You are on page 1of 9

This article has been accepted for inclusion in a future issue of this journal.

Content is final as presented, with the exception of pagination.


IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY

Model Predictive Control for Luminous Flux Tracking


in Light-Emitting Diodes
Silvio Baccari, Member, IEEE, Francesco Vasca, Senior Member, IEEE,
Massimo Tipaldi, and Luigi Iannelli, Senior Member, IEEE

Abstract Luminous flux tracking and junction temperature


stress minimization are typical objectives in the regulation
of high-brightness light-emitting diodes (LEDs). In this brief,
a solution based on a model predictive controller approach
is proposed. The state estimation is obtained using an LED
photoelectrothermal dynamic model whose parameters are tuned
through datasheets and dedicated experiments. By exploiting the
dynamic separation within the thermal state variables of the
model, the junction temperature is predicted using LED current
and heat sink temperature measurements. The effectiveness
of the proposed approach is verified through simulations and
experiments on different heat sinks and samples of the same
LED family, thus confirming the achievement of the luminous
flux tracking under current and temperature constraints typically
featuring LED applications.
Index Terms DCDC buck converter, light-emitting
diodes (LEDs), luminous flux tracking, model predictive control,
photoelectrothermal dynamic model.

I. I NTRODUCTION

IGHT-EMITTING DIODES (LEDs) represent a


promising solution for a wide spread replacement of the
traditional lighting components with more efficient and flexible
devices [1][4]. Different topologies and strategies have been
proposed in the literature for the LED drivers design
(see [5][7] and the references therein). Concisely one
could say that there exist two approaches for defining the
current that the LED driver should provide [8]: the LED
current is controlled such that its average value provides the
desired luminous flux [9][11]; the instantaneous current
produces an instantaneous illumination that is averaged by
the human eye filtering capabilities thus providing the desired
illumination [5], [9]. In both cases, the driver control problem
consists in regulating the LED current to a reference trajectory.
Nonetheless, it is not a trivial task to determine the LED driver
reference signal able to provide the desired luminous flux by
also taking into account the electrical and thermal constraints.
Regardless of the current regulation strategy, the reference
current is usually obtained using a static map relating the luminous flux to the LED current. Yet the nonlinear dependence of

Manuscript received September 24, 2015; revised January 19, 2016;


accepted April 8, 2016. Manuscript received in final form April 26, 2016. This
work was supported by the Ministero dellIstruzione, dellUniversit e della
Ricerca under Grant SFERE PON 01_00595. Recommended by Associate
Editor S. M. Savaresi.
The authors are with the Department of Engineering, University
of Sannio, Benevento 82100, Italy (e-mail: silvio.baccari@unisannio.it;
vasca@unisannio.it; mtipaldi@unisannio.it; luigi.iannelli@unisannio.it).
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TCST.2016.2560122

the luminous efficacy on the device forward current presents a


typical degradation for large currents [12], [13]. Moreover, the
temperature also affects such a nonlinear characteristic, and a
proper static and dynamic thermal management is a key factor
for counteracting the undesired photometric variations [14].
The construction of a model that is able to capture and
combine the luminous, the electrical, and the thermal phenomena, i.e., a photoelectrothermal model, is a crucial preliminary
step for the LED reference current design [15], [16]. Static
models are analyzed, among others, in [13], [17], and [18],
while dynamics are taken into account in [19] and [20].
The photoelectrothermal model here proposed is based on a
1-D thermal flow from the LED junction to the environment
using equivalent thermal resistors and capacitors [21], [22].
The dynamic separation among the thermal parts of the model
is exploited in order to define a reduced-order model suitable
for the controller design and implementation. Moving on from
such models, the typical static characteristics can thus be
interpreted as the steady-state behavior resulting from dynamic
photoelectrothermal models [15], [19].
The proposed controller is characterized by two nested
loops: the inner loop consists of a current-controlled dcdc
buck power converter [5], while the outer loop provides the
reference current to be tracked by the inner loop. In particular,
the outer loop consists of a model predictive controller (MPC)
aimed at regulating the LED luminosity by solving an optimization problem formulated on the reduced-order dynamic
model. By exploiting the typical MPC features [23], [24], the
proposed controller provides a unifying multivariable system
framework where objective functions based on the typical LED
performances, e.g., luminosity, heating, and electric efficiency,
can be integrated and tuned according to the specific goals.
In the presence of flux measurements, it is possible to include
the integral of the LED luminous flux error in the cost
function and the proposed MPC can be interpreted as an extension of the most common sensor-based approaches with flux
feedback and temperature feedforward compensations [25].
The proposed solution can be customized according to the
design requirements, guarantees a wide operating range, and
thanks to the abovementioned extension, can counteract the
LED aging process (usually faster than the one of the light
sensors). Moreover, it allows dealing with constraints that
can be handled during the control design phase, e.g., current
peak and maximum temperature. Finally, the computation of
the control signal is based on an optimization procedure,
which takes into account the desired reference trajectories for
luminosity and temperature.

