Wallace MPC

Embed Size (px)

Citation preview

  • 8/18/2019 Wallace MPC

    1/14

    Energy efficient model predictive building temperature control$

    Matt Wallace a, Ryan McBride a, Siam Aumi a, Prashant Mhaskar a,, John House b, Tim Salsbury b

    a Department of Chemical Engineering, McMaster University, 1280 Main Street West, Hamilton, Ontario, Canada L8S 4L8b Johnson Controls Inc., 507 E. Michigan Street, Milwaukee, WI 53202, United States

    a r t i c l e i n f o

     Article history:

    Received 4 March 2011

    Received in revised form

    5 July 2011Accepted 8 July 2011Available online 29 July 2011

    Keywords:

    Vapor compression cycle

    Temperature control

    Building control

    Energy efficient control

    Model predictive control

    EnergyPlus

    a b s t r a c t

    Many systems used in buildings for heating, ventilating, and air-conditioning waste energy because of 

    the way they are operated or controlled. This paper explores the application of model predictive control

    (MPC) to air-conditioning units and demonstrates that the closed-loop performance and energy

    efficiency can be improved over conventional approaches. This work focuses on the problem of 

    controlling the vapor compression cycle (VCC) in an air-conditioning system, containing refrigerant

    which is used to provide cooling. The VCC considered in this work has two manipulated variables that

    affect operation: compressor speed and the position of an electronic expansion valve. The system is

    subject to constraints, such as the range of permissible superheat, and also needs to regulate

    temperature variables to set points. An MPC strategy is developed for this type of system based on

    linear models identified from data obtained from a first-principles model of the VCC. The MPC strategy

    incorporates economic measures in the objective function as well as control objectives. Tests are carried

    out on a simulated VCC system that is linked to a simulation of a realistic building that is developed in

    the U.S. Department of Energy Computer Simulation Program, EnergyPlus. The MPC demonstrated

    significantly better tracking control relative to conventional approaches (a reduction of 70% in terms of 

    the integral of squared error for step changes in the temperature set-point), while reducing the VCC

    energy requirements by 16%. The paper describes the control approach in detail and presents results

    from the tests.

    &   2011 Elsevier Ltd. All rights reserved.

    1. Introduction

    Environmental concerns as well as increased fuel prices have

    brought energy efficiency to the forefront of research priorities.

    Canada currently ranks as the world’s sixth largest user of 

    primary energy (such as fossil fuels, nuclear fuels, hydro power,

    etc.). In Canada, approximately 30% of the energy obtained from

    primary sources of energy is consumed in the commercial and

    residential sectors of the economy (Behidj et al., 2009). In these

    sectors, a significant portion of the energy is used towards

    meeting the thermal and electrical energy demands in buildings.

    Recent government reports estimate that through more efficientbuilding operation, the total energy consumption by the com-

    mercial and residential sectors can be reduced by 15–20% (Behidj

    et al., 2009).

    The operating efficiency of a building is influenced by many

    factors and can be improved at various points over its lifespan.

    Prior to construction, using design standards that incorporate

    energy and environmental concerns is often the first step to

    achieving an energy efficient building design. However, these

    design standards alone are not sufficient to ensure that a building

    remains energy efficient in response to changing energy and

    environmental standards. Once constructed, with the advent of 

    more energy efficient technology (i.e., EnergyStar certified tech-

    nology), the building can be appropriately retrofitted to meet

    more stringent energy and environmental standards. Finally, the

    energy efficiency of existing buildings can be improved through

    better control of their heating, ventilation, and air-conditioning

    (HVAC) systems (see  American Society of Heating,  for a detailed

    description of the common components of an HVAC system),

    which regulate building comfort (temperature and humidity) andaccount for 30–50% of the total energy consumption in buildings

    (Albieri et al., 2009). The focus of the present work is to

    demonstrate (via simulations) this improved efficiency achievable

    through use of advanced (model based) control techniques. To

    this end, we utilize an existing model of a VCC and couple it with

    a building model to approximate a (reasonably) realistic scenario

    of a roof top unit providing cooling to a room in a building using

    air as the only cooling medium.

    A vapor compression cycle (VCC) refers to a type of thermo-

    dynamic machinery that transfers heat using a compressible fluid

    referred to as the refrigerant. The most common realization of a

    VCC consists of four components: a compressor, condenser,

    Contents lists available at  ScienceDirect

    journal homepage:  www.elsevier.com/locate/ces

    Chemical Engineering Science

    0009-2509/$- see front matter  &  2011 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.ces.2011.07.023

    $This work is a collaborative effort between Johnson Controls Inc. and the

    McMaster Advanced Control Consortium. Corresponding author.

    E-mail address:  [email protected] (P. Mhaskar).

    Chemical Engineering Science 69 (2012) 45–58

    http://-/?-http://www.elsevier.com/locate/ceshttp://localhost/var/www/apps/conversion/tmp/scratch_1/dx.doi.org/10.1016/j.ces.2011.07.023mailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_1/dx.doi.org/10.1016/j.ces.2011.07.023http://localhost/var/www/apps/conversion/tmp/scratch_1/dx.doi.org/10.1016/j.ces.2011.07.023mailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_1/dx.doi.org/10.1016/j.ces.2011.07.023http://www.elsevier.com/locate/ceshttp://-/?-

  • 8/18/2019 Wallace MPC

    2/14

    expansion valve, and an evaporator. In the VCC, the refrigerant

    circulates through the four components, undergoing various

    thermodynamic changes, which, in turn, influence external con-

    ditions. Mathematically, the dynamics of the refrigerant states

    and external conditions are modeled using a set of coupled

    nonlinear ordinary differential and algebraic equations, resulting

    in a complex differential-algebraic equation system. The control

    objectives are typically defined in terms of degrees of superheat

    in the refrigerant at the evaporator exit and the air temperature atthe evaporator exit (the supply air temperature). Ensuring that

    superheated refrigerant exits the evaporator is of utmost impor-

    tance in preventing physical damage in the VCC, as liquid

    refrigerant can damage the mechanical components used in the

    compressor. The manipulated variables include the compressor

    speed, the air flow rates and the expansion valve opening. While

    research activity has been strong for a long time on the design/

    material and, to a smaller extent, modeling, of the various

    components of the VCC (see  Rasmussen, 2005   for details), there

    has been a recent trend towards improved control of the VCC for

    improved energy efficiency.

    Traditional VCC control strategies have included PID/PI decen-

    tralized control (i.e., multiple independent single-input–single-

    output (SISO) controllers) and simple on/off control. The latter

    limits overall efficiency due to large power requirements and

    significant thermal inertia during start-up transients while the

    former’s efficiency is limited by extensive interactions and non-

    linearity in the VCC system dynamics and the presence of input

    constraints. Approaches to improve the performance of conven-

    tional PID/PI controllers can be categorized into those which

    attempt to decouple the VCC dynamics to improve SISO controll-

    ability (Keir and Alleyne, 2007;   Jain et al., 2010) and adaptive

    control approaches which attempt to account for the process

    nonlinearity (i.e. time-varying process gains) (Lin and Yeh, 2007;

    Zhu et al., 2001).

    Among the decoupling approaches, the most straightforward

    extension has been to employ linear decouplers to remove

    interactions among the individual control loops. However, the

    effectiveness of this approach is entirely contingent on the model

    accuracy, and a poor model can lead to closed-loop performance

    degradation. Another decoupling approach has been to use linear

    combinations of available VCC measurements as the controlled

    variables instead of traditional controlled variables ( Jain et al.,

    2010). The adaptive control approaches, on the other hand,

    attempt to account for the nonlinear nature of the VCC dynamics

    by updating the PI/PID tuning parameters online using a model of 

    the process. However, these approaches are typically restricted to

    linear models (for computational reasons), implying the tuning

    parameter updates may be erroneously updated since the true

    process is highly nonlinear, leading to poor control performance.

    Despite these improvements, PI/PID control designs remain

    inherently based on a single-input–single-output framework

    and do not account for the presence of constraints and optimality.The control action prescribed by a controller that does not

    account for input constraints can result in performance degrada-

    tion or even closed-loop instability.

    One control method well suited to handling constraints and

    optimality is model predictive control (MPC). MPC is an optimiza-

    tion-based control approach in which the coupled, multiple-input–

    multiple-output nature of complex systems can be accounted for in

    determining the control action by using a model of the process. In

    model predictive control, a nonlinear or linear process model is

    used to evaluate the effect of candidate manipulated input trajec-

    tories via an objective function, and an optimization problem is

    solved to yield the manipulated input trajectory that minimizes the

    objective function while satisfying any constraints. Only the first

    piece of the input trajectory is implemented and the problem is

    re-solved at the next sampling time, using the new measured

    values of the process variables. One of the strengths of the MPC

    framework is the flexibility (and scope) in specifying optimality

    objectives through various terms in the objective function, or

    through constraints on the variables of interest. This, and the

    results available on the stability and feasibility properties of MPC

    formulations (see, e.g., Mhaskar et al., 2005, 2006; Mhaskar, 2006)

    make MPC a preferred candidate to be evaluated for possible use

    within building control structures. Recently, there have been manyexamples in the literature of the application of MPC for regulating a

    wide range of VCC or HVAC systems. In addition to the nature of 

    the system being regulated, the major differentiating feature

    among these MPC approaches is the complexity of the model used

    for predictions. Specifically, the predictive model may be a linear-

    ized version of a non-linear state-space model (Schurt et al., 2009;

    Sandipan et al., 2010;   Morosan et al., 2010), an empirically

    identified linear (Huang et al., 2009; Ma et al., 2010) or nonlinear

    model (Xi et al., 2007), or a first-principles non-linear model

    (Leducq et al., 2006;   Ma et al., 2010;   Sarabia et al., 2007). The

    majority of the MPC application examples have utilized a linear

    VCC model (either a linearized state-space model or an empirically

    identified input–output model). For example, in   Sandipan et al.

