Boost Paper Unified Rev NMG 24-02-2016

Embed Size (px)

Citation preview

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    1/30

     Energies 2016, xx, 1-x; doi:10.3390/——OPEN ACCESS

    energiesISSN 1996-1073

    www.mdpi.com/journal/energies

     Article

    A Framework to Power Electronic Converters Design and

    Control By Including Dynamical System Response Criteria:

    Non-ideal Boost DC/DC Converter Application

    Jorge H. Urrea-Quintero   1, Nicolás Muñoz-Galeano   1,* and Lina M. Gómez   2

    1 Universidad de Antioquia, Medellín 05001000, Colombia; e-mail: [email protected] Universidad National de Colombia, Medellín 05001000, Colombia; e-mail: [email protected].

    * Author to whom correspondence should be addressed; [email protected], tel:

    +57-4-2196445.

     Received: xx / Accepted: xx / Published: xx

    Abstract:   This paper presents a framework to design and control Power Electronic

    Converters (PECs). For this purpose, a non-ideal Boost DC-DC converter is dynamically

    modeled and analyzed in steady-state. A non-linear average dynamical model is derived in its

    general and bilinear form applying the Kirchhoff’s laws. Obtained average model is suitable

    for implementation in any computational tool to large and small transients system analysis.

    Then, steady-state model is obtained from average model, which allows to derive expressions

    for equilibrium conversion ratio  M (D)  and efficiency  η   of the system. Later, conditions

    for Continuous Conduction Mode (CCM) and Discontinuous Conduction Mode (DCM)

    are found. Simulations were made to see how parasitics losses affect both equilibriumconversion ratio and efficiency of the system. Then, inductor current and capacitor voltage

    ripples analysis are carried out to find lower boundaries for inductor and capacitor values.

    Boost DC-DC converter passive elements are sizing taking into account both steady-state

    and zeros-based analyses. Finally, a non-ideal Boost DC-DC converter is designed and a

    PI-based Current Mode Control (CMC) structure is implemented. PI controllers are tuning

    by means of root-locus technique. Designed Boost DC-DC converter is implemented in

    PSIM and system operating requirements are satisfactorily verified.

    Keywords:  power electronic converters; losses analysis; dynamical modeling; steady-state;efficiency.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    2/30

     Energies 2016, xx   2

    1. Introduction

    Technological developments in power electronic fields have increased the use of DC-DC converters

    in a large variety of applications, from the simplest ones (power supply for mobile phones or laptops

    [1]) to more demanding ones (applications in the aeronautics field [2,3], automobiles industry [4–6],

    telecommunications [7,8], and renewable energy field [9,10]). DC-DC converters main role is to

    adjust voltage level, providing a regulated output voltage based on a variable voltage power supply.

    In consequence, DC-DC converters represent an interesting and active research domain [11].

    Design procedures of PECs must stablish a trade-off between passive elements sizing and dynamical

    performance due to the close dependence between them, in a way that dynamical performance is not

    deteriorated and operating requirements are satisfied [12]. This task represents a challenge for designers

    because it requires a deep understanding of system dynamics, which, generally, implies to construct a

    non-linear system dynamical model and its implementation in any computational tool [13,14].Modeling and simulation of PECs are essential steps that enable design verifications and control

    of numerous electrical energy systems including modern electric grid and its components, distributed

    energy resources, as well as electrical systems of ships, aircraft, vehicles, industrial automation, among

    others [15].

    Dynamical modeling and steady-state analysis of PECs have received a significant attention as tools

    to achieve a deep understanding to suitable system design [11,12,15–21]. Dynamical modeling provides

    a set of equations that describe dynamical system evolution, which allows to analyze dynamical system

    performance and its relation with passive elements values [12]. In the other hand, steady-state analysis

    provides expressions to determine PEC: (a) (M (D)), (b) (η), and (c) CCM and DCM boundaries [13,16,19].

    Multi-resolution PEC models can be constructed, where ideal as well as non-ideal passive elements

    are taken into account [22]. If only ideal passive elements are considered, PEC model is simplified. As

    a consequence, if parasitic losses are neglected, models do not adequately represent the PEC behavior in

    its entire operation range [23]. Moreover, ideal PEC models can not predict both M (D) and η limitations

    [19]. Furthermore, ideal PEC models might lead to failure in predicting fast-scale instabilities [ 24,25].

    Parasitics losses can be considered in the PEC design stage when system dynamical performance

    and  η   are taken into account [19]. Parasitics losses are typically modeled as appropriate Equivalent

    Series Resistances (ESR) associated with PEC passive elements [11,16,26,27]. Several works proposing

    different PEC modeling approaches have been carried out [11,22,26,28–32]. Nevertheless, in the

    reviewed literature, it does not exist a consensus about what is a suitable detailed level of the model,

    such that the PEC is accurately represented, without the model derivation become as a challenge for

    the designer. However, the trend remains the so-called average models, which describe low-frequency

    dynamics of the system, neglecting high-frequency dynamics due to the switching [12,14,15,33–36].

    From [15], it is possible to conclude that average models could be the best option to PECs

    representation due to themselves:   (a)   are continuous and allow conducting large-signal time-domain

    transient studies of the system with multiple power electronic modules, controllers, and mechanicalsubsystems; and   (b)  allow small-signal analysis, where the frequency-domain transfer functions and

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    3/30

     Energies 2016, xx   3

    impedance characteristics are typically required for the design of controllers and stability analysis of the

    overall system.

    Average models that regard as well as disregard, some or all, parasitics losses have been presented

    [12,14,34,36]. Recent works [26,36–40] show that a practical level of model detail for PECs includesparasitics losses as ESRs and disregards losses due to the switching. Models with this level of detail are

    suitable for system designed, M (D) derivation, η  analysis, and dynamical performance evaluation [38].

    Moreover, these models are suitable for control purposes [14,36].

    It is clear that based on average models, PECs can be designed and a dynamical performance analysis

    can be carried out. Nevertheless, it is needed a design procedure that comprise all necessary steps to

    design a PEC such that a set of given operating requirements are satisfied. Additionally, this design

    procedure must be simple but useful.

    In PECs field only few works that take into account the integration of system design and control

    have been carried out [41–43]. In [41], the integrated design and control of a Buck DC-DC converter isinvestigated. This paper simultaneously optimal design the converter and its control structure. Averaged

    linear model is employed as converter model. Covariance Control Theory (CCT) is adopted as method

    to control structure design. CCT is a method for parameterization of all-stabilizing controller in terms of 

    the closed-loop covariance matrix.

    Authors in [42] investigate how to maximize the dynamical performance of a Buck-Boost DC-DC

    converter using an integrated design and control approach. As the considered Buck-Boost DC-DC

    converter has nonlinear dynamics, a linearized averaged model is considered to simplify the problem. As

    controller, a PID is selected. Averaged model parameters and controller parameters to be tuned are state

    variables dependent, thus a nonlinear optimization problem is obtained. Several Buck-Boost DC-DC

    converter reference trajectories are taken as reference to measure the converter dynamical performance

    using a quadratic function cost. These reference trajectories are including in the function cost to be

    optimized. Sequential quadratic programming is then used to solve the nonlinear optimization problem.

    The two approaches presented by [41] and [42] fix the control structure. They provide optimal

    parameters for fixed control structure, but they do not, in general, provide the optimal controller. A

    reason that avoids finding out the optimal controller is the necessity to test iteratively all possible control

    structures to find the optimal one.

    Authors in [43] tackle the gap between methods that optimally design circuit parameters with anon-optimal controller and optimal control methods. The scope of this work is to optimally design

    circuit parameters for control purpose that give a performance close to the limits of the circuit, but

    without imposing any control structure. The considered design objectives are dynamical performance

    and energy efficiency. In this work is proposed a simultaneous optimization of the circuit parameters

    and control input using a variable substitution technique and a decomposition of the general min-max

    problem into two simple min-max problems. The first min-max optimization problem yield the optimal

    sequence of state for the worse case load, which is unique, one of the sought circuit parameter and

    an auxiliary control input which embeds the effect of the other disturbances. These results are used

    in a second min-max optimization problem to obtain real optimal control input and remaining circuit

    parameters.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    4/30

     Energies 2016, xx   4

    Although in [43] none control structure is fixed, optimization techniques in this work are more

    complex than optimization techniques adopted in [41] and [42]. In consequence, still is needed a

    framework to easy design and control PECs. In this paper, a framework to easy design and control

    PECs is provided. The detailed analysis of a non-ideal Boost DC-DC converter operating in CCM [44]is carried out since is one of the basic DC-DC converter topologies.

