economic model predictive control of transport-reaction processes

29
Economic Model Predictive Control of Transport-Reaction Processes Liangfeng Lao, Matthew Ellis and Panagiotis D. Christofides Department of Chemical & Biomolecular Engineering Department of Electrical Engineering University of California, Los Angeles AIChE Meeting San Francisco, CA November 7, 2013

Upload: donga

Post on 08-Feb-2017

221 views

Category:

Documents


0 download

TRANSCRIPT

Economic Model Predictive Control of Transport-ReactionProcesses

Liangfeng Lao, Matthew Ellis and Panagiotis D. Christofides

Department of Chemical & Biomolecular Engineering

Department of Electrical Engineering

University of California, Los Angeles

AIChE Meeting

San Francisco, CA

November 7, 2013

INTRODUCTION

Transport-Reaction Processes

• Examples:

⋄ Fixed bed and tubular reactors

⋄ Fluidized reactors

⋄ Chemical vapor deposition processes

• Process state variables are characterized by strong spatial variations and

nonlinear behavior

⋄ Diffusion and convection phenomena

⋄ Complex reaction mechanisms / Arrhenius dependence of reaction rates on

temperature

• First-principles modeling leads to nonlinear parabolic partial differential equations

(PDEs)

TUBULAR REACTOR EXAMPLE

Concentration Trajectory

• Diffusion-convection-reaction described by a quasi-linear parabolic PDE system:

∂T

∂t= −v

∂T

∂z+

k

ρCp

∂2T

∂z2+

(−∆H

ρCp

)e−E/RTC2

A −hAs

ρCp(T − TC)

∂CA

∂t= −v

∂CA

∂z+DA

∂2CA

∂z2− k0e

−E/RTC2A

BACKGROUND ON CONTROL OF PARABOLIC PDES

• Standard approach: (Ballas, IJC, 1979; Ray, McGraw-Hill, 1981; Curtain, Springer-Verlag, 1978)

⋄ Derivation of ODE models using eigenfunction expansions

⋄ Controller design using methods for ODEs

⋄ High-dimensionality of the controller

• Synthesis of nonlinear low-order controllers: (Christofides, Birkhauser, 2001)

⋄ Derivation of low-order ODE models using Galerkin’s method and

approximate inertial manifolds

⋄ Nonlinear and robust controller synthesis

• Other approaches:

⋄ Passivity-based control approach (Ydstie et al., Syst. & Contr. Lett., 1997)

⋄ Backstepping boundary control (Krstic and Smyshlyaev, SIAM, 2008)

• Model predictive control with quadratic cost functions (Dubljevic et al., Inter. J. Rob. & Non.

Contr., 2006, Comp. & Chem. Eng., 2005)

• Economic model predictive control of transport-reaction processes is an open issue

BACKGROUND ON CONTROL OF PARABOLIC PDES

• Standard approach: (Ballas, IJC, 1979; Ray, McGraw-Hill, 1981; Curtain, Springer-Verlag, 1978)

⋄ Derivation of ODE models using eigenfunction expansions

⋄ Controller design using methods for ODEs

⋄ High-dimensionality of the controller

• Synthesis of nonlinear low-order controllers: (Christofides, Birkhauser, 2001)

⋄ Derivation of low-order ODE models using Galerkin’s method and

approximate inertial manifolds

⋄ Nonlinear and robust controller synthesis

• Other approaches:

⋄ Passivity-based control approach (Ydstie et al., Syst. & Contr. Lett., 1997)

⋄ Backstepping boundary control (Krstic and Smyshlyaev, SIAM, 2008)

• Model predictive control with quadratic cost functions (Dubljevic et al., Inter. J. Rob. & Non.