1063-6536 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
2

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY

The rest of this brief is organized as follows. In Section II,


the photoelectrothermal LED model is derived. A procedure
for the determination of the model parameters is presented
in Section III. The MPC is described in Section IV. The
numerical and experimental results are given in Section V
by considering different heat sinks and LED samples and
by showing the better performances obtained by the MPC
with respect to those achievable with a standard controller.
Section VI concludes this brief.
II. P HOTOELECTROTHERMAL M ODEL
It is well known that the LED current, say i f , gradually
increases with the forward voltage, say v f . This empirical
observation gives the idea of modeling an LED as a diode
connected to an equivalent series resistor, say Rd , resulting
from the resistance of the p-n junction, contacts, and current
spreading layer [26]. Thus, the LED voltagecurrent characteristic can be modeled as
 



if
Tj
nk B T0
1 ln 1 +
1 kv
v f (i f , T j ) =
q
T0
I0
+Rd i f
(1)
where I0 is the reverse saturation current, T j is the junction
temperature, k B is the Boltzmann constant, q is the electric
charge, n is the ideality factor, T0 is a nominal temperature,
and kv is a nonnegative constant used to represent the shift
of the voltagecurrent characteristic for different values of the
junction temperature [20], [27]. The thermal dependence of the
characteristic is concentrated into the linear term depending
on kv , and the saturation current is assumed to be constant.
The electric power supplied to the LED can be expressed as
Pel (i f , T j ) = v f (i f , T j )i f

(2)

where the characteristic v f (i f , T j ) is given by (1). A fraction


of the electrical power supplied to the LED is transformed into
the optical power. By assuming that the optical power Popt and
the electrical power Pel are measurable, one can define the
so-called wall-plug efficiency [18]
Popt
.
(3)
Pel
Definition (3) can be interpreted as an external efficiency
because it is based on quantities that are directly measurable
without considering any specific internal phenomena.
The power balance can be modeled by assuming that the
total electric power determines the lighting effects and heating.
Then, using (3), one can write
p 

Pel = Popt + Pheat = p Pel + (1 p )Pel .

(4)

The efficiency p depends on both the junction temperature


and the LED current. In the following, we assume that the
wall-plug efficiency is expressed by means of the bilinear
function:
 




if
Tj
1 k
1
(5)
p (i f , T j ) = p 1 ki 1
I
T
where p is the efficiency at the nominal temperature T j = T
and nominal current i f = I . The nonnegative constants

Fig. 1.

Heat transfer equivalent scheme for a typical LED implementation.

ki and k are such that 0 < p (i f , T j ) < 1 for any


i f and T j of interest. The bilinear model (5) is similar to
the expression used in [13] and [28]. It can be deduced from
the characteristics reported in [18] and can be interpreted
as a simplified version of the model adopted in [29]. The
specific expression of the efficiency p is not crucial for the
application of the methodology proposed in this brief, which
can be applied for any reasonable nonlinear model of the wallplug efficiency. Expression (5) is validated in Section III by
considering the LED characteristic of luminous flux versus
current for different temperatures, and it turns out to be in line
with flux measurements through the calibration of the sensor
used in the feedback as shown in Section V-A.
In order to represent the dynamic thermal model, the following subscripts are defined: j for the junction, c for the capsule,
s for the heat slug, h for the heat sink, and a for the ambient.
Denote by T the temperatures, assumed to be uniform in
each part, m the equivalent masses, and c the heat capacities.
The symbols R indicate the equivalent thermal resistances and
C = mc the corresponding thermal capacitances (see Fig. 1).
By considering the (electrical) series connection of
Nd LEDs that share only the common heat sink, the following
dynamic model for the heat transfer can be derived using the
energy conservation principle:
dTj
1
= (1 p (i f , T j ))Pel (i f , T j )
(T j Tc )
dt
R jc
1

(T j Ts )
(6a)
R js
1
d Tc
1
=
m c cc
(T j Tc ) +
(Ts Tc )
(6b)
dt
R jc
Rsc
1
d Ts
1
1
m s cs
=
(T j Ts )
(Ts Tc )
(Ts Th )
dt
R js
Rsc
Rsh
(6c)
Nd
d Th
1
=
m h ch
(Ts Th )
(Th Ta ).
(6d)
dt
Rsh
Rha
m jcj

Model (6) is linear with respect to Pheat [see (4)] and


nonlinear when the electrical current i f is considered as an
input. In particular, (6a) is nonlinear due to the dependence
of the wall-plug efficiency p and electric power Pel on
the junction temperature. Moreover, the LED current i f is
considered as an independent variable because it is directly
controlled by the power converter that drives the LED.
The luminous flux v is clearly an output of interest for
the model. To show how v is related to the thermal and
electrical variables of the model, the luminous efficiency E
of a light source is introduced and expressed as the product
of the luminous efficacy of the optical radiation, Er , and the

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
BACCARI et al.: MODEL PREDICTIVE CONTROL FOR LUMINOUS FLUX TRACKING IN LEDs

TABLE I

power efficiency [16]


E=

v
v Popt
=
= Er p .
Pel
Popt Pel

(7)

PARAMETERS AND C ONSTANTS OF THE P HOTOELECTROTHERMAL


M ODEL AND C ONTROLLER W ITH THE I NDICATION
OF THE S OURCE FOR T HEIR VALUES

By considering a series connection of Nd LEDs and


using (7), one can write
v (i f , T j ) = Nd Er p (i f , T j )Pel (i f , T j )

(8)

where the wall-plug efficiency p is given by (5) and the


electrical power can be expressed as (2).
The computational load capabilities of the typical commercial digital controllers for LEDs make model (6) too complex
to be adopted for a real implementation. In order to overcome
such an issue, a reduced-order thermal model can be obtained
from (6) by assuming
m c  m h ch