    (2010), an experimental chiller network, consisting of two chillers

    and multiple external heat exchangers, is regulated to satisfy the

    cooling load in addition to minimizing electricity costs. In  Huang

    et al. (2009), empirical first-order time-delay (FOTD) models are

    used in a robust MPC formulation for improving temperature

    regulation of an air-conditioning system. Specifically, several FOTD

    models are identified at various operating points, and based on the

    current operating conditions, the most appropriate FOTD model is

    used in the MPC optimization. Linear MPC applications in the

    context of building control include the work in   Morosan et al.

    (2010)   where a distributed MPC design is used to regulate the

    temperature of multiple zones in a building and minimize power

    consumption. That is, each zone is served by a separate HVAC unit

    under the control of a zone-specific MPC design to regulate the

    internal zone temperature. Another example of building control

    using (linear) MPC is available in  Ma et al. (2010) where weather

    data is incorporated into the design to determine the building zone

    temperature set-point. This allows for pre-cooling during non-peak

    periods and reduced power consumption compared to traditional

    pre-programmed HVAC unit control strategies. The existing results

    notwithstanding, there still exists a lack of results on the applica-

    tion of a MPC design to a detailed model of a VCC unit coupled with

    a realistic building model to evaluate the control performance in

    the presence of disturbances.

    Motivated by the above, this work evaluates the performance

    of an integrated temperature control framework via simulations.

    Specifically, we design a predictive controller for a stand-alone

    VCC unit and utilize it in a cascade control structure for tempera-

    ture control. The proposed control structure is implemented on a

    realistic building model that accounts for varying weather condi-tions and internal heat loads throughout the course of a day and

    the results are compared with a PI-based control structure. The

    rest of this manuscript is organized as follows. In  Section 2, we

    give an overview of the VCC and building models used in this

    work. Then, in   Section 3, we develop a control strategy for

    temperature control in a building zone. To this end, we first

    estimate an input–output model for the VCC in   Section 3.1   and

    then design an offset free predictive controller for a stand-alone

    VCC unit in   Section 3.2. This controller is subsequently incorpo-

    rated in a cascade control strategy to control a specific room

    temperature and then implemented on a realistic building model

    in Section 3.3. The performance of the proposed control strategy is

    compared with a conventional PI-based control strategy. Finally,

    we summarize our results in  Section 4.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–5846

  • 8/18/2019 Wallace MPC

    3/14

    2. Preliminaries

    In this section, we give an overview of an existing VCC model

    used in this work and point out the key limitations of the models

    and modifications. Next, we describe the building model and the

    software used for interfacing the VCC model with the building

    model. Note that model development is not the focus of the

    present work. A detailed model of the VCC and the building is

    used only to illustrate the control design and is described in thissection for completeness.

     2.1. VCC model overview

    An ideal VCC consists of four processes: isentropic compres-

    sion in a compressor, isobaric energy dissipation in a condenser,

    isenthalpic expansion in an expansion valve and isobaric energy

    absorption in an evaporator. An overlay of a VCC and the

    corresponding pressure-volume diagram of the refrigerant is

    shown in Fig. 1.

    In a VCC unit, the refrigerant enters the compressor as a

    superheated vapor and is compressed to a higher pressure,

    resulting in the superheated vapor having a higher temperature

    than the ambient temperature. From the compressor, the super-heated refrigerant vapor enters a condenser (typically placed

    outdoors), condensing to a sub-cooled liquid at the condenser

    exit as a fan blows the ambient air over the condenser. The high

    pressure sub-cooled liquid then flows into an expansion valve

    which decreases the pressure and temperature of the refrigerant,

    causing a liquid–vapor mixture to form. Then, the two-phase

    refrigerant mixture enters an evaporator that is exposed to the

    environment to be cooled. The environment temperature is above

    the temperature of the refrigerant, resulting in the evaporation

    and subsequent heating of the refrigerant to a superheated vapor

    at the evaporator exit. The air, in turn, is cooled and available as

    primary air to be distributed for cooling. The superheated vapor

    from the evaporator exit then flows into the compressor, com-

    pleting the cycle.

    The VCC model used in this work is adapted from the existing

    simulation package, Thermosys, developed at the Air Conditioning

    & Refrigeration Center (ACRC) at the University of Illinois at

    Urbana-Champaign. In this simulation package, the refrigerant is

    R-134a and is assumed to be cooling an air medium. The

    simulator consists of dynamic models for the condenser, eva-

    porator, and compressor and static models (i.e., algebraic equa-

    tions) for the expansion valve and piping. In the following

    subsections, a brief overview of the model components is pro-

    vided followed by a general mathematical representation. For a

    full description of the model components and a complete list

    of the equations and parameters, the reader is referred to

    Rasmussen (2005).

     2.1.1. Compressor 

    The compressor in the VCC model is a reciprocating compres-

    sor defined by its isentropic efficiency, Zk, which is the ratio of thework required for ideal adiabatic compression to the work

    required for actual compression, and its volumetric efficiency,ZV , which is the ratio of induced gas volume to the discharged gasvolume (swept volume). Note that in the present work we use a

    model of a variable speed compressor to illustrate the improved

    efficiency achievable by model-based control designs and the key

    interpretations remain applicable for other compressor types (on/

    off compressors, etc.). Demonstrating improved energy efficiency

    on other cooling units is the subject of future work and outside

    the scope of the present manuscript.

    The efficiency of the compressor depends on the pressure ratio

    of the outlet to inlet stream and the compressor RPM,  ok. Theyare obtained via (experimentally obtained) lookup tables of 

    efficiencies at various operating conditions. The mass flow rate

    in the compressor is modeled using the following static equation:

    _mk ¼ okV krkZV    ð1Þ

    where   V k   denotes the swept volume (a compressor parameter)

    and  rk   is the refrigerant density at the inlet. The term,   V krkZV ,characterizes the compressor capacity in terms of inlet refrigerant

    conditions. For the compressor energy balance, the dynamics of 

    the heat transfer during the transport of the refrigerant from the

    compression cylinder (where the compression takes place) to the

    shell (where the refrigerant exits) are taken into consideration.

    Specifically, the energy dynamics are modeled as a linear (in K),

    first-order differential equation:

    tk _h

    o

    k þhok ¼ K ðZk,h

    ik,h

    o

    kÞ ð2Þ

    where   hok   is the enthalpy of the outlet refrigerant,  tk   is a time

    constant (a compressor parameter), and  K ðÞ  is the gain. The gainis constant during integration and a nonlinear function of the

    isentropic efficiency, inlet refrigerant enthalpy,  hik, and the ideal

    isentropic outlet enthalpy,   ho

    k, which is determined by the

    refrigerant thermodynamic properties and inlet enthalpy.

    Remark 1.  In general, the compressor type (variable speed or on/

    off) is dependent on the specific application of the VCC (most

    existing compressors utilize an on/off strategy). Due to the

    limited range of operating speeds for an on/off compressor, the

    startup and shutdown of a setup equipped with this compressor

    type can draw considerably more energy during these operating

    conditions than a setup equipped with a variable speed compres-

    sor. Demonstrating the improvement over traditional on–off 

    setups (using a good model of such a unit) does remain an

    objective of future work, but is outside the scope of this manu-script. If an on/off type compressor is used in the VCC, modifica-

    tions can be made to the model (as presented in this section), as

    well as to the control design to ensure the implementation of the

    MPC control structure is feasible. In particular, any proposed

    control design for the VCC must account for the discrete nature of 

    the compressor operation. For instance, in the present work, the

    compressor RPM is a manipulated variable for VCC control and

    treated as a continuous variable. With an on/off type compressor,

    the compressor RPM is fixed when it is on and zero otherwise. In a

    model predictive control framework, this can be modeled either

    indirectly (by choosing the on/off durations as input variables) or

    explicitly using binary variables in the optimization problem. On

    the other hand, classical control approaches, such as PI control,

    have limited options in handling units with discrete operation.

    Volume

         P    r    e    s    s    u    r    e

    Gas

    Liquid

    Wet vapor (saturated conditions)

    Evaporator Compressor 

    Condenser 

    Heat from refrigerant

    Heat from process

    Valve

    Fig. 1.  VCC overlay on a pressure–volume diagram of a typical refrigerant.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–58   47

  • 8/18/2019 Wallace MPC

    4/14

    Another limitation of the current VCC model is the range of 

    compressor speeds over which the model is valid. In particular,

    the model does not remain valid for low RPM values, and as a

    result, the simulations have been carried out with the RPM

    restricted to this range to preserve the validity of the results.

    Enhancing the range of validity of the model also remains another

    direction of future work.

