J. Thomas. 1986a. Methods of quan- Moore, F.K., and E.M. Greitzer. n.d. Theory of post-stall tran-
titative validation for dynamic simulation models—Part 1: sients in axial compression systems: Part I—development of
Theory. Transactions of the Institute of Measurement and equations. 85-GT-171. American Society of Mechanical En-
Control 8(4):182–200. gineers (ASME), New York, NY.
Butterfield, M.H., and P.J. Thomas. 1986b. Methods of quan- Moore, F.K., and E.M. Greitzer. 1986. Theory of post-stall tran-
titative validation for dynamic simulation models—Part 2: sients in axial compression systems: Part I—development of
Examples. Transactions of the Institute of Measurement and equations. ASME Journal of Engineering for Gas Turbines
Control 8(4):201–19. & Power 108(1):68–76.
Dullerud, G., and R. Smith. 1996. Sampled-data model val- Murray-Smith, D.J. 1998. Methods for the external validation
idation: An algorithm and experimental application. In- of continuous system simulation models: A review. Mathe-
ternational Journal of Robust and Nonlinear Control matical & Computer Modelling of Dynamical Systems 4(1):
6(9–10):1065–78. 5.
Eborn, J. 2001. On model libraries for thermo-hydraulic appli- Pettit, N.B.O.L., M. Willatzen, and L. Ploug-Sorensen. 1998.
cations. Ph.D. Thesis, Lund, Sweden, Lund University. General dynamic simulation model for evaporators and
Eldredge, B.D., B.P. Rasmussen, and A.G. Alleyne. 2008. condensers in refrigeration. Part II: Simulation and con-
Moving-boundary heat exchanger models with variable out- trol of an evaporator. International Journal of Refrigeration
let phase. Journal of Dynamic Systems, Measurement, and 21(5):404–14.
Control 130(6):061003–12. Poolla, K., P. Khargonekar, and A. Tikku. 1994. A time-domain
Estrada-Flores, S., D.J. Cleland, A.C. Cleland, and R.W. James. approach to model validation. IEEE Transactions on Auto-
2003. Simulation of transient behaviour in refrigeration plant matic Control 39:951–9.
pressure vessels: Mathematical models and experimental val- Rasmussen, B.P., and A. Alleyne. 2004. Control-oriented mod-
idation. International Journal of Refrigeration 26(2):170–9. eling of transcritical vapor compression systems. ASME
Grald, E.W., and J.W. MacArthur. 1992. A moving-boundary Journal of Dynamic Systems, Measurement, and Control
formulation for modeling time-dependent two-phase flows. 126(1):54–64.
International Journal of Heat & Fluid Flow 13(3):266– Rasmussen, B.P., A.G. Alleyne, and A.B. Musser. 2005. Model-
72. driven system identification of transcritical vapor compres-
He, X.D., H. Asada, S. Liu, and H. Itoh. 1998. Multivariable sionsystems. IEEE Transactions on Control Systems Tech-
control of vapor compression systems. HVAC&R Research nology 13(3):444–51.
4(3):205–30. Smith, R., G. Dullerud, K. Poolla, and S. Rangan. 1997. Model
He, X.D., S. Liu, S. Liu, and H. Itoh. 1997. Modeling of va- validation for dynamically uncertain systems. Mathematical
por compression cycles for multivariable feedback control of Modelling of Systems 3(1):43–58.
HVAC&R RESEARCH 973

Smith, R., and G.E. Dullerud. 1996. Continuous-time control Zhang, W.-J., and C.-L. Zhang. 2006. A generalized moving-
model validation using finite experimental data. IEEE Trans- boundary model for transient simulation of dry-expansion
actions on Automatic Control 41(8):1094–105. evaporators under larger disturbances. International Journal
Tummescheit, H. 2002. Design and implementation of object- of Refrigeration 29(7):1119–27.
oriented model libraries using Modelica. Ph.D. Thesis, Lund, Zhang, W.-J., C.-L. Zhang, and G.-L. Ding. 2009. On three forms
Sweden, Lund Institute of Technology. of momentum equation in transient modeling of residential
Wang, H., and S. Touber. 1991. Distributed and non-steady- refrigeration systems. International Journal of Refrigeration
state modelling of an air cooler. International Journal of 32(5):938–44.
Refrigeration 14(2):98–111.

You might also like