    In this paper, a typical sequential form to PECs design is adopted, i.e., PEC structure is selected,

    passive elements are sizing, and control structure is designed. However, unlike typical approaches to

    PECs design, in this paper, a zeros-based analysis is carried out in order to include a dynamical system

    response criteria to select passive elements size.

    In a system, dynamical response is determined by poles and zeros location. Poles do not depend on

    either system inputs or outputs. However, poles determine whether the system is stable or unstable, as

    well as its natural frequency. Moreover, every pole generates a natural mode in the system response. In

    contrast, system zeros are determined by the selected system inputs and outputs. Zeros location is relatedto system performance limitations, additional overshoot in unit step response, and undershoot magnitude

    due to the system non-minimum phase behavior associated with Right Half Plane (RHP) zeros (unstable

    zeros) [45].

    Zeros modify system natural modes. PEC design process could takes into account the zeros location

    such that several system dynamical characteristics are avoided such as both large currents and voltages

    overshoots, or both sharp currents or voltages undershoots. Large currents or voltage overshoots can

    cause converter failures. Otherwise, sharp currents or voltages undershoots are undesired behavior when

    classical control structures are employed due to tracking limitations with feedback systems [46], [47].

    Accordingly, zeros location analysis objective is to find suitable values for  L and C  such that RHP zeros

    are avoided or that their impact are attenuated. In consequence, a trade-off between passive elements

    size and system dynamical performance is established.

    In this paper, widely accepted Current-Mode Control (CMC) structure for Boost DC-DC converter is

    designed [48], [49]. The Boost DC-DC converter contains a RHP zero, i.e, the Boost DC-DC converter

    could have a non-minimum phase behavior. Non-minimum phase behavior is the reason why a CMC

    structure is needed [48]. CMC structure employs cascaded loops. This cascaded structure allows the

    output voltage regulation while preserving the inductor current within specified safety limits. The outer

    control loop deals with voltage regulation imposing low-frequency dynamics and the inner loop concernsthe faster current control. The voltage controller provides the setpoint of the inductor current, and this

    latter acts as the control input of the outer voltage loop.

    Standard controllers are employed for the CMC structure. Simple and robust classical invariant PI

    controllers that are tuned based on linearized Single-Input-Single-Output (SISO) averaged models of the

    Boost DC-DC converter. These PI controllers generate a continuous control signal (duty ratio d), which

    needs a modulation PWM so it may be applied to the power switching gates.

    The remaining of the paper is organized as follows: in the next section, both time- and

    frequency-domain models of non-ideal Boost DC-DC converter is derived. In section 3, the non-ideal

    Boost DC-DC converter is panned out in steady-state. Section 3 is composed of subsections  3.1,  3.2,

    3.3, and 3.4. In subsection 3.1 an expression for M (D) is derived including ESRs. In subsection 3.2, an

    efficiency expression η  is derived taking into account ESRs. In subsection 3.3, conditions for operate in

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    5/30

     Energies 2016, xx   5

    CCM or DCM are found. Knowing these conditions, it is possible to design a Boost DC-DC converter

    that always operate in CCM if load variations are given. In subsection  3.4 inductor current and capacitor

    voltage ripples analysis are carried out to find lower boundaries for inductor and capacitor values such

    that ripples requirements are satisfied. In section 4, passive elements are sizing such that operatingrequirements are satisfied and system dynamical performance is achieved. Mathematical model is

    contrasted with a PSIM implementation of the Boost DC-DC converter. In section 5, widely accepted

    Current-Mode Control (CMC) structure for Boost DC-DC converter is designed. The aim is to design

    a suitable control structure such that control objective is achieved. This stage is composed of following

    subsections: in subsection 5.1,  controllers are tuned by means of root-locus technique; in subsection

    5.2, closed-loop system performance verification is carried out; and in subsection  5.3, system operating

    requirements are verified. Finally, conclusions are provided in section  6.

    2. Non-Linear Dynamical Modelling

    Figure 1 shows a circuital representation of a typical non-ideal Boost DC-DC converter supplying

    a dominant-current load represented as a Norton equivalent model. Engines and inverters are common

    dominant-current loads that can be supplied by a Boost DC-DC converter. In Figure 1, L  is inductor, C 

    is capacitor, and RL and RC  are ESR for L and C , respectively. RL and RC  represent all parasitic losses

    in adopted non-ideal Boost DC-DC converter for this paper.

    The Boost DC-DC converter operating in CCM can take two configurations according to the switch

    position as shown in Figure  1: configurations  (a)  and   (b)   correspond to switch  H   =   {h1, h2}  being

    turned on h1  = 1, h2  = 0 and turned off  h1  = 0, h2  = 1, respectively. Therefore, switching function ucan be defined as follows:  u  =  h1 = 1 − h2.

    State variables (inductor current iL and capacitor voltage vC ) are defined to represent energy variation

    of the system. System inputs are u, DC input voltage source  vg, and current source  io   that is useful

    to represent system current perturbations; system outputs are output voltage  vo   and  iL. By applying

    Kirchhoff’s laws to system in Figure 1, dynamical model given by equations (1)-(2) are derived.

    LdiLdt

      =   vg − (RL + φC (1 − u))iL + φC R  − 1

    (1 − u)vC  + φC (1 − u)io   (1)

    C dvC dt

      =

      1

    1 + αC 

    (1 − u)iL −

     vC R − io

      (2)

    Where αC  = RC 

    R  and φC  =

      RC 1 + αC 

    .

    In this paper, widely accepted PI controllers based Current-Mode Control (CMC) structure for Boost

    DC-DC converter is adopted [48], [49]. PI controllers tuning require a frequency-domain model. From

    non-linear dynamical model given by equations (1) and (2), it is possible to obtain a linear state-space

    model of Boost DC-DC. Next, frequency-domain model is obtained by means of the realization given

    by the equation (3).

    G(s) =  1

    det(sI − A)C  [adj(sI −A)] B + D   (3)

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    6/30

     Energies 2016, xx   6

    Figure 1.  Nonideal DC-DC Boost converter circuital diagram.

    Linear state-space model for Boost DC-DC converter is given by equation (10). Where x = [iL, vC ]T 

    ,

    u = [d, vg, io]T 

    , and y  = [iL, vo]T 

    .

    In equation (10), I L and V C , D  and I o are states and inputs in their rated values, respectively.

    Applying realization given by equation (3) to equations in (10), transfer functions given by equations

    (4) and (5) are obtained. Numerator of equation (4) has three components, one for each system input,

    i.e., d,  vg, and io, respectively. Numerator of equation (5) has three components, one for each system

    input, i.e., d, vg, and io, respectively.

    Gvo(s) =

    −(RLI Ls + φC RI o(1 −D) + φC (1 − D)V C  − R(1 − D)V C  + RRLI L)

    (φC C (1 + αC )s + 1)

    R(φC C (1 + αC )s + 1)(1 − D)

    R(φC (1 − D)2 − φC (1 −D) − RL)(φC C (1 + αC )s + 1)

    (RLC (1 + αC ))s

    2 + φC RC (1 −D)(1 + αC )s − (φC  − R)(1 − D)2

    +φC (1 − D) + RL

      (4)

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    7/30

     Energies 2016, xx   7

    GiL(s) =

    1

    R

    (((φC R2CI L) − (φC R2CI o) − (φC RCV C ) + (R2CV C ))(1 + αC )s

    −φC R(1 −D)I L + R2

    (1 −D)I L + φC RI L − φC RI o − (φC  −R)V C )

    RC (1 + αC )s + 1

    φC RC (1 −D)(1 + αC )s + R(1 −D)

    (RLC (1 + αC ))s2 + φC RC (1 −D)(1 + αC )s − (φC  − R)(1 − D)2

    +φC (1 − D) + RL

      (5)

    Once the current control loop in the CMC structure is closed, equivalent simplified representation of 

    the Boost DC-DC converter showed in Figure 2 is obtained. Large- and small-signal models of simplified

    Boost DC-DC converter are given by equations (6) and (7), respectively. Applying realization given by

    the equation (3), transfer functions of simplified model are given by the equation (9).

    Figure 2.  The equivalent simplified representation of the Boost DC-DC converter.