     2.1.2. Expansion valve

    The expansion valve is modeled as an isenthalpic process,

    meaning the inlet and outlet enthalpies are identical, with a mass

    flow rate given by the following expression:

    _mv ¼ C d

     ffiffiffiffiffiffiffiffiffiffiffiffiffiffirvDP v

    q   ð3Þ

    where  C d  is the valve discharge coefficient,  DP v  is defined as the

    pressure difference between the inlet and outlet refrigerant, and

    rv   is the maximum of either the sub-cooled liquid density or thesaturated liquid density at the inlet operating conditions. The

    discharge coefficient is determined by the valve opening and

    pressure differential and obtainable from experimental lookup

    tables. The sub-cooled and saturated liquid refrigerant densities

    are also obtained from lookup tables of the refrigerant’s thermo-

    dynamic properties.

    Remark 2.   An electronic expansion valve (EEV) is used in this

    VCC model, where the valve position is adjustable and in agree-

    ment with electronic expansion valves used in practice. The valve

    position of an actual EEV is proportionally adjusted through

    varying the frequency of a built-in step-motor, where the energy

    draw of this motor is minimal relative to the other energy-

    consuming components of a practical VCC (i.e., compressor and

    fan motors).

     2.1.3. Heat exchangers

    The dynamics of the VCC are dominated by the condenser and

    evaporator. Both heat exchangers are modeled as a long thin

    horizontal tube with one-dimensional fluid flow, negligible pres-

    sure drop (due to momentum change and viscous friction), andnegligible axial conduction.

    In both heat exchangers, the refrigerant may undergo multiple

    phase changes; accordingly, the refrigerant is modeled using a

    lumped parameter, moving boundary approach which accounts

    for different fluid regions (superheated vapor, saturated vapor–

    liquid, or sub-cooled liquid) and their time varying boundaries. In

    this approach, each fluid region is represented as a separate

    control volume (see Fig. 2) with corresponding states and para-

    meters. For the evaporator, there are two fluid regions: a two-

    phase region followed by a superheat region while the condenser

    has three fluid regions: a vapor region followed by a two-phase

    and a sub-cooled region. In each two-phase region, the refrigerant

    fluid properties are taken as the weighted combination of the

    saturated liquid and vapor properties. The mean void fraction,  g ,which is defined as the ratio of the vapor volume in a region to

    the total region volume, is used to weight the properties. For

    instance, the refrigerant density in a two-phase region is given by

    gr f  þ ð1gÞr‘  where r f   and r‘  are the saturated vapor and liquid

    densities, respectively. In the superheated and sub-cooled

    regions, the refrigerant properties, such as density and tempera-

    ture, are determined using the heat exchanger pressure (assumed

    constant) and the average regional enthalpy (the average of the

    inlet and outlet enthalpies).

    The mass and energy balance ordinary differential equations

    (ODEs) for each fluid region are derived from the governing

    partial differential equations (PDEs) for fluid flow in a tube. To

    yield a set of ODEs from the PDEs, the spatial dependence fromthe PDEs is removed after applying simplifying assumptions and

    Leibnitz’s rule on any differential with respect to the spatial co-

    ordinate, z . The full details of the modeling approach are available

    in   Rasmussen (2005). Eqs. (4) and (5) represent the governing

    refrigerant mass and energy balance PDEs (respectively) of a

    specific fluid region.

    @ðr Ac Þ@t 

      þ @  _m

    @ z   ¼ 0   ð4Þ

    @ðr Ac h Ac P Þ

    @t   þ

     @ð  _mhÞ

    @ z   ¼ piaiðT wT r Þ ð5Þ

    where r,   _m, and h  denote the refrigerant density, mass flow rate,and specific enthalpy (respectively),   Ac    is the heat exchanger

    cross-sectional area,  P  is the fluid region pressure,  pi  is the innerperimeter of the heat exchanger,  ai  is the heat transfer coefficientbetween the refrigerant and the heat exchanger inner wall,  T w is

    the wall temperature, and T r  is the refrigerant temperature. These

    PDEs are coupled with the following wall energy balance for each

    region:

    ðc  pr AÞw _T w ¼ piaiðT r T wÞ þ poaoðT aT wÞ

    where ðc  pr AÞw is the thermal capacitance of the tube wall per unitlength, po is the outer perimeter of the heat exchanger, and  ao   isthe heat transfer coefficient between the tube wall and the

    surrounding air with temperature   T a. After integrating   @  _m=@ z 

    and   @ð  _mhÞ=@ z   along the length of the tube using Leibnitz’s rule,

    the final set of ODEs for the heat exchangers can be arranged in

    the following matrix form:

    Zhð xh,uÞ _ xh ¼ f hð xh,u,d Þ

    where ZhðÞ and f hðÞ are a matrix and vector, respectively, and the

    heat exchanger state variables,   xh, include: the length of the

    superheat,   Lc ,1, and two-phase,   Lc ,2, regions in the condenser,

    the condenser wall temperatures in all three regions,  T w,c ,1, T w,c ,2,

    and   T w,c ,3, the constant condenser pressure,   P c , the condenser

    outlet refrigerant enthalpy, hoc , the length of the two-phase region

    in the evaporator, Le,1, the evaporator wall temperatures for both

    regions,   T w,e,1,   T w,e,2, the constant evaporator pressure,   P e, the

    evaporator outlet refrigerant enthalpy,   hoe, and the compressor

    outlet refrigerant enthalpy,  hok. The input vector,  u, elements are

    the compressor RPM, ok, and valve opening, vo. Note that in some

    VCC systems, the fan speeds for the air being blown over theevaporator and condenser (and therefore the mass flow rate of 

    air) may also be available for adjustment; however, for this VCC

    model, these are assumed constant. The disturbance vector,  d , is

    constituted of two measurable temperatures: the temperature of 

    Two-phase region Two-phase region Sub-cooled regionSuperheat region Superheat region

    Fig. 2.   Heat exchanger schematics showing the different fluid regions (adapted from  Rasmussen, 2005) which are modeled using the moving boundary approach.

    (a) Evaporator with two fluid regions. (b) Condenser with three fluid regions.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–5848

  • 8/18/2019 Wallace MPC

    5/14

    the air to be blown over the evaporator,  T ia,e, which is commonly

    referred to as the mixed air temperature, and the air temperature

    at the inlet of the condenser. The latter temperature is simply the

    ambient air temperature since the condenser is assumed to be

    outdoors and henceforth will be denoted by  T amb. The air supplied

    to the VCC evaporator is typically a mix of the zone (i.e., room)

    and ambient air. For example, the mixed air may be a mixture of 

    80% zone air and 20% ambient air.

     2.1.4. Mathematical representation

    The Thermosys VCC model comes in the form of a Simulink-

    based toolbox in Matlab. For this work, we extracted the VCC

    model ODEs and algebraic expressions from the source files and

    expressed the model as a differential-algebraic equation (DAE)

    system. In order to integrate the DAE system, algebraic equations

    are required to be satisfied at all integration steps. Integrating the

    DAE system in Matlab as opposed to running the Simulink model

    files yielded significant computational benefits with execution

    times for the same test period being reduced by over 70%. The

    VCC DAE can be expressed in the following general form:

    Zð x,uÞ _ x ¼ f ð x,u,d Þ

     g ð x,uÞ ¼ 0

     y  ¼ hð x,d Þ

    where ZðÞ and  f ðÞ again denote a matrix and vector, respectively,

     x   is the VCC state vector,   u   and   d   were previously defined in

    Section 2.1.3, g ð x,uÞ represents the algebraic expressions (used for

    modeling the piping and expansion valve), and  y  denotes the VCC

    outputs. The VCC outputs are defined to be the superheat of the

    refrigerant exiting the evaporator,  T s,e, and the air temperature at

    the evaporator exit, the so-called supply air temperature,  T oa,e. The

    superheat is defined as the number of degrees by which the

    refrigerant temperature at the evaporator exceeds its saturation

    temperature. As mentioned in   Section 1,   T s,e   is required to be

    maintained above 0   1C to protect against any liquid refrigerant

    entering the compressor and therefore required for safe andreliable compressor operation. In practice, the superheat is main-

    tained above 0   1C with a safety margin. The vector,  hðÞ, denotes

    the (nonlinear) output mapping function. The mapping function

    for the superheat is relatively straightforward whereas the func-

    tion to compute the supply air temperature consists of finding the

    root of a nonlinear equation as discussed next.

    The following discussion contains modifications of the supply

    air temperature calculation procedure found in the original

    Thermosys model. Specifically, we make corrections to the pro-

    cedure in the event of any condensation of the water vapor

    content in the air. To obtain the supply air temperature,  T oa,e, an

    energy balance for the wall side of the evaporator is solved.