(9)

for any { j, c, s}, i.e., the diode thermal capacities are


negligible in comparison with the heat sink thermal capacity.
Assumption (9) will be justified in Section III by exploiting
the model parameters tuning procedure. At a slow time scale,
one can assume that the small parameters m c are zero,
thus obtaining from (6a)(6c) the corresponding equations at
steady state. By solving such equations, one can write the
corresponding steady-state temperatures, say Tc and Ts , as
linearly dependent on the steady-state temperature T j and the
instantaneous temperature Th , thus obtaining
0 = 1 (1 p (i f , T j ))Pel (i f , T j ) + 2 T j + 3 Th

(10)

where 1 , 2 , and 3 are derived from (6a)(6c) at steady state


Rsh
Rsh
+
R j c + Rsc
R js
1
Rsc
1

2 =

R j c (R j c + Rsc )
R jc
R js
1
1
3 =
+
.
R j c + Rsc
R js
1 = 1 +

(11a)
(11b)
(11c)

By considering (5) and (2) with (1), it is simple to deduce


that (10) is quadratic in T j . For each pair (i f , Th ) within
realistic ranges, (10) has two real solutions of different signs.
Then, by considering the positive solution, one can conclude
that (10) represents a unique mapping from (i f , Th ) to T j .
By adding (6a)(6c) at steady-state, the dynamic equation (6d) can be rewritten as
m h ch

Model (12) with (10) and (5) will be used for the controller
design and for its implementation in the experimental setup.
The more complex model (6) will be used as the actual process
model for the numerical validation.

d Th
1
= Nd (1 p (i f , T j ))Pel (i f , T j )
(Th Ta )
dt
Rha
(12)

where T j can be analytically determined as a function of


Th and i f by taking the positive solution of the quadratic
equation (10). Model (12) corresponds to an equivalent electrical circuit that can be obtained from the one shown in Fig. 1
by solving it after assuming C j , Cc , and Cs to be zero,
i.e., the corresponding capacitors replaced by open circuits.
In particular, the heating power source Pheat is given by the
state-dependent function expressed by the first term in the
right-hand side of (12).

III. M ODEL PARAMETERS S ELECTION


The parameters of the model are selected starting from
typical information reported in the LED datasheets and based
on step-response experiments. The parameters are referred to
the datasheet of the LED CREE family Xlamp Xp-G (product code XPGWHT-L1-0000-00CE7). Table I reports all the
parameters and physical constants (and their values) adopted
in this brief. It also highlights whether they are available or
they need to be identified.
We now describe a procedure for the parameters determination of the complete model (6) and of the reduced-order
model (12). Note that only the latter one has to be identified for
the controller implementation. As regards the model (1) of the
voltagecurrent LED characteristic, the nominal temperature
is T0 , while k B and q are physical constants. The parameters
of (1) to be determined are n, I0 , Rd , and kv . The ideality
factor n, the current I0 , and the equivalent resistance Rd are
estimated through a linear regression approach applied using
model (1) and the LED data at the nominal temperature [note
that for T j = T0 the characteristic (1) does not depend on kv ].
The parameter kv can be computed from the information

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
4

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY

Fig. 3. Block scheme of the closed loop system. The outer feedback loop
provides the reference current for an inner and faster LED current control loop
that provides the duty cycle d to the power converter. i f m is the measured
LED current and v is the measured LED flux.

Fig. 2. Luminous flux v versus current i f for different junction temperatures. The stars indicate values from datasheet at T j = T = 298.1 K. Other
curves represent (8) for different values of T j . The filled circles (with the
corresponding interpolating curve) indicate the luminous flux v calculated
with (8) for different currents and T j obtained from (6) at steady state.

usually available from the LED datasheets, wherein the values


of the forward-biased voltage at T0 and T j = T0 + T j are
provided at a specific current test i f . By exploiting (1), it is
straightforward to show that
kv =

qv f

nk B T j ln 1 +

i f
I0

(13)

where v f = v f (i f , T j ) v f (i f , T0 ). The parameters


identification for (8) must be completed by selecting the
parameters p , ki , and k of the efficiency expression (5). The
values of the luminous flux v for different LED currents at
the nominal temperature T can be deduced from the LED
datasheets (see the stars in Fig. 2). Therefore, the nominal
luminous efficiency p is calculated by exploiting (5) and (8)
with i f = I and T j = T . The parameter ki is determined
using a least square (LS) algorithm to fit model (8) with
the LED datasheet values at T j = T . The parameter k is
evaluated from the LED datasheets, that is, using data on
the luminous flux gradient versus the junction temperature T j ,
at the nominal current i f = I .
The luminous flux characteristic curves in Fig. 2, which
present the typical concave shape discussed in [12], allow
making considerations suitable for the luminous flux control.
First of all, it should be noted that, by assuming a maximum
current equal to 1.5 A, at steady state, it is not possible
to provide a luminous flux larger than approximatively
238 lm, i.e., the maximum of the curve interpolating the filled
circles. On the other hand, larger fluxes are possible during
transients. For instance, starting from the steady-state point
(i f , T j ) = (0.5 A, 335.4 K), which corresponds to a luminous
flux around 130 lm, and by providing a current step till 1.5 A,
since the current dynamics are much faster than the thermal
ones, the operating point will approximatively move on the
characteristic at a constant temperature equal to T j = 335.4 K
by providing a luminous flux equal to 300 lm. Then the
operating point will change at a constant current equal
to 1.5 A, and the increasing of the junction temperature will
reduce the luminous flux till the steady-state point is reached,
i.e., i f = 1.5 A, T j = 410 K, and v = 238 lm. The best