    C dvC dt

      =

      1

    1 + αC 

    iLREF (1 − D) −

     vC R − io

      (6)

     v̇C    = −  1

    1 + αC   1

    RC  vC +

      1

    1 + αC 

     (1 −D)

    C   −

      1

    1 + αC 

     1

    iLREF 

    io

      (7)

    vo

      =

      1

    1 + αC 

    vC 

    +

      1

    1 + αC 

    RC (1 − D)   −

      1

    1 + αC 

    RC 

    iLREF 

    io

      (8)

    Gvo(s) =[(1 + αC )RCRC s + R + RC ] (1 −D)−(1 + αC )RCRC s + R + RC 

    (1 + αC ) [(1 + αC )RCs + 1]  (9)

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    8/30

     Energies 2016, xx   8

      ˙iLv̇C    =

    RL + φC 

    L

    φC R  − 1

    (1 −D)

    L

    (1 −D)

    (1 + αC )C   −

      1

    RC (1 + αC )

    iL

    vC  +

    (10)

    φC I L − V C 

    φC R  − 1

    − φC I o

    L

    1

    L

    φC (1 −D)

    L

    −  I L

    (1 + αC )C 

      0   −  1

    (1 + αC C )

    dvg

    io

    iL

    vo

      =

    1 0

    φC (1 −D) 1 − φC 

    R

    iL

    vC 

    +

      0 0 0

    −φC I L   0   −φC 

    dvgio

    3. Steady-State Analysis

    Once the system model is obtained following analysis might carried out:   (1)  derivation of  M (D)

    expression,  (2)  losses effect and efficiency η  expression derivation,  (3)  conditions analysis of CCM and

    DCM, and (4) inductor current ∆iL and capacitor voltage ∆vC  ripples analysis. Aim of these analysis isto determine suitable passive elements (L and C ) boundaries which satisfy design requirements.

    3.1. First Step: Derivation of the Equilibrium Conversion Ratio  M (D) Expression

    Steady-state model allows to obtain expressions for average rated values for both  vC   and   iL   as a

    function of system inputs and parameters. The steady-state model is obtained by setting to zero model

    given by equations (1) and (2). Thus, equations (11) and (12) are obtained.

    I L =

      V C 

    R (1 − D)   (11)

    V g  =

    RL + φC  (1 −D)

    R (1 −D)

    φC R  − 1

    (1 − D)

    V o   (12)

    Capital letters indicate the steady-state average values, thus:  V g  is input voltage, V o is output voltage,

    I L is inductor current, and D  is duty cycle.

    From equations (1)-(2), it is found that V o  = V C  in steady-state. Therefore, the expression of  M (D)

    for the non-ideal Boost DC-DC converter is given by the equation (13).

    M (D) = V oV g =

      (1 − D)1 −

     φC R

    (1 − D)2 +

     φC R

      (1 −D) + αL(13)

    where

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    9/30

     Energies 2016, xx   9

    αL = RL

    R

    M (D) indicates the Boost DC-DC converter elevation voltage factor in terms of  D, R, RL, and φC .

    It is important to remark that if  RL  = 0 and  RC   = 0  in equation (13), this equation coincides with

    ideal Boost DC-DC M (D), i.e., M (D) =  1

    (1 −D).

    3.2. Second Step: Losses Effect and Efficiency Expression Derivation

    If ideal conditions are considered in Boost DC-DC converter analysis, M (D) tend to ∞ when D tend

    to 1  and converter has an efficiency equal to  100%. However, these characteristics are not true in a real

    PEC application.

    This section shows how parasitics losses affect both  M (D)  and  η   in the non-ideal Boost DC-DC

    converter case. Losses effect and efficiency analysis are carried out in order to find suitable values for

    RL and RC  such that designed PEC satisfies desired operating requirements.

    The DC transformer is used to model ideal functions performed by a DC-DC converter [44]. This

    model correctly represents the relations between DC voltages and currents of the converter. The resulting

    model can be directly solved to find voltages, currents, losses and efficiency in the non-ideal Boost

    DC-DC converter.

    Equations (14) and (15) are obtained from equations (1) and (2). These equations stablish that average

    value of both iL and vC  are equal to zero in steady-state.

    0 =   V g − (RL − φC  (1 −D)) I L −

      1

    1 + αC 

    V C  (1 −D)   

    V  d

    (14)

    0 =   −

      1

    1 + αC 

     V C 

    R  +

      1

    1 + αC 

    I L (1 −D)

       I d

    (15)

    A circuital model based on equations (14) and (15) is built. This circuital model describes the DCbehavior of the Boost DC-DC converter with inductor and capacitor losses. A circuit of which Kirchoff 

    loop and node equation represent equations (14) and (15).

    Equation (14) describes the average inductor voltage in the time interval  T s  and DC components of 

    voltages around a loop containing the inductor  L, with loop current equal to the DC inductor current

    I L. Therefore, a circuit containing a loop with current I L, corresponding to equation (14), is built in

    Figure 3a. The first term in equation (14) is input voltage V g. The second term is a voltage drop of value

    RL − φC (1 − D)I L, which is proportional to current  I L   in the loop. The third term is the dependent

    voltage source V d =   11+αC  V C (1−D), which depends on V C . This term is modeled using a dependentvoltage source as shown in Figure 3a. The polarity of the source is chosen to satisfy equation (14).Equation (15) describes the DC components of currents flowing into a node connected to the capacitor

    C . Therefore, a circuit containing a node connected to the capacitor is built in Figure 3b, of which the

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    10/30

     Energies 2016, xx   10

    node equation satisfies equation (15). The first term in equation (15) is a current magnitude

      1

    1+αC 

     V  oR

     ,

    proportional to the DC capacitor voltage  V C   in steady-state. The second term is a dependent current

    source I d =   1

    1+αC I L(1−D), which depends on DC inductor current I L. This term is modeled using a

    dependent current source as shown in Figure 3b. The polarity of the source is chosen to satisfy equation(15).

    Figure 3.  Non-ideal Boost DC-DC Converter source depended equivalent circuit diagram.

    Figures 3a and 3b are combined into a singe circuit. Circuits of Figures 3a and 3b can be further

    simplified by acknowledging that   V d   and   I d   constitute an ideal DC transformer. In each case, thecoefficient is

      1

    1+αC 

    (1 − D). Hence, the fact that this coefficient appears on the primary rather than

    the secondary side is owing to the symmetry of the transformer. Therefore, there is an equivalent DC

    transformer, with turns ratio

      1

    1+αC 

    (1 −D) : 1, that represents the non-ideal Boost DC-DC converter.

    Substitution of the ideal DC transformer model for the dependent sources yields the equivalent circuit of 

    Figure 4.

    Figure 4.  Ideal DC transformer model of the Boost DC-DC converter.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    11/30

     Energies 2016, xx   11

    The main importance of the equivalent circuit model in Figure  4 is that it allows to compute converter

    efficiency η. Figure 4 predicts that the convert input power is given by the equation (16).

    P in = V gI g  = V g   V C (1 − D)R   (16)The load current is equal to the current in the secondary of the ideal DC transformer  I d. Hence the

    model predicts that the converter output power is given by the equation (17).

    P out =  V oI o = V o

      V C 

    (1 + αC )R

      (17)

    Therefore, η  = P out

    P inis given by equation (18).

    η =   11 + αC  (1 − D) M  (D)   (18)Simulations of equations (13) and (18) are shown in Figure 5  for several values of  αC  and αL ratios

    in order to see how much losses affect both M (D) and η.

    Figure 5 shows αC  and αL ratios effects on M (D) and η. Simulations on Figure 5a were carried out

    for every possible combination of  αC   = [0, 0.05, 0.1]  and  αL  = [0, 0.05, 0.1].   αC   = 0 and  αL  = 0 are

    the ideal case for the Boost DC-DC converter (without losses). The group of curves that corresponds

    to the black lines (above lines group) has  αL   = 0  and αC   = [0, 0.05, 0.1]. It can be noted that while

    αC  increases, curves move towards right. However,  M (D) in the converter has an increasing trend and

    eventually tend to infinite. For group of curves  αL  = 0.05 and αC  = [0, 0.05, 0.1] (middle group lines),it is noted that while αC   increases, M (D) decreases. M (D) also decreases for lines group αL = 0.1 and

    αC  = [0, 0.05, 0.1] (below lines group). From Figure 5, it can be concluded that M (D) has a tendency to

    decrease in higher proportion when αL increases than when αC  increases. Really, losses on the inductor

    determine quadratic trend of the M (D) curves; while αC , to a lesser extent, determines the Boost DC-DC

    converter gain. Both αL and αC  determine M (D) form, however, αL determines curve trend.

    From Figure 5b, it is observed that maximum η value reached by the converter is determined by losses

    and it is given in D  = 0  for every M (D)  curve. For the studied case, it is the combination of  αC   and

    αL what determines maximum value of  η. η  decreases while D increases dropping to 0 when D tends to

    1, hence, converter should operate, as far as possible, with low  D  values. Increases of  αC  or αL  causesdecrease of  η , therefore, as far as possible, αC   and αL  should tend to zero to guarantee high converter

    η. From Figure 5a it seen that for a value of  M (D)  there are two possible  D  values; however, from

    Figure 5b is seen that higher  D  value always corresponds to a less  η . Hence, if it is desired to operate

    the converter as efficient as possible, always lesser  D  values that satisfies  M (D)  requirement must be

    selected. Similar to Figure 5a, for Figure 5b, values were also clustered in groups of curves. For  αL = 0

    and αC   = [0, 0.05, 0.1]  (above lines group), it noted that while αC   increases,  η  decreases notoriously.