    Assuming no energy accumulation in the evaporator walls, the

    heat absorption by the evaporator wall must equal the heat loss of the air:

    ao Ac X2i  ¼  1

    Le,iLe

    ðT a,eT w,e,iÞ ¼ H loss   ð6Þ

    The term,   ao Ac P2

    i ¼  1 ðLe,i=LeÞðT a,eT w,e,iÞ, represents the energy

    absorption by the evaporator where   T a,e   is the average air

    temperature around the evaporator:

    T a,e ¼ 12ðT 

    ia,e þT 

    oa,eÞ

    In the most general case (i.e., assuming there is condensation of 

    the water vapor content in the air), the heat loss of the air, H loss , is

    given by

    H loss ¼  _

    ma,

    ec  p,

    a,

    eðT 

    i

    a,

    eT 

    o

    a,

    eÞ þ  _

    ma,

    eðw

    i

    a,

    eh

    i

    W ,

    ew

    o

    a,

    eh

    o

    W ,

    eÞh‘ ,

    e   ð7Þ

    where   _ma,e  and  c  p,a,e  denote the mass flow rate and specific heat

    capacity of the dry  air being blown over the evaporator (assumed

    to remain constant),   wia,e   and   woa,e   denote the humidity ratio,

    which is defined as the ratio of the mass of water vapor in the air

    to the total dry air mass, of the inlet and outlet air (respectively),

    and hiW ,e  and  hoW ,e  denote the specific water vapor enthalpy at the

    inlet and outlet air conditions (respectively). The first term in Eq.

    (7) is the heat loss of the dry air and the only unknown variable in

    this term is  T o

    a,e  (the variable of interest). The second term is theenergy loss of the water vapor content in the air. In this term, the

    inlet humidity and water vapor enthalpy are readily computable

    from the known temperature (and pressure). If no condensation

    occurs, there is no change in the humidity ratio and  wia,e ¼ woa,e. In

    the case of condensation, the outlet air is saturated, implying the

    relative   humidity at the outlet,   foa,e, is 1.

    1 To compute the

    humidity ratio, its relationship with the relative humidity can

    be used to derive (Dincer and Rosen, 2007, Chapter 6):

    woa,e ¼ 0:622  f

    oa,eP 

    oW ,e,sat

    P afoa,eP 

    oW ,e,sat

    where  P a   is the known air pressure and  P oW ,e,sat  is the saturation

    pressure of water at T oa,e, which can be computed using Antoine’s

    equation. Meanwhile, the outlet water vapor enthalpy is com-puted using the standard formula:

    hoW ,e ¼ h f ,sat þ

    Z   T oa,eT ref 

    c  p,W ðT Þ dT 

    where   h f ,sat   is the heat of saturated water vapor at the air

    pressure,   T ref    is a reference temperature, and   c  p,W ðT Þ   is the

    (possibly) temperature-dependent specific heat capacity of water

    vapor.2 The third term in Eq. (7), h‘ ,e, represents the heat content

    in the condensed water if condensation occurs. Note that the

    negative sign is required in front of   h‘ ,e   since heat   losses   are

    written as positive energies in Eq. (7). The heat content in the

    water is given by

    h‘ ,e ¼   _ma,eðwia,ew

    oa,eÞc  p,W T 

    oa,e ¼   _m‘ ,ec  p,W T 

    oa,e

    where the product   _ma,eðwia,ew

    oa,eÞ  equals the mass of condensed

    water (follows from the definition of humidity ratio) or   _m‘ ,e   and

    c  p,W  is the constant heat capacity of liquid water.

    Having defined all the terms/variables in Eq. (6) and their

    dependence on the unknown supply air temperature,  T oe,a, a root

    finding algorithm can be applied to Eq. (6) to compute   T oe,a.

    Alternatively, an iterative (i.e., direct-substitution) procedure

    can be used where an initial guess for the supply air temperature

    is made, H loss is computed, and then the left hand side of Eq. (6) is

    solved for   T oa,e. If the difference between the newly computed

    supply air temperature and the initial guess exceeds a pre-defined

    tolerance, the newly computed value can be used to initialize the

    next iteration. Note also that when solving Eq. (6), an assumption

    regarding the occurrence of condensation has to be made. In thiswork, we first solve Eq. (6), assuming no condensation (i.e., with

    no  h‘ ,w   term), which is correct only if   T oa,e4T dp   where  T dp   is the

    dew-point temperature (computable from the air pressure). If 

    T oa,erT dp, the necessary condensation term is added to  H loss prior

    to solving Eq. (6) and the equation system is re-solved.

     2.1.5. VCC cooling capacity

    For proper regulation of a building zone temperature, the

    corresponding VCC unit for the zone must meet the cooling

    1 The relative humidity is defined as the ratio of the partial pressure of the

    water vapor to the saturation pressure of water at the system temperature.2 The same reference temperature is used for the computation of  how,e  and  h

    iw,e

    such that  T ref  disappears in Eq. (7).

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–58   49

  • 8/18/2019 Wallace MPC

    6/14

    capacity dictated by the highest possible ambient conditions and

    heat load disturbances. The nominal Thermosys VCC model has a

    maximum cooling capacity of 1127 W or 0.32 ton of refrigeration.

    This capacity is in agreement with that of a small experimental

    VCC used to validate the nominal model. Exploratory simulations

    revealed that this cooling capacity is insufficient (even with

    perfect control) to achieve the desired control objectives in terms

    of temperature control (see  Section 3.3) for the ambient condi-

    tions and heat load disturbances considered in the simulations. Asa result, the VCC model parameters are re-scaled such that the

    cooling capacity increases to 0.85 ton. Specifically, the mass flow

    rate of the circulating refrigerant,   _mr , along with the compressor

    volume, V k, are first increased. Next, the length of the evaporator,

    Le, is increased to allow longer contact of the supply air with the

    evaporator wall. The inner and outer cross-sectional areas of the

    evaporator,   Ae,i   and   Ae,o   respectively, and the mass of the eva-

    porator,   M e, are then increased by the same factor. Then, the

    diameter of the evaporator pipe,  De, and (dry) air mass flow rate,_ma,e, are increased to allow for more heat transfer from the air

    passing over the evaporator. To ensure that the additional heat

    absorbed by the refrigerant in the evaporator could be dissipated

    into the surrounding environment at the condenser, the same

    parameters for the condenser are increased by the same factors.

    Table 1   lists the nominal model parameters and the new ‘re-

    scaled’ parameters.

    Remark 3.   While the VCC model includes an EEV and variable

    speed compressor, enabling the VCC to achieve a varying cooling

    capacity, our current model does not capture the total cooling

    range associated with either an experimental or an industrial

    cooling unit. Specifically, the operating conditions corresponding

    to operating the VCC near its upper and lower cooling extremes

    are not captured in the scaled VCC model, as inaccuracies arose

    due to two factors: (1) the experimentally populated lookup

    tables for certain component and thermodynamic parameters

    (ZV ,   Zk,   C d, etc.) corresponded to different operating conditionsfor the scaled and original VCC, resulting in a smaller feasible

    operating range for the scaled system, and (2) the limiting natureof the EEV caused the minimal operating conditions associated

    with the compressor to be higher. Further reductions in the

    compressor RPM cause the refrigerant mass flow rate to decrease,

    however, the refrigerant mass flow rate will only converge to a

    steady-state value as long as the static valve opening is able to

    achieve the same decrease in flow. Eventually, the compressor

    RPM will reach a value where the static valve opening will not be

    able to reduce the refrigerant mass flow rate to the exit conditions

    corresponding to the specific RPM value, causing the refrigerant

    to never reach a steady-state value throughout the cycle, which

    will eventually result in liquid flowing into the compressor (i.e.,

    evaporator superheat region going to zero). This factor is solely a

    contribution of the choice of valve opening in the VCC and not

    affected by the current VCC model used. Future work will explore

    the potential benefit of using a non-adjustable valve in the VCC.

     2.2. Building model

    The key disturbances in the VCC model are the ambient air

    temperature (the air temperature at the condenser inlet) andmixed air temperature (the air temperature at the evaporator

    inlet). The ambient air temperature is naturally dictated by the

    outdoor weather conditions while the mixed air temperature is

    influenced by a variety of interacting factors including the degree

    of active heating/cooling in the room, the heating/cooling in

    adjacent rooms (if any), and various heat load disturbances,

    including the ambient air temperature. In this work, we utilize

    the EnergyPlus simulation package to provide realistic mixed and

    ambient air conditions based on a detailed building model and

    actual weather data.

    The building model in EnergyPlus accounts for building con-

    struction, surface geometries, and HVAC systems with the details

    based on the U.S. Department of Energy reference small office

    building model (U.S. Department of Energy). An important featureof the EnergyPlus building model is that it accounts for the typical

    daily variation of the internal gains in a building. Internal gains

    capture heat variations caused by a variety of realistic heat loads

    such as the movement of people and lighting schedules. As the air

    in a building is exposed to these internal gains, varying amounts

    of heat transfer occur from/to the air, causing fluctuations in the

    temperature and humidity of the zone temperature. This is

    reflected in the VCC unit as variations in the mixed air conditions.

    Recall that the mixed air is a mixture of the zone temperature and

    the ambient air temperature.

    The EnergyPlus building model used in this work considers a

    small (511 m2) single story office building in Chicago, Illinois, on a

    typical July day. The building is assumed to be divided into five

    occupied thermal zones, in which there is a conditioned floor areaof 150 m2 in the core zone, 113 m2 in perimeter zones 1 and 3,

    and 67 m2 in perimeter zones 2 and 4. The ground-to-ceiling

    height in all zones is 3 m. In total, the building houses 28 people

    at a standard occupant density of 5:38=100 m2 per zone. During

    peak operation, the building is occupied between the hours of 

    8:00 and 18:00 with the highest levels of occupancy. In this work,

    we assume that the thermal environment of perimeter zone 2 is

    regulated by the detailed VCC model described in  Section 2.1. All

    remaining thermal zones are assumed to be controlled by

    separate air-conditioning units (pre-)modeled in EnergyPlus,

    and their zone temperatures are maintained at a constant set-

    point temperature of 24   1C (to minimize inter-zone heat transfer).