Contr., 2006, Comp. & Chem. Eng., 2005)

• Economic model predictive control of transport-reaction processes is an open issue

CONVENTIONAL MPC VS. ECONOMIC MPC

• Conventional MPC

⋄ Quadratic cost function for MPC (Q,R: positive definite matrices)

J =

∫ N∆

0

[(x(τ)− xset)TQ(x(τ)− xset) + (u(τ)− uset)

TR(u(τ)− uset)]dτ

⋄ Penalize the deviation of states and inputs from the steady-state

⋄ Force the system to the economically optimal steady-state

• Lyapunov-based Economic MPC (LEMPC)

⋄ General cost function which accounts for the process economics

J =

∫ N∆

0

L(x(τ), u(τ))dτ

⋄ Use MPC for economic optimization and process control

⋄ Need constraints for closed-loop stability

⋄ Key ideas of our approach:

Utilize Lyapunov-based techniques in the economic MPC formulation

Take advantage of the stability region of an explicit Lyapunov-based controller to

ensure closed-loop stability under LEMPC

CONVENTIONAL MPC VS. ECONOMIC MPC

• Conventional MPC

⋄ Quadratic cost function for MPC (Q,R: positive definite matrices)

J =

∫ N∆

0

[(x(τ)− xset)TQ(x(τ)− xset) + (u(τ)− uset)

TR(u(τ)− uset)]dτ

⋄ Penalize the deviation of states and inputs from the steady-state

⋄ Force the system to the economically optimal steady-state

• Lyapunov-based Economic MPC (LEMPC)

⋄ General cost function which accounts for the process economics

J =

∫ N∆

0

L(x(τ), u(τ))dτ

⋄ Use MPC for economic optimization and process control

⋄ Need constraints for closed-loop stability

⋄ Key ideas of our approach:

Utilize Lyapunov-based techniques in the economic MPC formulation

Take advantage of the stability region of an explicit Lyapunov-based controller to

ensure closed-loop stability under LEMPC

LYAPUNOV-BASED ECONOMIC MPC

Two Mode Control Strategy

(M. Heidarinejad, et al., AIChE J. 2012, JPC 2012; S&CL, 2012; 2013)

minimizeu∈S(∆)

∫ tk+N

tk

L(x(τ), u(τ))dτ

subject to ˙x(t) = f(x(t), u(t), 0)

x(tk) = x(tk)

u(t) ∈ U, ∀ t ∈ [tk, tk+N )

V (x(t)) ≤ ρ ∀ t ∈ [tk, tk+N ),

if V (x(tk)) ≤ ρ and tk < t′

∂V

∂xf(x(tk), u(tk), 0)

≤ ∂V

∂xf(x(tk), h(x(tk)), 0)

if V (x(tk)) > ρ and tk ≥ t′

• Economic cost function

• Dynamic model

• Initial condition

• Input constraints

• Mode 1: Boundedness

of closed-loop state

• Mode 2: Convergence

to the origin

LYAPUNOV-BASED ECONOMIC MPC

Two Mode Control Strategy

Ωρ

Ωρ

xs

x(t0)

x(t′)

• Mode 1: Boundedness

of closed-loop state

• Mode 2: Convergence

to the origin

MOTIVATION OF EMPC FOR PDE SYSTEMS

Tubular reactor

• Control objective: maximize the production rate of the product B subject to a

constraint on the available reactant material

⋄ Formulate a model predictive controller that works to directly optimize

objective:

J(x, u) =

∫ tf

0

∫ L

0

k exp

(−E

RT (z, t)

)(CA(z, t))

2 dz dt

⋄ Account for constraint on available reactant material:

1

tf

∫ tf

0

u(τ)dτ = uavailable

• Challanges of applying EMPC to a PDE system

⋄ Construct a finite dimensional model using finite difference method can result

in a high-order model (can be computational inefficient)

⋄ Formulate EMPC with a finite dimensional model of the system

PRESENT WORK

(Lao et al., Ind. & Eng. Chem. Res., 2013; J. Process Contr., submitted)

• Scope:

⋄ Nonlinear parabolic PDEs

⋄ State-feedback Economic Model Predictive Control for PDE systems using

low-order/high-order finite-dimensional models of the PDE system

⋄ Output-feedback Economic Model Predictive Control for PDE systems using

low-order/high-order finite-dimensional models of the PDE system

• Objective:

⋄ Optimally operate a nonlinear parabolic PDE system with EMPC

• Approach:

⋄ Formulate the transport-reaction process as an infinite-dimensional system

⋄ Use Galerkin’s method for modal decomposition of the system

⋄ Take advantage of the stability region of the Lyapunov-based controller

⋄ Perform state-feedback/output-feedback low-order/high-order EMPC design

⋄ Application to a tubular reactor

CLASS OF NONLINEAR PARABOLIC PDE SYSTEMS

• System description

∂x

∂t= A

∂x

∂z+B

∂2x

∂z2+Wu(t) + f(x(z, t))

yj(t) =

∫ 1

0

cj(z)x(z, t)dz, j = 1, · · · , p

⋄ Boundary conditions and initial condition:

∂xj

∂z= g0xj , z = 0;

∂xj

∂z= g1xj , z = 1; x(z, 0) = η(z, 0)

⋄ x ∈ Rn: state vector of the system

⋄ u(t) ∈ Rm: manipulated input vector

⋄ f(x(z, t)) ∈ Rn: nonlinear vector function

⋄ yj(t): the jth measured output

⋄ cj(z): jth sensor distribution function

⋄ z ∈ [0, 1]: the spatial coordinate

⋄ u(t) ∈ U = u ∈ Rm | |ui(t)| ≤ ui,max, i = 1, . . . ,m⋄ A, B, W , g0 and g1: constant matrices and vectors

APPLICATION TO A TUBULAR REACTOR

• A non-isothermal tubular reactor where an irreversible second-order reaction of

the form A → B (exothermic) takes place

• Dynamic model of the process:

∂T

∂t= −v

∂T

∂z+

k

ρCp

∂2T

∂z2+

(−∆H)

ρCpexp

(−E

RT

)CA

2 − hAs

ρCp(T − TC)

∂CA

∂t= −v

∂CA

∂z+DA

∂2CA

∂z2− k0 exp

(−E

RT

)CA

2

• Boundary conditions:

z = 0 :∂T

∂z=

ρCpv

K(T − Tf ),

∂CA

∂z=

v

DA(CA − CAf )

z = L :∂T

∂z= 0,

∂CA

∂z= 0

• Manipulated input: the inlet concentration of species A, CAf

DIMENSIONLESS PDE DYNAMIC MODEL

• Dimensionless form (combining the non-homogeneous part of the boundary

conditions with the differential equation):

∂x1

∂t= −∂x1

∂z+

1

Pe1

∂2x1

∂z2+ βT (Ts − x1) +BTBC exp

(γx1

1 + x1

)(1 + x2)

2

+ δ(z − 0)Ti

∂x2

∂t= −∂x2

∂z+

1

Pe2

∂2x2

∂z2−BC exp

(γx1

1 + x1

)(1 + x2)

2 + δ(z − 0)u

where δ is the standard Dirac function

• Boundary conditions:

z = 0 :∂x1

∂z= Pe1x1,

∂x2

∂z= Pe2x2

z = 1 :∂x1

∂z= 0,

∂x2

∂z= 0

• Process Parameters: Pe1 = 7, P e2 = 7, BT = 2.5, BC = 0.1, βT = 2, Ts = 0, Tf = 0

and γ = 10

GALERKIN’S METHOD

• Solve the eigenvalue problem of the spatial differential operator, Ai subject to:

Aiϕij = −dϕij

dz+

1

Pei

d2ϕij

dz2= λjϕij

dϕij

dz= Peiϕij , z = 0;

dϕij

dz= 0, z = 1; i = 1, 2, j = 1, . . . ,∞

• Apply Ps, Pf to get the infinite-dimensional nonlinear system

x = Ax+ Bu+ f(x), x(0) = x0

⋄ Ps, Pf : orthogonal projection operators in the space of ϕij

• Derive out set of infinite ODEs

dxs

dt= Asxs +Bsu+ fs(xs, xf ) (m)

dxf

dt= Afxf +Bfu+ ff (xs, xf ) (l)

y(t) = Csxs(t) + Cfxf (t)