transient path from a given steady-state operating point to


another one is not obvious, and to this aim, the MPC will
show its potentialities.
In order to complete the model parameters identification,
the parameters of the dynamic equations must be determined.
By taking into account the separation among the thermal
dynamics, the most important thermal parameters are those
characterizing the heat sink dynamics represented by the
reduced-order thermal model (12). By considering the experimental setup described in Section V and by forcing the LED
with a step variation of its current, the values of the parameters
m h ch and Rha are selected so as to provide a good fitting
between the evolution predicted by model (12) and the experimental data. The mass m h is available from the datasheet.
Finally, the values m j c j , m c cc , and m s cs , and the thermal
resistances R j s , Rsh , R j c , and Rsc are chosen coherently with
similar models presented in [19], [21], and [30]. The identified
values of the parameters confirm the validity of (9).
IV. M ODEL P REDICTIVE C ONTROLLER
A block scheme of the proposed controller is shown
in Fig. 3. Two nested loops are used. In the outer and slower
loop, the LED current i f is considered as the control variable
for an optimization problem constrained by the dynamics (12)
together with the relation (10), which provide the prediction
of the evolutions of the thermal state variables. The current
computed by the MPC, say i f , is the reference signal for an
inner and faster control loop on the measured LED current.
The proposed MPC is obtained by solving a receding horizon optimization problem constrained by a discretized version
of (10)(12) and an integral controller of the LED luminous
flux error, the latter in case of luminous flux measurements.
The continuous dynamics (12) are discretized with a
sampling period h s , and the constraints are active for each
sampling time instant kh s with k being a nonnegative
integer. It is assumed that the controller requires a sampling
period h c = Ns h s > h s , with Ns being an integer, in
order to compute a good suboptimal solution; without loss
of generality, assume that the prediction horizon is given
by Nc h c , where Nc is a positive integer. In other words, the
controller provides a sample of the control variable i f each
Ns integration cycles of the model, and the prediction horizon
is an Nc multiple of the controller sampling period h c .
By discretizing (12) with a forward technique and using (10),
the following expression can be obtained:
Th (k + 1) = f d (Th (k), i f (k), Ta )

(14)

where the state is the heat sink temperature whose samples


are denoted by Th (k) = Th (kh s ).

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
BACCARI et al.: MODEL PREDICTIVE CONTROL FOR LUMINOUS FLUX TRACKING IN LEDs

A further equation is included in the optimization


problem, which implements a controller on the integral
of the flux error. In particular, the following equation is
considered:


x e (k + 1) = x e (k) + h s ref
v (k) v (k)

(15)

where v is the measured luminous flux and x e is the new


state variable corresponding to the cumulative flux error. The
MPC can be formulated through the following optimization
problem:

k
k0 +N1
k0 +N+1
0 +N


J1 (k) +
J2 (k)+
J3 (k)
min
i f ()

k=k0

k=k0 +1

k=k0

s.t.: model (14) (15) with (10)

rewritten as

k
k0 +N1
k0 +N+1
0 +N


J1 (k) +
J2 (k)+
J3 (k) + J (k)
min
i f ()

k=k0

k=k0

k=k0 +1

s.t.: model (14) (15) with (10)


i f (k0 + + j Ns ) = i f (k0 + j Ns )
= 1, . . . , Ns 1,

j = 0, 1, . . . , Nc 1

i f ()

(19)

which can be more easily implemented on a digital platform


equipped with standard algorithms for multivariable function
optimization.

i f (k0 + + j Ns ) = i f (k0 + j Ns )
j = 0, 1, . . . , Nc 1

(18)

with J (k) = i f i f (i f (k0 ), i f (k0 + Ns ), . . . , i f (k0 + N)) +


T j T j (T j (k0 ), T j (k0 + 1), . . . , T j (k0 + N)), where
i f and T j are the weights of the barrier functions
i f and T j . By substituting recursively (8), (14), and (15)
in (18), problem (18) can be rewritten in the following
unconstrained form:
min 
i f (i f (k0 ), i f (k0 + Ns ), . . . , i f (k0 + N))

i min
i f (k) i max
f
f
T jmin T j (k) T jmax
= 1, . . . , Ns 1,

(16)

for k = k0 , . . . , k0 + N, N = Nc Ns , and T j (k) is given by (10)


with Th and i f replaced by Th (k) and i f (k), respectively. The
cost functions are given by

V. E XPERIMENTS
We have verified the effectiveness of the proposed MPC,
implemented using (19), through numerical and experimental
comparisons with a closed-loop scheme similar to that in Fig. 3
where the MPC is replaced by a feedforward controller.