    Fig. 3  shows the ambient temperature,   T amb, relative humidity,

    and the zone 2 internal gains over the course of the July day

    considered for the building model. The ambient conditions are

    obtained from historical data (in a data file) consisting of hourly

    measurements of the temperature and relative humidity.

     2.2.1. VCC-building model interface

    To link the building model in EnergyPlus with the VCC unit

    model in Matlab, data is exchanged between the two environ-

    ments over sockets using the Building Controls Virtual Test Bed

    (BCVTB) middle-ware (Wetter and Haves, 2008). This exchange is

    accomplished using a Matlab script file (exchangeDoublewith-

    Socket.m), which is included in the BCVTB library (see Fig. 4).

    In the EnergyPlus client, the ambient air and the air of zone two

    are mixed (80% zone air with 20% ambient air) to form the mixed

    air conditions for the VCC. Ideally, the data exchange sequence

     Table 1

    Nominal and re-scaled VCC model parameters to allow for a greater cooling

    capacity.

    Parameter Nominal Re-scaled Units

    _mr    7.76 103 1.13 102 kg/s

    V k   3.04 105 1.52 104 m3

    Le   11.46 57.29 m

     Ae,o   3.07 15.34 m2

     Ae,i   0.32 1.60 m2

    M e   1.55 7.74 kg

    De   8.90 103 3.57   102 m

    _me,a   0.243 2.43 kg/s

    Lc    10.7 53.5 m

     Ac ,o   2.79 13.97 m2

     Ac ,i   0.28 1.38 m2

    M c    4.66 23.30 kg

    Dc    8.10 103 3.24 102 m

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–5850

  • 8/18/2019 Wallace MPC

    7/14

    should be as follows: (1) subsequent to computing the mixed air

    conditions, the EnergyPlus client is paused momentarily, (2) the

    mixed and ambient air conditions are sent to Matlab, (3) in Matlab,

    the VCC model is integrated (with computed input values from a

    given control algorithm) and the corresponding cooling load iscomputed, (4) the cooling load is sent to the EnergyPlus model, and

    (5) the EnergyPlus model is un-paused and integrated forward

    using the newly computed cooling load. However, one limitation of 

    the interfaced environment is that concurrent  to sending the mixed

    and ambient air conditions to Matlab, the BCVTB software requires

    a cooling load from Matlab. That is, data is sent to and read from

    Matlab simultaneously because there is no effective way to pause

    the EnergyPlus model until the newly computed cooling load

    (corresponding to the sent data) is computed. Instead, during the

    simultaneous data exchange, the cooling load from the previous

    time step is read and implemented in EnergyPlus, thereby intro-

    ducing an input delay (of one sampling instant). To minimize the

    effects of this delay, the fastest available sampling time of 60 s is

    used for the EnergyPlus model.

    The cooling load of a VCC quantifies the heat absorption by the

    refrigerant in the evaporator from the inlet air. The accurate

    computation of the cooling load is essential for properly interfa-

    cing the VCC model. The total VCC cooling load is the sum of the

    sensible and latent cooling loads, which correspond to changes in

    the evaporator dry air temperature and humidity, respectively.

    The sensible cooling load is equivalent to the first term in  H loss  in

    Eq. (7). If the inlet air has a sufficiently high water vapor content,

    condensation may result, causing a humidity change and there-

    fore a non-zero latent cooling load. If no condensation occurs, the

    humidity ratio of the air does not change (as mentioned in  Section

    2.1.4); thus, the sensible cooling load equals the total cooling

    load. The energy change associated with the humidity change (or

    equivalently the condensation of the water vapor content in the

    air) is the latent cooling load,  X‘ , and is given by

    X‘  ¼   _m‘ ,eDh f ‘ 

    where   Dh f ‘    is the enthalpy of water condensation and   _m‘ ,erepresents the mass of condensed water and depends on the

    supply air temperature, which can be computed using the

    procedure described in Section 2.1.4.

    3. Temperature control

    In this section, we propose a temperature control framework

    for regulating the air temperature of zone 2 in the EnergyPlus

    building model (interfaced with Matlab). To this end, we first

    identify an auto-regressive exogenous (ARX) model for the VCC

    outputs using simulation data. Next, we utilize the model in an

    offset free predictive control design for the stand-alone VCC unit

    and compare its performance against PI control. Finally, we

    integrate the proposed predictive controller in a cascade control

    structure for regulating the zone temperature and implement the

    control structure on the interfaced building model.

    8

    24

    26

    28

    30

    Time (h)

       A  m   b   i  e  n   t   t  e

      m  p  e  r  a   t  u  r  e ,

       T  a  m   b

       (            °   C   )

    60

    70

    80

    90

       A  m   b   i  e  n   t

      r  e   l  a   t   i  v  e   h  u  m   i   d   i   t  y   (   %   )

    200

    300

    400

    500

       I  n   t  e  r  n  a   l   G  a   i  n  s   (   W   )

    10 12 14 16 18 8

    Time (h)

    10 12 14 16 18

    8

    Time (h)

    10 12 14 16 18

    Fig. 3.  Variations in the ambient temperature, relative humidity, and internal gains, which act as disturbances in the zone 2 EnergyPlus building model.

    Cooling load

    Read values

    Integrate

    VCC model

         E    x    c     h    a    n    g    e     D    o    u     b     l    e    w     i     t     h     S    o    c     k    e     t .    m Read values

    Output

    variables

    Integrate

    building model

    Matlab BCVTB EnergyPlus

    Fig. 4.  Schematic of the energy Plus-Matlab interface.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–58   51

  • 8/18/2019 Wallace MPC

    8/14

     3.1. ARX VCC model

    In the ARX type modeling approach, the process outputs at a

    specific sampling instant are assumed to depend linearly on the

    previous process conditions (defined by the process outputs and

    inputs). Mathematically, ARX models are defined as

     y ðkÞ ¼ Xn y

    i  ¼  1

     Ai y ðkiÞ þ Xnu

    i  ¼  1

    BiuðkiÞ þ Xnd

    i  ¼  1

    Cid ðkiÞ þv ðkÞ ð8Þ

    where  y ðkÞ   and  uðkÞ  are the process output and input vectors at

    sampling instant  k  (respectively),  d ðkÞ   is a vector of measurable

    disturbances, Ai, Bi, and Ci are model coefficient matrices (that are

    estimated using least-squares regression),  v ðkÞ is the noise vector,

    and n y, nu, and  nd denote the (maximum) number of time lags in

    the outputs, inputs, and disturbances (respectively) and define

    the order of the ARX model. For specific outputs, inputs, or

    disturbances which do not require the maximum number of lags,

    the appropriate elements in the coefficient matrices can be set to

    zero. For the VCC, the outputs, inputs, and measurable distur-

    bances were previously defined in   Section 2.1.4   as follows:

     y  ¼ ½T s,e   T oa,e

    T,  u ¼ ½ok   voT, and  d ¼ ½T amb   T ia,e

    T.

    To identify the ARX model coefficient matrices, pseudo ran-

    dom binary sequences (PRBS) are generated for the inputs anddisturbances for the typical operating range (see   Fig. 5   for a

    portion of the PRBS data) and subsequently implemented on the

    nonlinear stand-alone VCC model. Using the System Identification

    Toolbox in Matlab (which essentially solves the linear regression

    problem to compute the model coefficient matrices), the ARX

    model coefficient matrices for numerous lag choices are esti-

    mated. Among these models, the lag choice representing a good

    trade-off between the prediction accuracy and number of model

    parameters is summarized in Table 2. Fig. 6 compares the output

    prediction by the ARX model with the training data from the

    nonlinear model, demonstrating the prediction capability of the

    identified model.

    Remark 4.   A key objective of this work was to study the

    applicability of a predictive control based scheme for temperature

    control in the presence of realistic disturbances. In this work, we

    opt for an empirically identified input–output VCC model as the

    predictive model in the control design instead of a linearized

    state-space model (coupled with a state estimator). In general, a

    linearized state-space model of a nonlinear system at a specific

    operating point only captures the local dynamics around the

    linearization point and therefore calls for successive linearizationtechniques when used in an MPC framework to maintain reliable

    predictions. Another limitation of using a first-principles (deter-

    ministic) model as the foundation of the control design is that the

    model’s reliability is subject to the accuracy of numerous physical

    parameters (i.e., thermodynamic properties, etc.), which may not

    be known accurately. Additionally, many of the simplifying

    assumptions made during the model development can be violated

    in practice, further decreasing its validity. These reliability issues

    together with the inherent error introduced by linearizing a

    nonlinear model motivated the use of an empirical model for this

    work. From an industrial perspective, if a sufficiently large

    number of identical packaged units are produced, it may make

    economic sense to invest in the effort to develop a dedicated first

    principles model, or alternatively, generation of enough data to

    capture the model characteristics in a data-driven model.