    For curves group αL  = 0.05  y  αC   = [0, 0.05, 0.1]  (middle lines group) and for curves group αL  = 0.1

    y  αC   = [0, 0.05, 0.1]   (below lines group), something similar occurs,  η  decreases while αC   increases.

    However, decrease in η is more notable when αL increases than when αC  increases. Combined effect for

    high αL and αC  leads to highly inefficient system with high losses.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    12/30

     Energies 2016, xx   12

    Figure 5.  (a) the conversion ratio M (D) v.s. the duty cycle D .  (b) Efficiency η v.s. the duty

    cycle D

    3.3. Third Step: Conditions Analysis of Continuous Conduction Mode (CCM) and Discontinuous

    Conduction Mode (DCM)

    The CCM is suggested since DCM causes larger voltage ripple [50], [51]. Furthermore, the peak 

    inductor current in DCM is higher than that in CCM with the same power level and current stress on the

    power switch is also higher [52].

    By [44], the sufficient conditions for operation in the CCM and DCM are given by equations (19) and

    (20), respectively.

    |I L| > |∆iL|   for CCM    (19)

    |I L| < |∆iL|   for DCM    (20)

    Where |I L| and |∆iL| are found assuming that the converter operates in CCM.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    13/30

     Energies 2016, xx   13

    The DCM operation condition for the Boost DC-DC converter is given by the equation ( 21).

    D

    (1 −D)

    M  (D)  −

      1

    R >

      2L

    RT s(21)

    where

    K (D) = D

    (1 −D)

    M  (D)  −

      1

    R

      (22)

    K  =  2L

    RT s(23)

    Dimensionless parameter K  is a converter measure of the tendency to operate in the DCM [44]. Large

    values of  K  lead to CCM, while small values of  K  lead to the DCM for some values of  D. The critical

    value of  K  is the boundary between modes, K crit(D), and it is a function of  D  and R, RC , RL, i.e, load

    and ESRs of  C  and L.

    Function K crit(D) has a critical point given by the equation (24).

    D   =  1

    3 [−3αC  + αC αL + αL]   (24)

    There is a maximum for the function in critical point given by the equation ( 24) due to the fact that

    the second derivative of  K crit(D) given by the equation (25) is always negative for D ∈ [0, 1].

    d

    2

    K (D)dD2   = −   11 + αC  (6D + 2αC )   (25)

    From result given by equations (24) and (25), the sufficient condition for always operate in CCM is

    given by the equation (26).

    max (K (D)) < min (K )   (26)

    where

    min (K ) = min  2LRT sTherefore, if a value for  R   was given in the system specifications, a condition for the maximum

    possible value of  L  is given by the equation (27) such that equation (26) is assured.

    L > RT s

    2  max(K (D))   (27)

    Simulations of equations (22) and (23), varying R and L, are shown in Figure 6. Figure 6 shows how

    variations of  R  and L  affect both K crit(D) and K  functions.

    Figure 6 shows K (D) limits (continuous gray line) between CCM and DCM for the non-ideal Boost

    DC-DC converter. Above K (D), converter works in CCM, while below converter works in DCM.  Rand L  factors influence operating mode of the system.

    Figure 6a shows the effect on K  when varying L  = [20, 40, 60, 80]mH , setting R  in  50Ω.  R = 50Ω

    represents an adverse case for the system (it is a heavy load). From Figure 6a, it is observed that converter

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    14/30

     Energies 2016, xx   14

    Figure 6.  (a) K (D) v.s. K  varying R  .  (b) K (D) v.s. K  varying L.

    works in CCM for L values equal to 40, 60y80mH , while for L = 20mH  the line intercepts K (D) limit,

    hence, it works in DCM for several  D  values.

    Figure 6b shows the effect on  K   when varying  R   = [125, 150, 175, 200]Ω, setting  L   in  100mH .L  = 100mH  represents a favorable case for the converter since  ∆iL value decreases when L   increases

    and, therefore, probability of  ∆iL  > I L   (DCM condition). From Figure 6b, it is observed that while R

    decreases, the system is less likely to DCM. For  R  values equal to 150, 175y200Ω, the converter can or

    cannot operate in CCM depending on  D  value. To guarantee CCM operation, independent of  D  value,

    R value equal to 125Ω must be selected.

    3.4. Fourth Step: Inductor Current and Capacitor Voltage Ripple Analysis

    ∆iL  and ∆vC  analysis is carried out to determine constraint equations for a suitable choice of bothL  and  C   values. The analysis carried out in this section is suitable for the Boost DC-DC converter

    operating in CCM.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    15/30

     Energies 2016, xx   15

    Figure 7  shows both typical inductor voltage  vL   and iL   linear-ripple approximations. Slope, with

    iL  increasing or decreasing, is deduced from the analysis of  V L  at each subinterval of time taken into

    account. Typical values of current inductor ripple ∆iL  lie under  10%  of the full-load value of  I L  [44].

    From Figure 7 is seen thatiL begins at initial value iL(0). Then, iL increases with a constant slope duringthe first subinterval (DT s), h1  = 1 and h2  = 0. Switch changes to position h1  = 0 and  h2  = 1 at time

    t =  DT s. Then, iL decreases with a constant slope. Switch changes back to position  h1 = 1 and h2 = 0

    at time t  =  T s and the process repeats.

    Figure 7.   a.   Typical inductor voltage linear-ripple approximation.   b.   Typical current

    inductor linear-ripple approximation.

    As illustrated in Figure 7b, the peak inductor current I  pk is equal to I L plus the peak-to-average ripple

    ∆iL. I  pk flows through inductor and semiconductor devices that comprise the switch. The knowledge of 

    I  pk  is necessary when specifying the rating of the device.

    The ripple magnitude can be calculated knowing both the slope of   iL   and the length of the first

    subinterval (DT s). The iL  linear-ripple approximation is symmetrical about  I L, hence, during DT s, iL

    increases by 2∆iL  (since ∆iL  is the peak ripple, the peak-to-peak ripple is  2∆iL). Thus the change in

    current, 2∆iL, is equal to the slope with iL  increases.

    The inductor value L  can be chosen from equation (28).

    L =

    V g −   αL(1 − D)V o2∆iL

    DT s   (28)

    Where T s =  1

    f swand f sw is the converter switching frequency.

    Equation (28) is a lower boundary for  L  value, where L  can be choose such that in the worst Boost

    DC-DC operating condition a maximum ∆iL is attained.

    Likewise, vC  linear-ripple approximation is depicted in Figure  8b and an expression derived for the

    output voltage ripple peak magnitude ∆vC  is obtained. Capacitor current linear-ripple approximation I C 

    is depicted in Figure 8a.  vC  begins at initial value vC (0).  vC  decreases with a constant slope during the

    first subinterval (DT s), h1  = 1 and  h2  = 0. The switch change to position h1  = 0  and  h2  = 1 at timet  =  DT s. Then, vC   increases with a constant slope. The switch changes back to position h1   = 1  and

    h2 = 0 at time t  =  T s and the process repeats.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    16/30

     Energies 2016, xx   16

    Figure 8.   a.   Typical capacitor current linear-ripple approximation.   b.   Typical capacitor

    voltage linear-ripple approximation.

    The ripple magnitude can be calculated knowing both the slope of  vC  and the length of  DT s. Thechange in vC , −2∆vC , during DT s, is equal to the slope multiplied by  DT s. Equation (29) can be used

    to select capacitor value C  to obtain a given ∆vC .

    C  =  V o

    2R∆vC DT s   (29)

    Equation  29  is a lower boundary for  C  value, where C  can be choose such that in the worst Boost

    DC-DC operating condition a maximum ∆vC  is attained.

    4. Passive Elements Sizing

    In this section, non-ideal Boost DC-DC converter operating requirements are specified. Then, passive

    elements are sizing such that operating requirements are satisfied and system dynamical performance is

    achieved. Finally, mathematical model is contrasted with a PSIM implementation of the Boost DC-DC

    converter.

    4.1. System Operating Requirements

    In Boost DC-DC application, typical requirements are: input voltage range, output voltage range,

    output power range, output current range, operating frequency, output ripple and efficiency. Unless

    otherwise noted, continuous operating mode is assumed.

    The set of operating requirements for the Boost DC-DC converter are specified in Table 1.