2

J1 (k) = qT j T j (k) T jref (k)
2

+qv v (k) ref
v (k)

(17a)

A. Numerical and Experimental Setups

J2 (k) = qi f (i f (k) i f (k Ns ))

(17b)

In the simulations, the LED thermal model (6) together


with the averaged model of the dcdc buck power converter
is used as the real process to be controlled, and it will
be indicated as the photoelectrothermal model. The photoelectrothermal model with the MPC has been implemented
on the MATLAB/Simulink platform with four time domains,
each featured by a fixed integration time step. The first
(and smallest) step size h sim is chosen for the simulation of
the photoelectrothermal model. The second sampling period
h PID > h sim is dedicated to the PI current controllerthat
calculates the duty cycle on the basis of the error between
the set point i f provided by the MPC and the measured
LED current. The third time step h s > h PID is dedicated to
the simulation of the thermal part of LED lamp inside the
MPC optimization problem, while the last sampling period
h c > h s refers to the MPC. The PI controller sampling period
is h PID = 0.1 ms. The step size for the model integration
used by the MPC is h s = 1 ms, the MPC sampling period is
h c = 15 s, and the length of the prediction horizon is Nc = 10.
The PI parameters are 0.045 and 13.5 for the proportional and
integral gains, respectively.
The experimental setup aims at verifying the performance
of the proposed MPC. The setup follows the conceptual block
scheme reported in Fig. 3. The inner control loop is based
on a PI current regulator that drives the power MOSFET of a
dcdc buck converter [36]. This regulator is implemented on
a 32-bit microcontroller PIC32mx795f512l micro-controller
unit, provided by Microchip. It receives the reference current
(set every h c seconds by the MPC) and provides the commands

J3 (k) = qxe x e (k)2

(17c)

where the superscripts min, max, and ref are used for indicating minimum, maximum, and reference signals, respectively,
of the corresponding variables, and qT j , qv , qi f , and q xe are
weighting parameters of the quadratic cost functions. In the
last sum of (16), by considering (17c) and (15), the luminous
flux v (k0 ) is available from the last measurement, while
the remaining values v (k) for k = k0 + 1, . . . , k0 + N are
predicted using (8) with T j (k) given by the thermal model and
i f (k) being the optimization variable. The equality constraints
in the optimization problem (16) ensure that the controller
output i f is piecewise constant with a period equal to h c .
The term J2 permits to reduce the variation of the control
signal i f in the prediction horizon. The term J3 is useful
when flux measurements are available, and it introduces in the
MPC formulation the action typical of a classical proportional
integral (PI) controller. Indeed, using qxe = 0 and thanks
to (15), the past and present flux errors are also weighted
in the minimization problem. Instead, the LED luminous flux
measurements do not affect the optimization problem solution
when qxe = 0, because in this case, the objective function
depends only on model predictions.
The optimization problem (16) can be reformulated in order
to include the temperature and current inequality constraints in
the cost function. To this aim, a barrier function approach [35]
is used, which prescribes a high cost for violation of
the constraints. The minimization problem (16) can be

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
6

to the switch of the converter. The current measurement is


obtained by a series resistor.
The MPC minimization problem (19) is implemented by
means of the Real Time Windows Target toolbox and runs
on a digital control platform based on Intel Core i7 processor
at 3.4 GHz. The reference current generated by the MPC is
passed to the microcontroller by a serial port interface and
in particular by the use of 4 B of data that are combined to
obtain a 10-bit reference. The temperature measurements are
forwarded to the MPC by means of the microcontroller board
connected through the serial RS232. The heat sink temperature
is measured by means of a sensor TMP05, which generates a
modulated serial digital output that varies proportionally to the
measured temperature. The LED luminous flux is measured by
means of the light-to-frequency transducer TSL230. The light
sensor is placed in correspondence of the maximum emitting
flux direction, according to the geometrical distributions of the
luminous intensity provided by the LED and sensor datasheet.
Dedicated flux measurement tests are carried out for the sensor
calibration. An open-loop test is done at a given current. The
corresponding temperature is measured and the flux provided
by (8) is obtained. Then the sensor is calibrated, i.e., a
proportional gain is determined and used as normalization
factor during the experiments in order to make coherent the
measured LED luminosity intensity with its flux values as
provided by the datasheet at those operating conditions. The
proposed implementation allows a conceptual validation of the
approach, while an industrial/commercial realization should
be achieved by embedded platforms and by calibration that
considers the spectral responses of LED and sensor as well.

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY

Fig. 4.
Luminous fluxes obtained with the MPC and the feedforward
controller in a luminous flux regulation test with ref
v = 200 lm.

Fig. 5.

Heat sink temperatures for the same test of Fig. 4.

B. Luminous Flux Regulation


Let us consider a luminous flux regulation test case. The
parameters of the cost function are qTh = 0, qv = 100,
qi f = 0, qxe = 0, and i f = Th = 104 . The maximum
current is (also for the feedforward controllers) i max
= 1 A.
f
The real-time solution of the optimization problem is realized
with a MATLAB code that uses the fmincon method with the
not scaled sequential quadratic programming algorithm and the
following parameters: termination tolerances for the objective
function TolFun = 105 and for the state TolX = 107 ,
constraints tolerance TolCon = 105 , maximum number of
function evaluations MaxFunEvals = 100, maximum number
of iterations MaxIter = 10, and lower bound of the objective
limit ObjectiveLimit = 103 . The cost function (19) is
implemented using a logarithmic barrier function.
Figs. 4 and 5 show the experimental and simulation results
(luminous flux and heat sink temperature, respectively) for
the case of a luminous flux regulation to 200 lm obtained via
the MPC and the feedforward controller. In accordance to
what described above, the feedforward controller provides a
reference current based on the nominal fluxcurrent map, and
therefore the corresponding current is constant. Figs. 4 and 5
highlight a satisfactory consistency between experiments
and numerical results as well as the coherence of the
experimental steady state with the maps presented in Fig. 2.
The steady-state flux difference between the MPC and the