    0

    1

    1.5

    ·103

    Time (h)

       R   P   M ,    ω   k

    0

    11

    11.5

    12

    12.5

    Time (h)

       V  a   l  v  e  o  p  e  n   i  n  g ,       v  o

       (   %   )

    20

    22

    24

    26

       M   i  x  e   d  a   i  r   t  e  m  p  e  r  a   t  u  r  e ,   T   i  a

     ,  e   (      °   C   )

    20

    25

       A  m   b   i  e  n   t  a   i  r   t  e  m  p  e  r  a   t  u  r  e ,

       T  a  m   b   (      °   C   )

    10 20 30 40 50 10 20 30 40 50

    0

    Time (h)

    0

    Time (h)

    10 20 30 40 50 10 20 30 40 50

    Fig. 5.  Portion of the input profiles used to generate output data for ARX model identification.

     Table 2

    Final ARX model lag structure.

    Output Lags

    T s,e   T oa,e   ok   vo   T amb   T ia,e

    T s,e   2 2 2 2 2 2

    T oa,e   1 1 1 1 1 1

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–5852

  • 8/18/2019 Wallace MPC

    9/14

    Remark 5.  Using an ARX model with a measurable disturbance

    vector as one of the predictors in a predictive control design

    effectively incorporates an element of feed-forward control into

    the design. That is, the control algorithm utilizes the measured

    disturbance vector to anticipate its effect and takes corrective

    action and further improve upon the achieved energy efficiency.

    However, for an MPC design with a prediction horizon greater

    than one, future disturbance measurements are required to make

    predictions over the horizon. In this case, the current measure-

    ments of the disturbances can be assumed to hold for the length

    of horizon. This is a common assumption used in MPC formula-

    tions that utilize disturbance measurements and is meaningful in

    the present context due to the different time scales at which the

    VCC evolves (small time scale) and the disturbance variables

    change (larger time scale).

    Remark 6.   The estimated ARX model predicts the VCC output

    behavior relatively well; however, the nonlinear nature of the

    process dynamics (i.e., varying process gains) cannot be fully

    captured using a single linear model. One approach to capture

    this nonlinearity is to identify multiple local linear models at

    various operating points and combine them with an appropriateweighting function during prediction. Recently, in   Aumi and

    Mhaskar (in press), a data-driven modeling methodology was

    proposed that unifies the concepts of ARX modeling, latent

    variable regression techniques, fuzzy   c -means clustering, and

    multiple local linear models in an integrated framework capable

    of capturing process nonlinearities. Specifically, plant data is first

    clustered using fuzzy   c -means clustering to identify the most

    suitable points for linearization and come up with a continuous

    weighting function for the individual models. Using this weight-

    ing function, the local linear model coefficients are  simultaneously

    estimated using latent variable regression tools, which allow for

    dimensionality and noise reduction. The same weighting function

    is then utilized to weight the individual models given an initial

    condition and inputs. The proposed modeling methodology hasbeen shown to be effective for identifying accurate models for use

    in MPC formulations (Aumi and Mhaskar, in press;   Aumi et al.,

    submitted), and represents one possibility for developing

    improved data-based models for use in the control design.

     3.2. Stand-alone VCC control

    In this section, we design and implement a predictive controller

    on the nonlinear VCC model using the model identified in the

    previous section and compare the closed-loop simulation results

    against PI control. The control objectives for stand-alone VCC

    control are to track a given set-point trajectory of the supply air

    temperature, to maintain reliable/safe operation by maintaining

    the superheat above 0   1C (see Section 2.1.4), and to maximize the

    energy efficiency by minimizing the compressor energy consump-

    tion (the largest energy consumer in the VCC). The closed-loop

    performance is evaluated in terms of the integral of squared error,

    ISESA, between the supply air temperature,  T oa,e, and its set-point

    trajectory, T oa,e,SP:

    ISESA ¼D

    t X

    i  ¼  1½T 

    o

    a,

    e,

    SPðiÞT 

    o

    a,

    eðiÞ2

    where   i   indexes the sampling instant,  Dt   is the sampling period

    (60 s), and   K   is the total number of sampling instants in the

    simulation. To quantify the energy demand associated with a

    control design, the instantaneous compressor power is summed

    over the simulation time, yielding a measure of the total energy

    consumption, TEC (see Section 2.1.1 for the variable definitions):

    TEC¼Dt XK i  ¼  1

    _mkðiÞ½hokðiÞh

    ikðiÞ

    Z

    where Z  is the combined total efficiency of the compressor, whichis the product of the power and the mechanical efficiencies (known

    parameters).

     3.2.1. MPC control design and implementation

    Consider a VCC system for which the ARX model for its outputs

    has been computed. For the proposed predictive control design,

    the inputs to the VCC at sampling instant   i   are computed by

    solving the following constrained optimization problem:

    minuminruðkÞrumax

    XP k  ¼  1

    J ^ yn

    2ðkÞ y2,SPðkÞJQ  þJu1Jr  þJDuJR 

    subject to  :   DuminrDuðkÞrDumax

    ^ y ðkÞ ¼Xn yi ¼  1

     Ai ^ y ðkiÞ þXnui  ¼  1

    BiuðkiÞ þXndi  ¼  1

    Cid ðkiÞ   for  k ¼ 1, . . .  , P 

    ^ y n

    ðkÞ ¼   ^ y ðkÞ þaþbðiÞ

     y1,minr ^ yn

    1ðkÞr y1,max

    a ¼ k½ y ð0Þ ^ y ð0Þ

    b1ðiÞ ¼ b1ði1Þ þ g 1 maxf0,½ y1,min y1ð0Þgþ g 2  maxf0,½ y1ð0Þ y1,maxg

    b2ðiÞ ¼ b2ði1Þ þ f ½ y2ð0Þ y2,SPð0Þ

    where the notation,  J    JQ , refers to the weighted norm, defined by

    J xJQ  ¼ xTQ  x and Du denotes a vector in which each element is the

    difference between successive input moves. The weighting

    matrices are diagonal and used to trade-off the relative impor-

    tance of the different control objectives. The plant measurement

    at the current sampling instant  i  corresponds to  k ¼0 or  y ð0Þ.

    In this MPC formulation, the control objective of supply air

    temperature set-point tracking is addressed by penalizing the

    0

    10

    15

    20

    25

    Time (h)

       S  u  p  e  r   h  e  a   t ,   T

      s ,  e

       (            °   C   )

    Nonlinear model

    ARX model

    14

    16

    18

    20

    22

       S  u  p  p   l  y

      a   i  r   t  e  m  p  e  r  a   t  u  r  e ,

       T

      o  a ,  e

       (            °   C   )

    Nonlinear model

    ARX model

    10 20 30 40 50 0

    Time (h)

    10 20 30 40 50

    Fig. 6.  Comparison of the output prediction by the ARX model with the nonlinear model for the input and disturbance profiles in  Fig. 5.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–58   53

  • 8/18/2019 Wallace MPC

    10/14

    deviation between the predicted supply-air temperature from its

    set-point,  y2,SPðkÞ, weighted by  Q . The predicted superheat is also

    bounded between y1,min and  y1,max. To reduce the energy consump-

    tion associated with the control action, the absolute value of RPM is

    also penalized using the weight  r . The inputs are constrained in a

    range for which the nonlinear VCC model is known to be valid. In

    addition to using hard constraints for the input rates, excessive

    input movements are penalized in the objective function using a

    move suppression factor with the weighting matrix,   R . Whentuning the different weighting matrices, the highest importance

    was initially given to tracking the supply air set-point. Subse-

    quently, the remaining weighting matrices were adjusted appro-

    priately to achieve relatively smooth input behavior.

    To achieve offset free performance, a disturbance/bias term is

    added to the model predictions that is expressed by combining

    two constant  terms, a  and  bðiÞ. The first term, a, is the disturbancedue to plant-model mismatch at the current sampling instant,

    multiplied by a tuning parameter,  k. Specifically,  a  is defined asthe difference between the predicted outputs at sampling instant

    i  from the previous control calculation and the measured outputs

    at  i . The  bðiÞ  term is the summation of tracking errors up to and

    including sampling instant i. For the superheat (output 1 or  y1), a

    non-zero tracking error at  i  is used only if the current measure-

    ment exceeds the minimum or maximum superheat. The  b   term

    essentially ‘persists’ and influences the control action until the

    offset is eliminated. It can be understood as operating the same

    way as the integral mode in a PI controller. The tuning para-

    meters, g 1, g 2, and f , are used to trade-off the input aggressiveness

    and the amount of offset. The list of tuning parameters which

    resulted in offset free performance while maintaining relatively

    smooth input behavior is tabulated in   Table 3   along with the

    constraint bounds. Fig. 7 demonstrates the effect of the  a and b. Inthe nominal case (no corrections), there is considerable offset in

    the supply air temperature. After adding the feedback term to

    account for plant-model mismatch, this offset is significantly

    reduced but not eliminated. Zero offset is only achieved after

    also including the error accumulation term in the formulation.

    Next, closed-loop simulation results for MPC and PI control are

    compared. For these simulations, constant disturbances are

    assumed. That is, the ambient air conditions (temperature and

    humidity) and the inlet air temperature to the evaporator (the

    mixed air temperature) are maintained at constant values. Using

    the results in Keir and Alleyne (2007), for the PI loop pairing, the

    supply air temperature is paired with the compressor RPM while

    the superheat is paired with the expansion valve opening. The

    superheat set-point for the PI controller is specified to be 10   1C

    (see   Remark 7). The PI controllers are initially tuned using the

    internal model control tuning method and fine-tuned to minimize

    the integral of absolute error while maintaining relatively smooth

    input trajectories.