    4.2. Passive elements sizing

    Once operating requirements are established, passive elements are sizing. Concerning to passive

    elements value selection, the main interest is to select suitable values for inductors and capacitors

    such that constraints like maximum physical admissible currents and voltages, converter efficiency, and

    converter conduction mode are satisfied by keeping an acceptable dynamical system performance.

    Via steady-state analysis expressions given by (28) and (29) were deduced for lower inductor  L  and

    capacitor C  boundaries, respectively. These deduced expressions are suitable to choose L and C  values in

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    17/30

     Energies 2016, xx   17

    Table 1.  Boost DC-DC converter operating requirements.

    RequirementValues

    Min Typ Max

    Input voltage range   30V 35V 40V

    Output voltage range   50V 70V 95V

    Output power range   0W 100W 300W

    Output current range   0A 2A 8A (At 50V)

    Operating frequency   100kHz

    Output current ripple   1% 5% 10%

    Output voltage ripple   0.1% 0.5% 1%

    Steady-State efficiency   90% 95% 98%

    Load   25Ω   50Ω   100Ω

    function of  ∆iL, ∆vc, states, and inputs values in steady-state. Additionally, from converter conduction

    mode analysis was deduced the expression given by (27) that allows to guarantee the Boost DC-DC

    CCM operation in all operating range by choosing L value given the load impedance value R.

    In PECs are desired that  ∆iL  ≤   max(∆iL)  and ∆vo  ≤   max(∆vo)  are assured in entire operation

    range. Then, based-on worst condition for ∆iL and ∆vo, L and C  lower boundaries can be deduced such

    that ripple constraints are satisfied. Equations (30) and (31) give lower boundaries for  L  and C  values,

    respectively.

    L ≥

    max(V g) −

      RL

    max(R)(1 −D) min(V o)

    2max(∆iL)

      DT s   (30)

    C  ≥  max(V o)

    2 min(R) max(∆vo)  (31)

    Equation (27) also gives a minimum boundary for  L  value. Then, equations (30) and (27) must be

    evaluated and maximum L value must be selected as lower boundary.

    Equation (30) depends of αC . Equation (27) depends of M (D), which depends of RL. Therefore, RC 

    and RL must be defined in order to choose L  value.

    From steady-state analysis, instead of calculate RC   and RL  values, it is suitable to establish  αC   and

    αL   values.   αC   and  αL   values can be chosen such that the system efficiency is  η   ≥   90%   in entire

    operating range. Worst Boost DC-DC condition is when M (D)  is maximized.   M (D) =  V o/V g, then

    max(M (D)) = max(V o)/ min(V g). According to operating requirement in Table 1,  min(V g) = 30V 

    and max(V o) = 95V , thus max(M (D)) ≈ 3.17. Losses ratios must be αC  

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    18/30

     Energies 2016, xx   18

    Figure 9.  (a) Conversion ratio M (D).  (b) Efficiency η.

    If   max(∆iL) = 0.1max(I L), in order to keep a converter safe operation [52],   L   ≥   326.34µH 

    according to equation (30). Evaluating equation (27),   L   ≥   40µH   to always operate in CCM.   iL

    ripple-based condition is a less restrictive boundary for   L   than CCM-based condition. Therefore,

    L ≥ 326.34µH  is the lower boundary for this element.

    If  max(∆vC ) = 0.01max(V o), in order to keep a converter safe operation [52],  C   ≥   14.120µF 

    according to equation (31).

    Minimum L  and C   values are selected as system parameters. By [36],  RL   = 150mΩ  and RC   =

    70mΩ  are selected due to the fact that in this paper a Boost DC-DC converter with similar operating

    requirements are designed and experimental tested with these  RL   and RC   vales. Next, a simulation

    of the designed Boost DC-DC converter is carried out. Figure 10 shows the step system response forV g  = 35V , V o  = 70V , I o  = 0A, L  = 326.34µH , C  = 14.120µF , R  = 50Ω, αC  = 0.0034, αL  = 0.006,

    and f sw = 100kH Z .

    Figure 10.  (a) Gvd  Step system response.  (b) GiLd Step system response.

    From Figure   10   is seen that, with minimum values of   L   and   C , Voltage Overshot   O.S.Gvd   =

    57.3427%, Current Overshot O.S.GiL

    d = 190.0448%, and system setting time ts = 2.3ms. From Figure

    10a is seen that the peak voltage value is 214.2V , while the final voltage value is 135.5V . From Figure

    10b is seen that the peak current value is  34.0A, while the final current value is  11.41A. A designed

    system with these overshoots needs oversized electronic devices such that these devices support both

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    19/30

     Energies 2016, xx   19

    peak voltages and current values without a system damage. However, the electronic devices oversizing

    can be expensive and inconvenient. Therefore, in this paper, an analysis of Boost DC-DC dynamical

    characteristics is carried out. Boost DC-DC dynamical characteristics analysis test the passive elements

    values impact in overshoots and system setting time.In Boost DC-DC converter applications,   d   is chosen as control input, while   vg   and   io   are

    considered disturbances. Thus,   d   variations effect is of primary interest over system output. Then,

    duty-ratio-to-voltage-output Gvd  and duty-ratio-to-inductor-current GiLd  transfer functions are studied,

    i.e., characteristics of equations (4) and (5) are studied.

    It is possible to obtain zeros of  Gvd  from system transfer function in equation (4). Zeros of  Gvd  are

    given by equations (32) and (33).

    s1 = −  1

    RC C   (32)

    s2 =  1

    L

    [(φC  − R)V C  − φC RI o]

    I LR  (1 −D) + RL

      (33)

    Zero given by the equation (32) is negative because it depends on circuit parameters, all of which are

    positive. However, zero given by the equation (33) is positive because it depends on circuit parameters,

    V C , and I L, all of which are positive. Also, in steady-state current source I o is equal to zero. Zero given

    by the equation (33) is then placed in the RHP of Laplace domain, which mean that if  vo  is selected as

    system output, the system could have a non-minimum phase behavior.

    Non-minimum phase behavior is a well-known result derived of the Boost DC-DC converter

    study [53]. To avoid this system behavior a cascade control structure has been proposed [53],

    [48]. Non-minimum phase behavior is avoided with this control structure since both   GiLd   and

    inductor-current-to-output-voltage GvoiLREF  transfer functions have a minimum phase behavior as will

    be shown.

    It is possible to obtain zeros of  G(s)iLd  from transfer functions in equation (5). Zero of  G(s)iLd   is

    given by the equation (34).

    s = −  1

    RC   1

    1 + αC   (R − φC )RI L(1 −D)

    φC RI L − φC RI o + (R − φC )V C + 1

      (34)

    Zero given by the equation (34) is negative because it depends on circuit parameters, all of which are

    positive and, in steady-state, I o is equal to zero. This zero is then placed in the Left Half Plane (LHP) of 

    Laplace domain, which means that the system has a minimum phase behavior.

    From equations (32) and (33) is seen that zeros of Gvd are in function of L and C  vales. From equation

    (34) also is seen than zero of  GiLd is in function of  L  and C  values. Therefore, a simulation was carried

    out to evaluate effects of large values for both  L  and  C . Figure 11 shows Gvd  and GiLd  overshoots and

    setting time for several values of  L and C .

    From Figure 11 is seen that minimum possible value for  C  causes maximum overshoot in vo. While

    a minimum possible value for L causes maximum overshoot in iL. Moreover, minimum C  and L valuesgive minimum system setting time.

    In contrast, large values for   C   cause high overshoot for   iL. While large values for   L   cause

    high system setting time. In consequence, two additional designed requirements are given in order

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    20/30

     Energies 2016, xx   20

    Figure 11.   (a) Gvd  Step system response varying C  and L: overshoot with zeros.   (b) GiLd

    Step system response varying C  and L: overshoot with zeros. c Step system response varying

    C  and L: setting time.

    to establish maximum possible values for   L   and   C   such that system overshoots and setting time

    are suitable:   (a)   maximum duty-ratio-to-output-voltage overshoot   max(O.S.Gvd)   and   (b)  maximum

    duty-ratio-to-inductor-current overshoot max(O.S.GiLd).

    From Figure 11, any effect over system performance is achieved if the values of both  L  and  C   are

    increased simultaneously. However, if either L  or C  values are increased both O.S.Gvd  and O.S.GiLd  are

    decreased. Nevertheless, larger values of  L  have a major impact that larger values of  C .

    L   = 1mH   and  C   = 15µF   are selected by results showed in Figure   11 since with these values

    O.S.Gvd

      ≈   52%  and O.S.GiLd

      ≈   100%, i.e.,  O.S.GiLd

      is reduced approximately  90%. Further,  ts   =

    3.4ms, i.e., the system setting time is increased  1.1ms. Thus, these L  and C  values establish a trade-off 

    between system overshoots and performance.