⋄ xs(t) = Psx(t): state vector corresponding to slow eigenmodes

⋄ xf (t) = Pfx(t): state vector corresponding to fast eigenmodes

⋄ Cs: measurement operator corresponding to slow eigenmodes

EMPC FORMULATION

Economic Cost & Reactant Material Constraint

• Control Objective: Maximize the total reaction rate along the length of the

reactor over one operation period of tf = 1

• The economic cost function:

J(x, u) =

∫ tf

0

L(CA, T, u) dt

L(CA, T, u) =

∫ L

0

k exp

(−E

RT (z, t)

)(CA(z, t))

2 dz

• Since the reaction is second-order, can maximize the rate by feeding in the most

amount of material

• Consider the total amount of reactant material during one period to be fixed

(practical consideration)

• The reactant material constraint:

1

tf

∫ tf

0

u(τ)dτ = 0.5

EMPC FORMULATION 1

Low-order State-feedback EMPC Formulation

• EMPC formulation:

maxu∈S(∆)

∫ tk+N

tk

L(as(τ), u(τ)) dτ

s.t. as(t) = Asas(t) + Fs(as(t), 0) +Bsu(t)

umin ≤ u(t) ≤ umax, ∀t ∈ [tk, tk+N )

as(tk) =

∫ 1

0

x(tk)ϕsdz

u ∈ gk

[as(t)− ass]TP [as(t)− ass] < ρ, ∀ t ∈ [tk, tk+N )

⋄ Model based on first two modes (m = 2) computed from 200 measured

state-feedback points

⋄ ass: steady states value of the coefficients as

⋄ gk: the reactant material constraint

⋄ Prediction horizon and sampling time: N = 3, ∆ = 0.02 hr = 72 sec,

umin = −1 and umax = 1

CASE 1: LOW-ORDER STATE-FEEDBACK EMPC

Concentration Trajectory

Temperature Trajectory

0 0.2 0.4 0.6 0.8 1−1.5

−1

−0.5

0

0.5

1

1.5

u(t)

t

Low−order EMPC with Input ConstraintUniform Input Distribution

Manipulated Input Trajectories

• Initially, feed in the maximum allow-

able reactant material

• At t = 0.75, feed minimum allowable

reactant material to satisfy material

constraint

• Able to obtain the optimal trajectory

with a reduced-order model (2 modes)

CASE 1: LOW-ORDER STATE-FEEDBACK EMPC

Comparison with Uniform Material Distribution

Concentration Trajectory

Temperature Trajectory

0 0.2 0.4 0.6 0.8 10

10

20

30

40

50

60

70

t

Rea

ctio

n R

ate,

L(x

,u)

Low−order EMPC with Input ConstraintUniform Input Distribution

Reaction Rate Trajectory

• Average reaction rate from the sys-

tem under EMPC is 9.45% greater

than that from the system under

uniform in time reactant material

distribution

CASE 1: LOW-ORDER STATE-FEEDBACK EMPC

Comparison with Finite-difference EMPC

0 0.2 0.4 0.6 0.8 1−1.5

−1

−0.5

0

0.5

1

1.5

u(t)

t

Low−order EMPCFinite−Difference EMPCUniform Input Distribution

Manipulated Input Trajectory

0 0.2 0.4 0.6 0.8 10

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

Simulation Time

Eva

luat

ion

Tim

e (s

)

Finite Difference EMPCLow−Order EMPC

EMPC Computational Time Trajectory

• Finite-difference EMPC uses 400th− order ODE model

• Input, state and reaction rate trajectories are overlapping

• Low-order state-feedback EMPC shows significant computational efficiency

improvement

• Finite-difference EMPC based on a 4th− order ODE model does not compute the

optimal trajectory

EMPC FORMULATION 2

High-order State-feedback EMPC Formulation with State and Input Constraints

• Consider a state constraint on the maximum allowable temperature in the reactor

• EMPC formulation:

maxu∈S(∆)