Fig. 6. Experimental and simulation results for the MPC in luminous flux
regulation test with ref
v = 220 lm.

feedforward controller is due to the fact that the latter has no


compensation of the measured temperatures. In order to show
the performance of the MPC when the constraints are active,
an experiment with the flux reference set to 220 lm has been
carried out. The corresponding simulation and experimental
results are reported in Figs. 68.
The quantitative (minor) differences between the simulations and experiments are due to the maximum number of
iterations fixed for the experiments. The luminous flux is
maintained at the maximum value compatible with the current
and temperature constraints (see Fig. 6). The LED current
evolution is due to the increase in the heat sink temperature
(see Fig. 7), and it saturates at i max
f . At around 800 s, the
current starts decreasing because the junction temperature

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
BACCARI et al.: MODEL PREDICTIVE CONTROL FOR LUMINOUS FLUX TRACKING IN LEDs

TABLE II
C OMPARISON OF THE F EEDFORWARD (FF) C ONTROLLER AND MPCs
W ITH D IFFERENT C OST F UNCTIONS : MPC0 I S W ITH L UMINOUS F LUX
T RACKING AND qTh = 0; MPC1 I S W ITH qv = 0 AND
T EMPERATURE T RACKING ; AND MPC2 I S W ITH
T EMPERATURE AND L UMINOUS F LUX
T RACKING S IGNALS

Fig. 7.

LED currents for the same test of Fig. 6.

Fig. 8.

Junction and heat sink temperatures for the same test of Fig. 6.

becomes close to its maximum value that for this experiments


has been set to 355 K (see Fig. 8). The (low amplitude) spikes
of the simulated current are due to the very small difference
between the simulated junction temperature and its maximum
value.
C. Different Reference Signals
Table II shows a comparison of the feedforward controller
and MPCs for different cost functions. The reference current
of the feedforward controller is obtained via the steady-state
map reported in Fig. 2 (filled circles with the corresponding
interpolating curve) using as entry the reference flux. Each
column corresponds to a different value of the reference
flux ref
v . The variables reported are the steady-state junction
ss with respect
temperature T jss , the flux steady-state error e
to the reference flux, the steady-state wall-plug efficiency ss
p,
the average value int over the simulation time interval of the
difference between the instantaneous and the steady-state
wall-plug efficiency (the highest the better), and the average
int over the simulation time interval of the flux error.
value e
Inspired by Fig. 8, the reference signal for the junction
temperature, when used, is a piecewise linear signal starting
from the initial value of the temperature with two breaking
timetemperature points at (500 s, 0.9T jss ) and (1500 s, T jss ).
The MPC0 performs always better than the feedforward
one. The results of MPC1 show the importance of considering
the flux error in the cost function. By comparing the values
of int corresponding to MPC2 and MPC0 at ref
v = 150 lm,

Fig. 9. Measured luminous flux (continuous line) with time-varying reference


flux (dotted line) obtained by adding a constant (125 lm) to a sinusoid (period
equal to 36 min and amplitude equal to 25 lm). The dashed and dashed-dotted
lines indicate the estimated junction temperature and the measured heat sink
temperature, respectively.

it is evident that the use of a reference temperature signal


in the cost function can provide better performance when
the luminous flux reference value corresponds to intervals
where the nonlinearity of the fluxcurrent characteristic is
more evident (see Fig. 2).
The effectiveness of the MPC nonconstant flux reference
signals is confirmed by the experimental results in Fig. 9.
D. Different LEDs and Heat Sinks
In order to analyze the robustness of the proposed MPC,
a luminous flux regulation test has been repeated for three
different setups. The setup indicated by exp1 is the nominal
plant used for the parameters tuning, while exp2 and exp3 are
featured by different LED pads and a different type of heat
sink (exp3). In all three setups, the same MPC designed for the

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
8

Fig. 10. LEDs currents and luminous fluxes for the three experimental
setups and for a flux regulation test with ref
v = 155 lm. For flux, continuous
line represents the nominal case. Different values of the current are due to
different values of the junction temperatures.

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY

Fig. 12.
Evolution of xe (k) and LED luminous flux measurement in
the luminous flux feedback experiments with qxe = 1 and qxe = 0 for
ref (k) = 210 lm. qTh = 0. qv = 100. qi f = 0. i f = Th = 104 .

microcontroller can retrieve the LED reference forward current


based on the measured Th and the desired luminous flux value.
The map in Fig. 11 provides results that are coherent with the
experimental data in Fig. 10.
F. LED Luminous Flux Feedback

Fig. 11. The current i f (k0 ) (z-axis) that solves problem (16) with constant
reference luminous flux values (y-axis) and heat sink temperature (x-axis).