    4.3. System Frequency Response Verification

    Frequency response of both mathematical model and a PSIM circuital implementation are contrasted

    in order to validate dynamical model of the designed Boost DC-DC converter via simulation. The Boost

    DC-DC converter was parameterized with L  = 1mH ,  C   = 15µF ,  V g   = 35,  V o   = 70, I o   = 0,  αC   =

    0.0034, αL

     = 0.006, R  = 50Ω, and f sw

     = 100kH z . In consequence, I L

     = 2.8812A and D  = 0.5141 in

    the equilibrium point.

    Figure  12 presents the Boost DC-DC converter Bode diagrams of the mathematical model given

    by equations (4) and (5) and PSIM circuital implementation. Frequency response of the PSIM

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    21/30

     Energies 2016, xx   21

    Figure 12.  (a) Gvd  Bode diagram.  (b) GiLd Bode diagram.

    circuital implementation coincides with mathematical model. Then, PSIM circuital implementation is

    satisfactorily reproduced by mathematical model.

    5. Control Structure Design

    Here, widely accepted Current-Mode Control (CMC) structure for Boost DC-DC converter is

    designed [48], [49]. The aim is to design a suitable control structure such that control objective

    is achieved. This stage is composed of following subsections:   (1)   control structure selection,

    (2)   controllers tuning,   (3)   closed-loop system performance verification, and   (4)   system operating

    requirements verification.

    5.1. Controllers Tuning

    Generally speaking, converter control design is focused on imposing desired low-frequency behaviorto the system by means of specified closed-loop requirements. In the Boost DC-DC converter operating

    in a switch-mode power supply, feeding a certain variable load, d needs adjustments in order to ensure a

    constant  vo for the entire operating range (voltage regulation). Besides, against any system disturbance,

    d value needs to be adjusted such that the system can be driven back to the operating point.

    The controller tuning task begins with a set of design specifications. The set of specifications are a

    group of goals for the behavior of the controlled system. Specifications are composed for both transient

    behavior (e.g. rise time, setting time, and maximum closed-loop system overshot) and stability margins

    (e.g., relative stability, gain margin, phase margin).

    Time domain specifications are placed by the system performance specifications. The transient

    response of a regulated system is typically limited in terms of maximum deviation from the rated output

    value and setting time in response to a transient. The goal for the Boost DC-DC converter controller

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    22/30

     Energies 2016, xx   22

    design presented is to control the output voltage vo  to within 2%  of rated value (i.e.  68.6V   to 71.4V ) in

    response to unit step transients in both input voltage vg  and current source io. Also, the controller should

    be able to maintain the rated output voltage within the tolerances as the input varies over a range from

    30V   to 40V , though this should be considered a steady-state, not transient, operating requirement. As afinal specification, steady-state error in the output voltage vo should be eliminated.

    Inner loop bandwidth (i.e., current loop) must be  20kH z  or less due to the fact that the Boost DC-DC

    converter switching frequency is 100kH z , and outer loop bandwidth (i.e., voltage loop) must be smaller

    than 1

    5 inner loop bandwidth [54], i.e., smaller than 5kH z . Additionally, a robustness index M s  <  2 is

    desired to establish a trade-off between control performance and robustness [55].

    A PI controller, acting directly on  d, has been designed to track the inductor current   iLREF    since

    GiLd  exhibits a minimum phase behavior. The inductor current PI controller was designed by means

    of the root-locus technique, adopting following design specifications: damping factor  ζ  equal to 0.707

    and a 20kH z  closed loop bandwidth. The designed PI controller transfer function GC iL (s) is given by

    the equation (35). These PI controller design specifications ensure:   (a)  zero steady-state error and a

    satisfactory reference tracking for frequencies below 20kH z  observed on transfer function T iLiLREF  in

    Figure 13.   (b) Effective disturbance rejection for both input voltage  vg  and current source  io   variations

    observed on transfer functions T iLvg  and T iLio  in Figure 13, respectively.  (c) A M s  = 1.2.

    GC iL (s) = 1.27s + 55218

    s  (35)

    Figure 13.  Inner current control loop transfer functions.

    A PI controller, that provides the setpoint of the inductor current control loop, has been designed to

    regulate vo  since the GvoiLREF  (s) transfer function given by the equation (9) exhibits a minimum phase

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    23/30

     Energies 2016, xx   23

    behavior. The output voltage PI controller was designed by means of the root-locus technique, adopting

    following design specifications: damping factor ζ  equal to 0.707  and a  5kH z  closed loop bandwidth.

    The designed PI controller transfer function  GC vo (s) is given by the equation (36). These PI controller

    design specification ensure:   a.  zero steady-state error observed on transfer function  T vovoREF  in Figure14.   b.  Effective disturbance rejection for the current source io  variations observed on transfer function

    T voio  in Figure 14.  c.  A  M s = 1.2.

    GC vo (s) = 0.07994s + 235.1

    s  (36)

    Figure 14.  Outer output voltage control loop transfer functions.

    5.2. Closed-Loop System Performance Verification

    The non-ideal designed Boost DC-DC converter with its control structure has been implemented in

    PSIM to assess the closed-loop system performance. Figure  15a shows the closed-loop behavior at unit

    steps of  io  around the operating point corresponding to the full load. Two current source io  unit steps

    were applied to evaluate the control structure performance. First unit step was applied at  t   = 10ms

    and for 10ms, then the current source returns to its rated value io  = 0. Second unit step was applied at

    t  = 30ms and for 10ms, then the current source returns to its rated value  io   = 0. From Figure 15a is

    observed a satisfactory tracking of  iLREF   provided by the outer PI voltage controller and a satisfactory

    regulation of  vo to reject the load disturbances depicted as changes in io.

    Figure 15b shows the closed-loop behavior at unit steps of the input voltage vg. Two vg unit steps were

    applied to evaluate the control structure capabilities to regulate vo and to evaluated the boost capabilities

    of the designed Boost DC-DC converter. Firs unit step was applied at  t = 10ms and for 10ms. This first

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    24/30

     Energies 2016, xx   24

    unit step was equal to vg  = −5V , i.e., the final value of the input voltage was  vg  = 30 that corresponds

    with its lower boundary. Second unit step was applied at  t  = 30ms and for 10ms. This second unit step

    was equal to vg  = +5V , i.e., the final value of the input voltage was  vg  = 40V  that corresponds with its

    upper boundary. From Figure 15b is observed a satisfactory tracking of  iLREF  provided by the outer PIvoltage controller and a satisfactory regulation of the  vo  to changes in vg. It is important to remark that

    under the worst condition for input voltage  vg, the Boost DC-DC convert was able to keep the output

    voltage in its rated value.

    Finally, Figure   15c shows the closed-loop behavior at random unit steps of both   io   and  vg. This

    unit steps were applied such that the designed control structure performance could be evaluated against

    any random disturbance. From Figure 15c is possible to see that the designed control structure has

    a satisfactory performance against multiple disturbances within specified design requirements for the

    Boost DC-DC converter in Table 1.

    Figure 15.  Closed-loop behavior at unit steps system disturbances.

    5.3. System Operating Requirements Verification

    Figure 16 shows:  (a) P in, P out, and η  and  (b)  iL and vo, when case c  of Figure 15 is considered.

    From Figure 16a is seen that  P out  never exceeds the maximum admissible output power and it is

    always less that P in. It is also seen that the Boost DC-DC converter is never 100% efficient. Furthermore,

    from Figure 16a is seen that only between t = 0.01s and t = 0.02s the efficiency is below 0.9. However,

    in this time interval vg  is equal to 20V , i.e., in this time interval the Boost DC-DC converter operates in

    a not considered condition in Table 1. Accordingly, the designed Boost DC-DC converter satisfies both

    power and efficiency requirements.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    25/30

     Energies 2016, xx   25

    Figure 16.  (a) Instantaneous Power and efficiency verification. (b) Ripples verification.

    Figure 16b shows iL and vo, a zoom was made for the worst simulated system condition. From Figure

    16b is seen that even in the worst iL and vo condition, ripples are below to 1%. Accordingly, the designed

    Boost DC-DC converter satisfies both iL and vo ripples condition.

    In conclusion, Figure 16 shows that the Boost DC-DC converter system operating requirements given

    in Table 1 are satisfied even in the worst simulated case ( see Figure  15c).

    6. Conclusions

    In this paper, a non-ideal Boost DC-DC converter was dynamically modelled and analized in

    steady-state. In contrast with several adopted topologies to study the Boost DC-DC converter, in this

    work, parasitic resistances in the passive elements (L  and C ) were included in order to have a more

    realistic approach of the system.