∫ tk+N

tk

L(as(τ), af (τ), u(τ)) dτ

s.t. as(t) = Asas(t) + Fs(as(t), af (t)) +Bsu(t)

af (t) = Afaf (t) +Bfu(t)

umin ≤ u(t) ≤ umax, ∀ t ∈ [tk, tk+N )

amin ≤ a(t) ≤ amax, ∀ t ∈ [tk, tk+N )

as(tk) =

∫ 1

0

x(tk)ϕsdz, af (tk) =

∫ 1

0

x(tk)ϕfdz

a(t) = as(t) + af (t)

u ∈ gk

[a(t)− ass]TP [a(t)− ass] < ρ, ∀ t ∈ [tk, tk+N )

⋄ Slow subsystem based on the first 10 modes

⋄ Fast subsystem based on the next 190 modes

CASE 2: HIGH-ORDER STATE-FEEDBACK EMPC

State Constraint

0 0.2 0.4 0.6 0.8 1−1.5

−1

−0.5

0

0.5

1

1.5

u(t)

t

High−order EMPC with State & Input ConstraintsUniform Input Distribution

Manipulated Input Trajectory

0 0.2 0.4 0.6 0.8 11.4

1.6

1.8

2

2.2

2.4

2.6

2.8

3

3.2

x 1,m

ax

t

High−order EMPC with State & Input Constraints

Maximum Allowable Temperature

Maximum Temperature Trajectory

0 0.2 0.4 0.6 0.8 10

1

2

3

4

5

6

t

Rea

ctio

n R

ate,

L(x

,u)

High−Order EMPC with State & Input ConstraintsUniform Input Distribution

Reaction Rate Trajectory

• The reactor temperature must not ex-

ceed the maximum allowable operat-

ing temperature (state constraint)

• EMPC determines the optimal time-

varying operating strategy

• Average reaction rate from the system

under EMPC is 6.91% greater

EMPC FORMULATION 3

State Estimation Using Output Feedback Methodology

• State Estimation: Use a finite number, p, of measured outputs yj(t) (here, we

choose state-feedback measured points) (j = 1, · · · , p) to compute estimates of asand af with the assumption:

p = m

C−1s exists

• The estimates of the slow modes, as(t) is obtained from a direct inversion of the

measured output map:

as(t) = C−1s y(t)

• The fast dynamics af (t) can be ignored (given that Af includes eigenvalues with

large negative real part) compared with the slow dynamics as(t)

• An explicit form for the estimated fast modes, af :

af (t) = −A−1f [Bfu(t) + ff (as(t), 0)]

EMPC FORMULATION 3

Low-order Output-feedback EMPC

• EMPC formulation:

maxu∈S(∆)

∫ ttk+N

tk

L(as(τ), u(τ))dτ

s.t. ˙as(t) = Asas(t) + Fs(as(t), 0) + Bsu(t)

as(tk) = C−1s y(tk)

umin ≤ u(t) ≤ umax

u ∈ gk

as,min ≤ as(t) ≤ as,max

[as(t)− ass]TP [as(t)− ass] < ρ, ∀ t ∈ [tk, tk+N )

⋄ Model based on first 11/21 modes (m = 11/21) computed from 11 measured

state-feedback points (p = 11/21)

⋄ as(t): predicted slow mode function of as(t)

⋄ Prediction horizon and sampling time: N = 3, ∆ = 0.01 hr = 36 sec,

umin = −1 and umax = 1

CASE 3: LOW-ORDER OUTPUT-FEEDBACK EMPC

Maximum Temperature and Manipulated Input Trajectories

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−1.5

−1

−0.5

0

0.5

1

1.5

t

u(t)

11 Slow Modes21 Slow ModesUniform Input, u(t) = 0.5

Manipulated Input Trajectory

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 11.4

1.6

1.8

2

2.2

2.4

2.6

2.8

3

3.2

t

x 1,m

ax

21 Slow Modes11 Slow ModesMaximum Allowable Temperature

Maximum Temperature Trajectory

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

1

2

3

4

5

6

t

Rea

ctio

n R

ate,

L(x

,u)