The last experiment verifies the MPC performance in the


presence of luminous flux measurements, i.e., with a nonzero
term J3 in (16). The experimental results in Fig. 12 show that,
after a transient, a zero steady-state error on the luminous flux
is obtained, which is also confirmed by the constant steadystate value of the state variable x e .
VI. C ONCLUSION

nominal case has been adopted. The results of the experiments


are shown in Fig. 10. Good reference flux regulations are
also obtained in the nonnominal scenario. The LED currents
show how the MPC can compensate for the LED efficiency
reduction due to the different evolutions of the heat sink
temperature, independently from the ambient temperature.
E. Lookup Table Implementation
The solution of the minimization problem (19) can be
approximated by discretizing the flux and the heat sink temperature in certain intervals. In Fig. 11, the minimization
problem has been uniformly discretized with a 30-step grid
over the interval [262.15 K, 379.9 K] for Th and [10 lm,
240 lm] for ref
v . The offline calculated values are stored into a
2-D array and can be accessed by a simple indexing operation.
For a microcontroller, retrieving a value from memory is
faster than the online minimization procedure that can be
executed on a powerful PC. The table is precalculated with
a better precision than the online computation because in the
offline optimization, the maximum execution time constraints
are not used. The table can be either implemented in the
source code or stored in static memory support/Application
Specific Integrated Circuit hardware according to the specific
application needs. The precomputed values of i f correspond
to the first sample of the current in the control horizon with
a constant reference flux. During the operating phase, the

An MPC for LEDs modeled by a set of equations capturing


their electric, thermal, and optical behaviors has been presented. The proposed approach offers a comprehensive control
framework, where different and conflicting objectives are
handled simultaneously and extends the most common sensorbased approaches based on flux feedback and temperature
feedforward compensation. In particular, the control strategy
is capable of optimizing performance related to luminosity
by satisfying the typical constraints on temperature and
current. The simulation and experimental results demonstrate
the effectiveness of the proposed control strategy that
allows obtaining high wall-plug efficiencies for luminous flux
regulation problems also during transients when the constraints
on critical process variables become active. A lookup table
controller implementation has shown its practicality also for
digital control platforms whose computational capabilities do
not allow the online solution of the optimization problem.
R EFERENCES
[1] B. Bowers, Lights out, IEEE Spectr., vol. 48, no. 4, pp. 4452,
Apr. 2011.
[2] D. Tran and Y. K. Tan, Sensorless illumination control of a networked
LED-lighting system using feedforward neural network, IEEE Trans.
Ind. Electron., vol. 61, no. 4, pp. 21132121, Apr. 2014.
[3] M. G. Craford, R. D. Dupuis, M. Feng, F. A. Kish, and J. Laskar,
50th anniversary of the light-emitting diode (LED): An ultimate lamp
[scanning the issue], Proc. IEEE, vol. 101, no. 10, pp. 21542157,
Oct. 2013.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
BACCARI et al.: MODEL PREDICTIVE CONTROL FOR LUMINOUS FLUX TRACKING IN LEDs

[4] S. Scalzi, S. Bifaretti, and C. M. Verrelli, Repetitive learning control


design for LED light tracking, IEEE Trans. Control Syst. Technol.,
vol. 23, no. 3, pp. 11391146, May 2015.
[5] S.-C. Tan, General n-level driving approach for improving electrical-tooptical energy-conversion efficiency of fast-response saturable lighting
devices, IEEE Trans. Ind. Electron., vol. 57, no. 4, pp. 13421353,
Apr. 2010.
[6] F. J. Azcondo, J. M. Alonso, and S. Y. R. Hui, Special section on
modern ballast technology and lighting applications, IEEE Trans. Ind.
Electron., vol. 59, no. 4, pp. 16891759, Apr. 2012.
[7] M.-S. Lin and C.-L. Chen, An integrated lighting unit with regulated
pulse current driving technique, IEEE Trans. Ind. Electron., vol. 60,
no. 10, pp. 46944701, Oct. 2013.
[8] Y.-L. Lin, H.-J. Chiu, Y.-K. Lo, and C.-M. Leng, LED backlight driver
circuit with dual-mode dimming control and current-balancing design,
IEEE Trans. Ind. Electron., vol. 61, no. 9, pp. 46324639, Sep. 2014.
[9] X. Lv, K. H. Loo, Y. M. Lai, and C. K. Tse, Energy-saving driver design
for full-color large-area LED display panel systems, IEEE Trans. Ind.
Electron., vol. 61, no. 9, pp. 46654673, Sep. 2014.
[10] D. G. Lamar, M. Arias, A. Rodriguez, A. Fernandez, M. M. Hernando,
and J. Sebastian, Design-oriented analysis and performance evaluation
of a low-cost high-brightness LED driver based on flyback power factor
corrector, IEEE Trans. Ind. Electron., vol. 60, no. 7, pp. 26142626,
Jul. 2013.
[11] X. Wu, C. Hu, J. Zhang, and C. Zhao, Seriesparallel autoregulated
charge-balancing rectifier for multioutput light-emitting diode driver,
IEEE Trans. Ind. Electron., vol. 61, no. 3, pp. 12621268, Mar. 2014.
[12] K. H. Loo, W.-K. Lun, S.-C. Tan, Y. M. Lai, and C. K. Tse, On driving
techniques for LEDs: Toward a generalized methodology, IEEE Trans.
Power Electron., vol. 24, no. 12, pp. 29672976, Dec. 2009.
[13] V. C. Bender et al., An optimized methodology for led lighting system
designers: A photometric analysis, in Proc. 38th Annu. Conf. IEEE Ind.
Electron. Soc., Montreal, QC, Canada, Oct. 2012, pp. 45214526.
[14] P. Pickering, Towards a systematic methodology for the design, testing and manufacture of high brightness light emitting diode lighting
luminaires, Ph.D. dissertation, Faculty Eng. Phys. Sci. School Mech.,
Aerosp. Civil Eng., Univ. Manchester, Manchester, U.K., 2013.
[15] S. Y. R. Hui, H. Chen, and X. Tao, An extended photoelectrothermal
theory for LED systems: A tutorial from device characteristic to system
design for general lighting, IEEE Trans. Power Electron., vol. 27,
no. 11, pp. 45714583, Nov. 2012.
[16] V. C. Bender, O. Iaronka, W. D. Vizzotto, M. A. Dalla Costa,
R. N. do Prado, and T. B. Marchesan, Design methodology for lightemitting diode systems by considering an electrothermal model, IEEE
Trans. Electron Devices, vol. 60, no. 11, pp. 37993806, Nov. 2013.
[17] S. Y. Hui and Y. X. Qin, A general photo-electro-thermal theory for
light emitting diode (LED) systems, IEEE Trans. Power Electron.,
vol. 24, no. 8, pp. 19671976, Aug. 2009.
[18] A. Poppe and C. J. M. Lasance, On the standardization of thermal
characterization of LEDs, in Proc. 25th Annu. IEEE Semiconductor
Thermal Meas. Manage. Symp., San Jose, CA, USA, Mar. 2009,
pp. 151158.
[19] X. Tao and S. Y. R. Hui, Dynamic photoelectrothermal theory for lightemitting diode systems, IEEE Trans. Ind. Electron., vol. 59, no. 4,
pp. 17511759, Apr. 2012.