    Based on the circuital configurations of the Boost DC-DC converter, Kirchhoff’s law was applied to

    obtain a non-linear dynamical model. The obtained non-linear dynamical model was composed of a set

    of coupled non-linear differential equations in function of state variables, input variables, and system

    parameters. The obtained dynamical model was suitable to describe dynamical behavior of the system

    and to derive the Boost DC-DC converter steady-state model.

    Steady-state analysis was composed of:   (a) M (D) expression derivation,  (b)  ∆iL and  ∆vC  analysis,

    (c) losses effect analysis and η  expression derivation, and (d)  conditions analysis of CCM and DCM.

    From expression of  M (D), it was possible to find the maximum Boost DC-DC converter conversion

    ratio in function of losses. An important conclusion from this result was that converter has a restricted

    conversion ratio, which depends not only on losses, but also on both load and D

    . ∆iL

     and ∆vC 

      analysis

    allowed to find lower boundaries to select  L  and  C  values to design the converter such that maximum

    ripples can be guaranteed. Losses effect analysis allowed to obtain an expression to the system efficiency

    in function of losses. From this expression, it was possible to conclude that the efficiency converter

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    26/30

     Energies 2016, xx   26

    depends not only on losses, but also on both load and  D. Moreover, it was concluded that the Boost

    DC-DC convert never reaches an efficiency equal to  100%. Finally, conditions analysis of CCM and

    DCM allowed to find boundaries such that the Boost DC-DC converter can be designed to operate in

    CCM or DCM.Main features of dynamical modeling and steady-state analysis were: (a)  developed dynamical model

    allowed to represent both high and low frequency converter dynamics and allowed to carry out the

    non-ideal Boost DC-DC converter steady-state analysis, which allowed to derive the expression of M (D)

    that showed the converter conversion ratio limitations;  (b)  from the converter steady-state model, it was

    possible to construct a suitable circuital converter representation that depicted the relations between DC

    voltages and currents, which allowed to derive an Boost DC-DC converter efficiency expression that

    showed the converter efficiency limits.

    Based on non-linear dynamical model and steady-state analysis, it was possible to design a Boost

    DC-DC converter such that a set of operating requirements are achieved in the entire operating range,even if bounded disturbances appear.

    Finally, CMC structure was implemented for the designed Boost DC-DC converter. PI controllers

    were tuning by means of root-locus controller design method. Finally, designed Boost DC-DC converter

    was implemented in PSIM and system operating requirements were verified.

    Acknowledgments

    Authors wish to thank the support offered by the “Estrategia de sostenibilidad 2014 − 2015" and the

    “Beca Estudiante Instructor" of Universidad de Antioquia (Colombia).

    Author Contributions

    Jorge H. Urrea-Quintero carried out most of the work presented here, Nicolás Muñoz-Galeano was

    the advisor of this work, and Lina M. Gómez had a relevant contribution with her extensive knowledge

    about systems theory.

    Conflicts of Interest

    The authors declare no conflict of interest.

    References

    1. Yousefzadeh, V.; Wang, N.; Popovic, Z.; Maksimovic, D. A digitally controlled DC/DC converter

    for an RF power amplifier.  Power Electronics, IEEE Transactions on 2006, 21, 164–172.

    2. Naayagi, R.; Forsyth, A. Bidirectional DC-DC converter for aircraft electric energy storage

    systems. Power Electronics, Machines and Drives (PEMD 2010), 5th IET International

    Conference on, 2010, pp. 1–6.

    3. Naayagi, R.; Forsyth, A.; Shuttleworth, R. High-Power Bidirectional DC-DC Converter forAerospace Applications.  Power Electronics, IEEE Transactions on 2012, 27 , 4366–4379.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    27/30

     Energies 2016, xx   27

    4. Chiu, H.J.; Lin, L.W. A bidirectional DC-DC converter for fuel cell electric vehicle driving

    system.  Power Electronics, IEEE Transactions on 2006, 21, 950–958.

    5. Sharma, R.; Gao, H. Low cost high efficiency DC-DC converter for fuel cell powered auxiliary

    power unit of a heavy vehicle.  Power Electronics, IEEE Transactions on 2006, 21, 587–591.6. Camara, M.; Gualous, H.; Gustin, F.; Berthon, A.; Dakyo, B. DC/DC Converter Design for

    Supercapacitor and Battery Power Management in Hybrid Vehicle Applications - Polynomial

    Control Strategy.  Industrial Electronics, IEEE Transactions on  2010, 57 , 587–597.

    7. de Souza Oliveira, D., J.; Barbi, I. A three-phase ZVS PWM DC/DC converter with asymmetrical

    duty cycle for high power applications.   Power Electronics, IEEE Transactions on   2005,

    20, 370–377.

    8. Oliveira, D.; Barbi, I. A three-phase ZVS PWM DC/DC converter with asymmetrical duty

    cycle associated with a three-phase version of the hybridge rectifier.   Power Electronics, IEEE 

    Transactions on 2005, 20, 354–360.9. Liang, Z.; Guo, R.; Li, J.; Huang, A. A High-Efficiency PV Module-Integrated DC/DC Converter

    for PV Energy Harvest in FREEDM Systems.   Power Electronics, IEEE Transactions on  2011,

    26 , 897–909.

    10. Wang, Z.; Li, H. An Integrated Three-Port Bidirectional DC-DC Converter for PV Application

    on a DC Distribution System.  Power Electronics, IEEE Transactions on  2013, 28 , 4612–4624.

    11. Vlad, C.; Rodriguez-Ayerbe, P.; Godoy, E.; Lefranc, P. Advanced control laws of DC-DC

    converters based on piecewise affine modelling. Application to a stepdown converter.   Power 

     Electronics, IET  2014, 7 , 1482–1498.

    12. Liu, J.; Hu, J.; Xu, L. Dynamic Modeling and Analysis of Z Source Converter – Derivation of 

    AC Small Signal Model and Design-Oriented Analysis.   Power Electronics, IEEE Transactions

    on 2007, 22, 1786–1796.

    13. Arango, E.; Ramos-Paja, C.A.; Calvente, J.; Giral, R.; Serna, S. Asymmetrical Interleaved

    DC/DC Switching Converters for Photovoltaic and Fuel Cell ApplicationsŮPart 1: Circuit

    Generation, Analysis and Design.   Energies 2012, 5, 4590.

    14. Arango, E.; Ramos-Paja, C.A.; Calvente, J.; Giral, R.; Serna-Garces, S.I. Asymmetrical

    Interleaved DC/DC Switching Converters for Photovoltaic and Fuel Cell ApplicationsŮPart 2:

    Control-Oriented Models.   Energies 2013, 6 , 5570.15. Chiniforoosh, S.; Jatskevich, J.; Yazdani, A.; Sood, V.; Dinavahi, V.; Martinez, J.; Ramirez, A.

    Definitions and Applications of Dynamic Average Models for Analysis of Power Systems. Power 

     Delivery, IEEE Transactions on  2010, 25, 2655–2669.

    16. Liang, T.; Tseng, K. Analysis of integrated boost-flyback step-up converter.   Electric Power 

     Applications, IEE Proceedings - 2005, 152, 217–225.

    17. Davoudi, A.; Jatskevich, J.; De Rybel, T. Numerical state-space average-value modeling of PWM

    DC-DC converters operating in DCM and CCM.  Power Electronics, IEEE Transactions on 2006,

    21, 1003–1012.

    18. Wai, R.J.; Shih, L.C. Design of Voltage Tracking Control for DC-DC Boost Converter Via Total

    Sliding-Mode Technique.  Industrial Electronics, IEEE Transactions on  2011, 58 , 2502–2511.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    28/30

     Energies 2016, xx   28

    19. Galigekere, V.; Kazimierczuk, M. Analysis of PWM Z-Source DC-DC Converter in CCM for

    Steady State.  Circuits and Systems I: Regular Papers, IEEE Transactions on  2012, 59, 854–863.

    20. Thao, N.; Thang, T.; Ganeshkumar, P.; Park, J.H. Steady-state analysis of the boost converter for

    renewable energy systems. Power Electronics and Motion Control Conference (IPEMC), 20127th International, 2012, Vol. 1, pp. 158–162.

    21. Hu, G.Y.; Li, X.; Luan, B.Y. A Generalized Approach for the Steady-State Analysis of 

    Dual-Bridge Resonant Converters.   Energies 2014, 7 , 7915.

    22. Davoudi, A.; Jatskevich, J.; Chapman, P.; Bidram, A. Multi-Resolution Modeling of Power

    Electronics Circuits Using Model-Order Reduction Techniques.  Circuits and Systems I: Regular 

    Papers, IEEE Transactions on  2013, 60, 810–823.