    Fig. 8   displays the closed-loop VCC input and output variable

    responses for the two control strategies and   Table 4  summarizes

    their control performances using the metrics previously discussed

    and also the settling times for the supply air temperature,  t settSA , for

    the different set-point step changes. As shown in   Fig. 8, the

    proposed MPC design is able to provide better tracking performance

    of the supply air temperature for the different set-point changes

    with similar settling times and lower energy consumption. The

    third supply air set-point change (to approximately 23.7  1C) is an

    infeasible set-point for the VCC cooling capacity, but note that the

    predictive controller is able to drive the supply air temperature

    closer to this set-point compared to the PI controller. Note, how-

    ever, that the infeasibility is merely a result of the model not being

    valid at low RPM (or as low as required) to provide less cooling.

    For the MPC design, the superheat is permitted to ‘float’

    between its minimum and maximum value whereas for PI

    control, the superheat is maintained at the constant safety margin

    of 10   1C. This additional ‘degree of freedom’ for the predictive

    controller leads to more accurate tracking and better overall

    control performance. Note that if the superheat was prescribed

    to be maintained at a constant value of 10   1C for the MPC design

    as well, the corresponding closed-loop results would be similar to

    those obtained when using the PI controller. In regard to the

    energy efficiency, the MPC design required 8% less energy

    compared to the PI controller. This is a consequence of using

    higher valve openings and lower RPM values resulting from the

    multivariable nature of the MPC controller and the ability to allow

    the superheat to ‘float’ between acceptable values.

    Remark 7.  For the PI closed-loop simulation, the safety margin for

    the superheat is specified to be 10   1C. This represented a rough

    lower bound for the superheat set-point for reliable simulations

    under PI control. When the VCC model is interfaced with the

    building model and the superheat set-point is prescribed to be less

    than 10   1C, the PI controller drives the superheat to a negative

    value, resulting in a failed simulation. Note that in practice, a VCC

    unit has protections to ensure against negative superheat values;

    however, such protections are not considered in the existing VCC

    model. Simulation studies also revealed that by increasing the

    superheat set-point to 20   1C, the supply air tracking performance

    can be substantially improved. However, maintaining the superheat

    at a higher safety margin requires lower valve openings, which, in

    turn, results in the PI controller prescribing higher RPM values to

     Table 3

    MPC tuning parameters.

    Parameter Value

    P    4

    Q    950

    R    diag{0.004,0.5}

    r    350/17002

    f y1,min, y1,maxg   {3.5, 20}

    fumin,umaxg   {[678.8 6]T, [1700 15]T}

    fDumin,Dumaxg   f½200   1T,½200 1Tg

    k   ½0:2 0:50T

    f g 1 , g 2g   {6, 0.3}

     f    0.01

    0

    23.1

    23.3

    23.4

    23.5

    Time (h)

       S  u  p  p   l  y  a   i  r   t  e  m  p  e  r  a   t  u  r  e ,

       T  o

      a ,  e   (      °   C   )

    SPNominal +

    1 2 3 4 5 6 7 8

    23.2

    23

    Fig. 7.   Supply air temperature responses using various combinations of the bias

    terms in the proposed MPC design.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–5854

  • 8/18/2019 Wallace MPC

    11/14

    track the supply air temperature. Thus, there is a trade-off between

    the improved tracking performance and increased power consump-

    tion. It is also worth noting that the MPC design still offered better

    tracking performance (in addition to being more energy efficient by

    23%) than the PI controller by 10% even with the increased super-

    heat set-point of 20   1C.

    Remark 8.  Note that if the model is allowed to operate over the

    entire range of RPM, it is possible that the PI controller could be

    used to keep the superheat value at a fixed, low set-point, this

    would result in the RPM being able to change over the entirerange to provide minimum cooling where required, and at other

    times, providing additional cooling using minimal RPM. In such a

    scenario, the energy efficiency of a PI control structure would be

    comparable with the MPC in the current form. In such a scenario,

    however, possible nonlinear (and more importantly, non-mono-

    tonic) dependence of energy efficiency on the RPM would (and

    could) be incorporated within the MPC controller to provide more

    efficient operation over conventional control structures.

     3.3. Energy efficient temperature control framework

    In this section, we integrate the proposed MPC design for

    stand-alone VCC control in a cascade control structure for energy

    efficient temperature control in zone 2 of the EnergyPlus building

    model. Note that the main purpose of the interfacing is to

    demonstrate superior control of the cooling device subject to

    realistic disturbances (induced by the interfacing and use of 

    weather data). The primary control objective we consider is to

    maintain the zone 2 temperature, T zone, within acceptable comfort

    standards in the presence of disturbances brought on by varying

    ambient conditions, changes in the internal gains, and zone

    interactions (see   Section 2.2). The secondary control objectives

    are the stand-alone VCC control objectives listed in   Section 3.2.

    The comfort standards we consider are inspired by those used by

    the American Society of Heating, Refrigeration and Air-Condition-

    ing Engineers (ASHRAE). For typical summer conditions, assuming

    that the room occupants are wearing light clothing, the ASHRAE

    comfort standards (ASHRAE 55-2004) entail maintaining the zone

    temperature between 22.3 and 24.7   1C and for this work, a zone

    temperature set-point,   T zone,SP, of 24   1C is selected. Another

    important ASHRAE comfort standard is that the zone temperature

    not drift, which is defined as the temperature violating a band

    around the set-point for longer than 15 consecutive minutes. In

    this work, a  70.5   1C band around the set-point is considered.

    The proposed control structure for meeting the temperature

    control objectives is shown in Fig. 9. The cascade control structurewas motivated by the time-scale of the VCC dynamics compared

    to the zone temperature dynamics. Step tests in the VCC inputs

    (compressor RPM and valve opening) revealed the supply air

    and superheat temperatures evolve roughly in the same time

    scale (1–10 min) whereas the zone air temperature dynamics

    were significantly slower (nearly 50 min). Varying internal gains

    and ambient conditions act as disturbances to the zone air

    temperature; however, by using a cascade control structure, the

    relatively faster dynamics of the inner loop are exploited to

    eliminate these disturbances (using the VCC inputs) before they

    significantly affect the zone temperature.

    In this cascade control structure, the inner loop consists of a

    stand-alone VCC controller (either the predictive controller or PI

    controllers designed in   Section 3.2.1). The outer loop is used to

    0

    5

    10

    15

    20

    25

    Time (h)

       S  u  p  e  r   h  e  a   t ,   T  s ,  e

       (            °   C   )

    MPCPI

    23

    23.2

    23.4

    23.6

    23.8

       S  u  p  p   l  y

      a   i  r   t  e  m  p  e  r  a   t  u  r  e ,

       T

      o  a ,  e

       (            °   C   )

    SP

    MPCPI

    1

    ·103

       R   P   M ,       k

    MPCPI

    6

    8

    10

    12

    14

       V  a   l  v  e  o  p  e  n   i  n  g ,       v  o

       (   %   )

    MPCPI

    1.2

    1.1

    0.9

    0.8

    0.7

    5 10 15 0

    Time (h)

    5 10 15

    0

    Time (h)

    5 10 15 0

    Time (h)

    5 10 15

    Fig. 8.  Closed-loop output and input profiles for the VCC under MPC and PI control.

     Table 4

    Stand-alone VCC closed-loop performance metrics.

    Metric Control strategy

    PI control MPC

    ISESA  (s  1C2)   837 222

    t settSA   (s)   1800, 1800, 900, 1620 1020, 1800, 1140, 4440TEC (kJ) 10017 9217

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–58   55

  • 8/18/2019 Wallace MPC

    12/14

    regulate the zone temperature and is connected to the inner loop

    via the supply air temperature set-point. Based on the error

    between the zone temperature and its set-point (T zone,SP ¼

    24   1C), the outer loop prescribes a supply air set-point,  T oa,e,SP, to

    the inner loop controller. The zone temperature is sampled every

    15 min, which also corresponds to the frequency of the supply air

    set-point updates. Faster sampling times led to excessive fluctua-

    tions in the prescribed supply air set-point, resulting in poor

    tracking performance by the inner loop controller. The outer loop

    was tuned iteratively such that it yielded trackable supply air set-

    points by the VCC. The outer loop tunings were kept consistent for

    both control strategies in the inner loop.

    The zone temperature response for each control strategy is

    shown in   Fig. 10, which compares the efficacy of each control

    strategy in meeting the primary control objective. The MPC-based

    strategy is able to satisfy the comfort standards for the entire test

    period (any zone set-point violations lasted less than 15 min)

    while the comfort standards are violated for approximately the

    last 40 min of the test period when using the PI-based controller.

    From Fig. 3, after 16:00 or 4:00 P.M., there is a significant decrease

    in the internal gains owing to a decrease in the zone occupancy

    and also a decrease in the ambient temperature. The zone

    temperature response after 4:00 P.M. indicates that the MPC-

    based design is able to respond to these disturbances more

    effectively than the PI-based control strategy.