[20] A. Poppe and T. Temesvlgyi, A general multi-domain LED model and


its validation by means of AC thermal impedance, in Proc. 29th Annu.
IEEE Symp. Semiconductor Thermal Meas. Manage. Symp., San Jose,
CA, USA, Mar. 2013, pp. 137142.
[21] X. Tao and D. Zhang, Thermal parameter extraction method for lightemitting diode (LED) systems, IEEE Trans. Electron Devices, vol. 60,
no. 6, pp. 19311937, Jun. 2013.
[22] X. Tao, H. Chen, S. N. Li, and S. Y. R. Hui, A new noncontact
method for the prediction of both internal thermal resistance and
junction temperature of white light-emitting diodes, IEEE Trans. Power
Electron., vol. 27, no. 4, pp. 21842192, Apr. 2012.
[23] Y. Wang and S. Boyd, Fast model predictive control using online
optimization, IEEE Trans. Control Syst. Technol., vol. 18, no. 2,
pp. 267278, Mar. 2010.
[24] J. S. A. Carneiro and L. Ferrarini, Preventing thermal overloads in
transmission circuits via model predictive control, IEEE Trans. Control
Syst. Technol., vol. 18, no. 6, pp. 14061412, Nov. 2010.
[25] C.-W. Tang, B.-J. Huang, and S.-P. Ying, Illumination and color control
in red-green-blue light-emitting diode, IEEE Trans. Power Electron.,
vol. 29, no. 9, pp. 49214937, Sep. 2014.
[26] J. Park and C. C. Lee, An electrical model with junction temperature for light-emitting diodes and the impact on conversion efficiency, IEEE Electron Device Lett., vol. 26, no. 5, pp. 308310,
May 2005.
[27] D. Gacio, J. M. Alonso, J. Garcia, M. S. Perdigo, E. S. Saraiva, and
F. E. Bisogno, Effects of the junction temperature on the dynamic
resistance of white LEDs, IEEE Trans. Ind. Appl., vol. 49, no. 2,
pp. 750760, Mar./Apr. 2013.
[28] P. S. Almeida, V. C. Bender, H. A. C. Braga, M. A. Dalla Costa,
T. B. Marchesan, and J. M. Alonso, Static and dynamic photoelectrothermal modeling of LED lamps including low-frequency current ripple effects, IEEE Trans. Power Electron., vol. 30, no. 7,
pp. 38413851, Jul. 2015.
[29] H. Chen and S. Y. Hui, Dynamic prediction of correlated color
temperature and color rendering index of phosphor-coated white lightemitting diodes, IEEE Trans. Ind. Electron., vol. 61, no. 2, pp. 784797,
Feb. 2014.
[30] L. Tan, J. Li, K. Wang, and S. Liu, Effects of defects on the thermal and optical performance of high-brightness light-emitting diodes,
IEEE Trans. Electron. Packag. Manuf., vol. 32, no. 4, pp. 233240,
Oct. 2009.
[31] E. F. Schubert, Light-Emitting Diodes. Cambridge, U.K.:
Cambridge Univ. Press, 2006.
[32] Y.-S. Tyan, Organic light-emitting-diode lighting overview, J. Photon.
Energy, vol. 1, no. 1, p. 011009, 2011.
[33] Y. Ohno, Spectral design considerations for white LED color rendering, Opt. Eng., vol. 44, no. 11, p. 111302, 2005.
[34] T. W. Murphy, Jr., Maximum spectral luminous efficacy of white light,
J. Appl. Phys., vol. 111, no. 10, p. 104909, 2012.
[35] S. Baccari, L. Iannelli, and F. Vasca, A parallel algorithm for
implicit model predictive control with barrier function, in Proc.
IEEE Int. Conf. Control Appl., Dubrovnik, Croatia, Oct. 2012,
pp. 14051410.
[36] F. Vasca and L. Iannelli, Eds., Dynamics and Control of Switched
Electronic Systems. London, U.K.: Springer, 2012.

You might also like