    23. Beldjajev, V.; Roasto, I. Efficiency and Voltage Characteristics of the Bi-Directional Current

    Doubler Rectifier.  Przegl ąd Elektrotechniczny 2012, 88 , 124–129.

    24. Mazumder, S.; Nayfeh, A.; Boroyevich, D. Theoretical and experimental investigation of thefast- and slow-scale instabilities of a DC-DC converter.   Power Electronics, IEEE Transactions

    on 2001, 16 , 201–216.

    25. El Aroudi, A.; Giaouris, D.; Ho-Ching Iu, H.; Hiskens, I. A Review on Stability Analysis

    Methods for Switching Mode Power Converters.   Emerging and Selected Topics in Circuits and 

    Systems, IEEE Journal on 2015, 5, 302–315.

    26. Geyer, T.; Papafotiou, G.; Frasca, R.; Morari, M. Constrained Optimal Control of the Step-Down

    DC-DC Converter.  Power Electronics, IEEE Transactions on 2008, 23, 2454–2464.

    27. Mariethoz, S.; Almer, S.; Baja, M.; Beccuti, A.; Patino, D.; Wernrud, A.; Buisson, J.; Cormerais,

    H.; Geyer, T.; Fujioka, H.; Jonsson, U.; Kao, C.Y.; Morari, M.; Papafotiou, G.; Rantzer, A.;

    Riedinger, P. Comparison of Hybrid Control Techniques for Buck and Boost DC-DC Converters.

    Control Systems Technology, IEEE Transactions on  2010, 18 , 1126–1145.

    28. Luo, F.L.; Ye, H. Small Signal Analysis of Energy Factor and Mathematical Modeling for Power

    DC ndash;DC Converters.  Power Electronics, IEEE Transactions on 2007, 22, 69–79.

    29. Wu, X.; Xiao, G.; Lu, Y.; Chen, F.; Lu, D. Discrete-time modeling and stability analysis of 

    grid-connected inverter based on equivalent circuit. Applied Power Electronics Conference and

    Exposition (APEC), 2015 IEEE, 2015, pp. 1222–1226.

    30. Mahery, H.M.; Babaei, E. Mathematical modeling of buck  ˝ Uboost dc

     ˝ Udc converter and

    investigation of converter elements on transient and steady state responses.  International Journal

    of Electrical Power & Energy Systems  2013, 44, 949 – 963.

    31. Davoudi, A.; Jatskevich, J. Parasitics Realization in State-Space Average-Value Modeling of 

    PWM DC-DC Converters Using an Equal Area Method.   Circuits and Systems I: Regular Papers,

     IEEE Transactions on 2007, 54, 1960–1967.

    32. Qiu, Y.; Chen, X.; Zhong, C.; Qi, C. Uniform Models of PWM DC-DC Converters

    for Discontinuous Conduction Mode Considering Parasitics.   Industrial Electronics, IEEE 

    Transactions on 2014, 61, 6071–6080.

    33. Chan, C.Y. A Nonlinear Control for DC-DC Power Converters.   Power Electronics, IEEE 

    Transactions on 2007, 22, 216–222.

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    29/30

     Energies 2016, xx   29

    34. Urtasun, A.; Sanchis, P.; Marroyo, L. Adaptive Voltage Control of the DC/DC Boost Stage in

    PV Converters With Small Input Capacitor.   Power Electronics, IEEE Transactions on  2013,

    28 , 5038–5048.

    35. Roshan, Y.; Moallem, M. Control of Nonminimum Phase Load Current in a Boost ConverterUsing Output Redefinition.  Power Electronics, IEEE Transactions on  2014, 29, 5054–5062.

    36. Trejos, A.; Gonzalez, D.; Ramos-Paja, C.A. Modeling of Step-up Grid-Connected Photovoltaic

    Systems for Control Purposes.   Energies 2012, 5, 1900.

    37. Kapat, S.; Krein, P. Formulation of PID Control for DC-DC Converters Based on Capacitor

    Current: A Geometric Context.  Power Electronics, IEEE Transactions on  2012, 27 , 1424–1432.

    38. Kabalo, M.; Paire, D.; Blunier, B.; Bouquain, D.; Godoy Simões, M.; Miraoui, A. Experimental

    evaluation of four-phase floating interleaved boost converter design and control for fuel cell

    applications.  Power Electronics, IET  2013, 6 , 215–226.

    39. Mapurunga Caracas, J.; De Carvalho Farias, G.; Moreira Teixeira, L.; De Souza Ribeiro,L. Implementation of a High-Efficiency, High-Lifetime, and Low-Cost Converter for an

    Autonomous Photovoltaic Water Pumping System.   Industry Applications, IEEE Transactions

    on 2014, 50, 631–641.

    40. Evzelman, M.; Ben-Yaakov, S. Simulation of Hybrid Converters by Average Models.   Industry

     Applications, IEEE Transactions on 2014, 50, 1106–1113.

    41. Gezgin, C.; Heck, B.S.; Bass, R.M. Integrated design of power stage and controller for switching

    power supplies. Computers in Power Electronics, 1996., IEEE Workshop on, 1996, pp. 36–44.

    42. Pomar, M.; Gutierrez, G.; de Prada Moraga, C.; Normey Rico, J. Integrated design and control

    applied to a buck boost converter. Control Conference (ECC), 2007 European, 2007, pp.

    948–954.

    43. Mariethoz, S.; Almer, S.; Morari, M. Integrated control and circuit design; an optimization

    approach applied to the buck converter. American Control Conference (ACC), 2010, 2010, pp.

    3305–3310.

    44. Erickson, R.W.; Maksimovic, D.   Fundamentals of power electronics; Springer Science &

    Business Media, 2007.

    45. Hoagg, J.; Bernstein, D. Nonminimum-phase zeros - much to do about nothing - classical control

    - revisited part II.  Control Systems, IEEE  2007, 27 , 45–57.46. Aguiar, A.; Hespanha, J.; Kokotovic, P. Path-following for nonminimum phase systems removes

    performance limitations.  Automatic Control, IEEE Transactions on 2005, 50, 234–239.

    47. Su, W.; Qiu, L.; Chen, J. Fundamental performance limitations in tracking sinusoidal signals.

     Automatic Control, IEEE Transactions on  2003, 48 , 1371–1380.

    48. Alvarez-Ramirez, J.; Espinosa-Pérez, G.; Noriega-Pineda, D. Current-mode control of DC-DC

    power converters: a backstepping approach.   International Journal of Robust and Nonlinear 

    Control 2003, 13, 421–442.

    49. Qiu, Y.; Liu, H.; Chen, X. Digital Average Current-Mode Control of PWM DC-DC Converters

    Without Current Sensors.  Industrial Electronics, IEEE Transactions on  2010, 57 , 1670–1677.

    50. El Aroudi, A.; Alarcon, E.; Rodriguez, E.; Leyva, R.; Villar, G.; Guinjoan, F.; Poveda, A.

    Ripple Based Index for Predicting Fast-Scale Instability of DC-DC Converters in CCM and

  • 8/18/2019 Boost Paper Unified Rev NMG 24-02-2016

    30/30

     Energies 2016, xx   30

    DCM. Industrial Technology, 2006. ICIT 2006. IEEE International Conference on, 2006, pp.

    1949–1953.

    51. Ping, L.; Xin, M.; Bo, Z.; Zhao-ji, L. Analysis of the Stability and Ripple of PSM Converter in

    DCM by EB Model. Communications, Circuits and Systems, 2007. ICCCAS 2007. InternationalConference on, 2007, pp. 1240–1243.

    52. Liu, S.L.; Liu, J.; Mao, H.; qing Zhang, Y. Analysis of Operating Modes and Output

    VoltageRipple of Boost DC - DC Convertersand Its Design Considerations.  Power Electronics,

     IEEE Transactions on 2008, 23, 1813–1821.

    53. Chen, Z.; Gao, W.; Hu, J.; Ye, X. Closed-Loop Analysis and Cascade Control of a Nonminimum

    Phase Boost Converter.  Power Electronics, IEEE Transactions on  2011, 26 , 1237–1252.

    54. Louganski, K.; Lai, J.S. Current Phase Lead Compensation in Single-Phase PFC Boost

    Converters With a Reduced Switching Frequency to Line Frequency Ratio.   Power Electronics,

     IEEE Transactions on 2007, 22, 113–119.55. Vilanova, R.; Alfaro, V.M. Control {PID} robusto: Una visión panorámica.   Revista

     Iberoamericana de Automática e Informática Industrial {RIAI} 2011, 8 , 141 – 158.

    c   2016 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article

    distributed under the terms and conditions of the Creative Commons Attribution license

    (http://creativecommons.org/licenses/by/4.0/).