    With regard to the secondary control objectives,   Fig. 11   dis-

    plays the closed-loop VCC input and output profiles for the two

    inner loop control strategies. The performance metrics of the

    inner loops are shown in   Table 5. Similar to the results in the

    stand-alone VCC case, for the MPC-based design, the superheat is

    allowed to ‘float’ between its bounds and ended up evolving

    closer to its lower bound, allowing for better supply air tempera-

    ture tracking using considerably less compressor power. As

    shown in the supply air profiles in Fig. 11, in contrast to the PI-

    based design, the MPC-based controller provides an offset freesupply air temperature profile for the majority of the  feasible set-

    point values prescribed by the outer loop controller. To achieve

    this offset free performance (in addition to improved zone air

    temperature regulation), aggressive control action is prescribed.

    We note, however, that the key idea in the results with the

    interfaced system is not so much to demonstrate improved

    control of the zone conditions (which depends on several factors,

    including the ‘outer loop’) but more to evaluate the performance

    of the VCC control structure subject to realistic disturbances.

    Remark 9.   One natural extension of the proposed control struc-

    ture is to replace the outer PI loop with a model predictive

    controller. The main requirement for this extension is to identify a

    model between the supply air temperature and the zone air

    temperature. This can be identified through step-tests or more

    desirably, by generating PRBS-like sequences of the supply air.

    However, in any case, the resulting model will be dependent on

    the closed-loop dynamics of the stand-alone VCC controller. Due

    to the large variation in the time scales between the zone and VCC

    dynamics, in addition to the zone air being affected through a

    single VCC output variable (supply air temperature), it is advan-

    tageous to use separate MPC designs for each level of the cascade

    rather than using one model predictive controller to regulate the

    zone conditions. As discussed in   Remark 5, a weather model/estimator can also be incorporated in the design by including an

    ambient temperature component in the model. In this case,

    through an economic objective function that considers varying

    electricity costs for the outer loop controller, operating costs

    can be reduced by pre-cooling as necessary. In addition to

    optimality, the benefits of using MPC in the outer loop include

    explicitly incorporating comfort specifications and accounting

    for the VCC cooling capacity in computing the supply air set-

    points. Finally, while we use temperature as the comfort measure

    in this work, other measures of comfort, such as a Predicted Mean

    Vote (Federspiel and Asada, 1994;  Hanna, 1997;   Brager and de

    Dear, 1998; Jones, 2002; Ye et al., 2003; Baus et al., 2008) can be

    readily incorporated in the objective function in the MPC

    formulation.

    +

    T zone,SPPI   MPC/PI

    T oa,e,SP

    Condenser   Evaporator

    Compressor

    Valve

    EnergyPlus

    Model

    vo

    T s,e ,T   oa,e

    T zone

    DisturbancesVCC

    Fig. 9.  Proposed cascade control structure for energy efficient temperature control.

    8

    24

    Time (h)

       Z  o  n  e  a   i  r   t  e  m  p  e  r  a   t  u  r  e   (            °   C   )

    MPCPI

    24.6

    24.4

    24.2

    23.8

    23.6

    23.4

    23.2

    9 10 11 12 13 14 15 16 17 18

    Fig. 10.  Air temperature in zone 2 when using a cascade control structure for

    temperature control with a PI or predictive controller in the inner loop.

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–5856

  • 8/18/2019 Wallace MPC

    13/14

    4. Conclusions

    In this work, control strategies were implemented on a

    realistic building model interfaced with a detailed VCC model

    which provided the cooling load for a specified zone in the

    building. The detailed VCC model was subject to realistic dis-

    turbances in the ambient and mixed air conditions as a result of 

    interfacing the two models. The zone air temperature in the

    building model was also subject to realistic ambient air condi-

    tions (obtained from actual weather data) and typical internal

    load variations. A cascade control structure was proposed to

    regulate the zone air conditions in the building model. In the

    proposed control structure, an outer loop regulates the zone air

    temperature by adjusting the set-point of the VCC supply air

    temperature using PI control. The inner loop regulates the VCC

    supply air and superheat by manipulating the compressor RPM

    and valve opening using an ARX-model-based predictive control-

    ler. The proposed control strategy demonstrated better distur-

    bance rejection ability in the zone air temperature than a PI-based

    cascade structure. Also the predictive control strategy was more

    energy efficient. The improved performance of the MPC-baseddesign stemmed from its incorporation of the coupled nature of 

    VCC dynamics (through the ARX model) in computing the control

    action. The ability in the MPC to ‘trade-off’ optimality considera-

    tions with tracking requirements enabled achieving improved

    set-point tracking while operating at lower RPM values (where

    possible) resulting in better energy efficiency.

    Notation

    Variable Description

    V    volume

    r   density

    Z   efficiency

    8

    21

    22

    23

    24

    Time (h)

       S  u  p  p   l  y  a   i  r   t  e  m  p  e  r  a   t  u  r  e ,   T

      o  a ,  e

       (            °   C   )

    SPMPC

    21

    22

    23

    24

    25

       S  u  p  p   l  y  a   i  r   t  e  m  p  e  r  a   t  u  r  e ,

       T  o

      a ,  e

       (            °   C   )

    SPPI

    5

    10

    15

    20

    25

       S  u  p  e  r   h  e  a   t ,   T

      s ,  e

       (            °   C   )

    MPCPI

    1

    ·103

       R   P   M ,       k

    MPCPI

    6

    8

    10

    12

    14

       V  a   l  v  e  o  p  e  n   i  n  g ,       v  o

       (   %   )

    MPC

    PI

    1.2

    0.8

    10 12 14 16 18 8

    Time (h)

    10 12 14 16 18

    8

    Time (h)

    10 12 14 16 18

    8

    Time (h)

    10 12 14 16 18

    8

    Time (h)

    10 12 14 16 18

    Fig. 11.  Closed-loop output and input profiles for the VCC when interfaced with the EnergyPlus building model under MPC and PI control.

     Table 5

    Inner loop performance metrics in the cascade control structure.

    Metric   Control strategy

    PI control MPC

    ISESA  (s  1C2)   70 978 20 987

    TEC (kJ) 5080 4284

    M. Wallace et al. / Chemical Engineering Science 69 (2012) 45–58   57

  • 8/18/2019 Wallace MPC

    14/14

    o   compressor RPMt   time constanth   specific enthalpy

    DP   pressure drop

    C d   discharge coefficient

    g   volumetric mean void fraction_m   mass flow rate

     A   cross-sectional area

    P    pressure p   perimeter

    a   heat transfer coefficientT    temperature

     x   state variable

    u   input variable

    d   disturbance variable

    vo   valve opening

    c  p   specific heat capacity

    L   length

    w   humidity ratio

    D   diameter

    M    mass

    t    time

    Subscript Description

    k   compressor

    v   valve

    h   heat exchanger

    e   evaporator

    c    condenser

     f    vapor

    ‘    liquid

    r    refrigerant

    w   wall

    a   air

    amb   ambient

    W    water

    sat   saturationdp   dew point

    V    volumetric

    s   superheat

    SA   supply air

    SP   set-point

    Superscript Description

    i   inlet

    o   outlet

     Acknowledgments

    Financial support from the Natural Sciences and EngineeringResearch Council of Canada through the Collaborative Research

    and Development Program (in collaboration with Johnson Con-

    trols Inc.) is gratefully acknowledged.

    References

    Albieri, M., Beghi, A., Bodo, C., Cecchinato, L., 2009. Advanced control systems forsingle compressor chiller units. Int. J. Refrig. 32 (5), 1068–1076.

    American Society of Heating, Refrigeration and Air-Conditioning Engineers, Inc.,

    ASHRAE Handbook-Heating, Ventillating and Air-Conditioning Systems and

    Equipment, ASHRAE.Aumi, S., Mhaskar, P. Integrating data-based modeling and nonlinear control tools

    for batch process control. AIChE J., in press.  doi:10.1002/aic.12720.Aumi, S., Corbett, B., Mhaskar, P., Clarke-Pringle, T. Data-based modeling and

    control of nylon-6,6 batch polymerization. IEEE Trans. Control Syst. Technol.,

    submitted for publication.Baus, Z.L., Nikolovski, S.N., Maric, P.Z., 2008. Process control for thermal comfort

    maintenance using fuzzy logic. J. Elec. Eng. 59 (1), 34–39.Behidj, N., Brugger, M., Demers, D., Kowal, A., Liu, Y., Warbanski, M., Yamada, F.,

    2009. Energy Efficiency Trends in Canada, Technical Report, Natural Resources

    Canada (09 2009). URL    /http://oee.nrcan.gc.ca/publications/statistics/

    trends09/S.Brager, G.S., de Dear, R.J., 1998. Thermal adaptation in the built environment: a

    literature review. Energy Build. 27 (1), 83–96.Dincer, I., Rosen, M., 2007. Exergy: Energy, Environment and Sustainable Devel-

    opment. Elsevier, Oxford, UK.Federspiel, C.C., Asada, H., 1994. User-adaptable comfort control for HVAC

    systems. J. Dyn. Syst. Meas. Control Trans. ASME 116 (3), 474–486.Hanna, R., 1997. Relationship between thermal comfort and user satisfaction in

    hot dry climates. Renew Energy 10 (4), 559–568.Huang, G., Wang, S., Xu, X., 2009. A robust model predictive control strategy for

    improving the control performance of air-conditioning systems. Energy Con-

    vers. Manage. 50, 2650–2658. Jain, N., Li, B., Keir, M., Hencey, B., Alleyne, A., 2010. Decentralized feedback