11 Slow Modes21 Slow ModesUniform Input, u(t) = 0.5

Reaction Rate Trajectory

• EMPC based on 21 slow modes computes

a smoother manipulated input profile

• Average reaction rate from the EMPC

based on 21 slow modes is 5.13% greater

than that of EMPC based on 11 slow

modes and 8.45% greater than that of the

system with uniform in time reactant ma-

terial distribution

EMPC FORMULATION 4High-order Output-feedback EMPC Formulation

• EMPC formulation:

maxu∈S(∆)

∫ ttk+N

tk

L(as(τ), af (τ), u(τ))dτ

s.t. ˙as(t) = Asas(t) + Fs(as(t), af (t)) + Bsu(t)

as(tk) = C−1s y(tk)

af (t) = −A−1f [Bfu(t) + ff (as(t), 0)]

a(t) = as(t) + af (t)

umin ≤ u(t) ≤ umax

u ∈ gk

amin ≤ a(t) ≤ amax

[a(t)− as]TP [a(t)− as] < ρ, ∀ t ∈ [tk, tk+N )

⋄ Model based on first 30 modes which include 11 slow modes and 19 fast modes

(m = 11, l = 19) computed from 11 measured state-feedback points (p = 11)

⋄ af (t): predicted fast mode function of af (t)

⋄ a(t): predicted finite mode function of a(t)

CASE 4: HIGH-ORDER OUTPUT-FEEDBACK EMPC

Maximum Dimensionless Temperature and Manipulated Input Trajectories

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1−1.2

−0.8

−0.4

0

0.4

0.8

1.2

t

u(t)

30 Modes (11 Slow Modes, Output Feedback)

11 Modes (11 Slow Modes, Output Feedback)

30 Modes (11 Slow Modes, State Feedback)

Uniform Input, u(t) = 0.5

Manipulated Input Trajectory

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

1.6

1.8

2

2.2

2.4

2.6

2.8

3

3.2

t

x 1,m

ax

30 Modes (11 Slow Modes, Output Feedback)11 Modes (11 Slow Modes, Output Feedback)30 Modes (11 Slow Modes, State Feedback)Maximum Allowable Temperature

Maximum Temperature Trajectory

• Additional 19 fast modes reduces the input fluctuation and increase the maximum

dimensionless temperature

• State-feedback EMPC with same number of slow and fast modes uses 100

measured points to construct the reduced-order model used in EMPC

• State-feedback EMPC shows its advantage of model accuracy due to more

available measurement points

CASE 4: HIGH-ORDER OUTPUT-FEEDBACK EMPC

Reaction Rate and Computational Time Trajectories

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

1

2

3

4

5

6

t

Rea

ctio

n R

ate,

L(x

,u)

30 Modes (11 Slow Modes, Output Feedback)

11 Modes (11 Slow Modes, Output Feedback)

30 Modes (11 Slow Modes, State Feedback)

Uniform Input, u(t) = 0.5

Reaction Rate Trajectory

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

10

20

30

40

50

60

70

t

Com

puta

tiona

l Tim

e Pr

ofile

/ se

c

30 Modes (11 Slow Modes, Output Feedback)11 Modes (11 Slow Modes, Output Feedback)30 Modes (11 Slow Modes, State Feedback)

EMPC Computational Time Trajectory

• Average reaction rate from the high-order output feedback EMPC is 0.89%

greater than that from the low-order output feedback EMPC system

• Average reaction rate from the high-order output feedback EMPC system is

0.73% less than that from the EMPC system with state feedback

• The average computational efficiency of the high-order output feedback EMPC

system is comparable to that of the state feedback EMPC system

CONCLUSIONS

• Proposed EMPC formulation for nonlinear parabolic PDE systems

⋄ Use Galerkin’s method to realize order reduction of nonlinear parabolic PDE

model

⋄ Perform low/high-order state/output-feedback EMPC formulations

• Application to a tubular reactor

⋄ Demonstrated the ability to maximize the economic cost by following a

time-varying input trajectory

⋄ Demonstrated the ability to satisfy a state constraint on the maximum

allowable temperature

⋄ Compared the output-feedback and state-feedback EMPC performance on

model accuracy, objective optimization and computational efficiency

ACKNOWLEDGEMENT

Financial support from the National Science Foundation and Department of Energy is

gratefully acknowledged