A HIGH VOLTAGE DC-DC CONVERTER FOR APPLICATIONS IN HARSH ENVIRONMENTS By Matthew Lee Gebben A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Electrical Engineering 2012 ABSTRACT A HIGH VOLTAGE GAIN DC-DC CONVERTER FOR APPLICATIONS IN HARSH ENVIRONMENTS By Matthew Lee Gebben The constant evolution of the use of power electronics today is bringing sensitive electrical devices into harsher environments than ever before. Today, the energy and automotive industries, among others, are seeking to utilize power electronics to increase the performance and efficiency of the systems in use. This often involves using converters in close proximity to sources of heat, such as internal combustion engines or photovoltaic panels. These situations also usually require small size and low weight. The result is a need for a high efficiency converter capable of operating in high temperatures with minimal size and weight. This paper seeks to explore the options available to achieve this when considering a typical photovoltaic system. In order to meet the above conditions while operating as a micro-converter attached to the solar panels, an additional requirement for high voltage gain is added. This limits the number of existing topologies capable of being used, and has resulted in the development of a new design, detailed in this paper, that seeks to meet all of the aforementioned constraints. To make the case of the new converter, other topologies are investigated in terms of their features and costs in order to determine their best and worst attributes. This involves discussion of their operation, calculation of their relative costs, and simulations to confirm their behavior. This is followed by a look at the NX dc-dc converter, the new topology, which includes a basic investigation into it operation, followed by an analysis of its behavior and relative cost, plus simulated and experimental results. DEDICATION Dedicated to my mother for her constant support and encouragement Dawn L. Gebben May 20, 1951 – July 12, 2012 iii ACKNOWLEDGEMENTS I would like to express my appreciation for the continual support, encouragement, and love of my parents during my time in graduate school. If not for everything they did, none of this would have happened. I would also like to thank Dr. Fang Z. Peng for his support and guidance during the past few years. Without his help, I never would have been able to conduct this research. Finally, I thank my colleagues at the Power Electronics and Motor Drives Laboratory. I received a great deal of help from everyone, and I can scarcely recall everything they taught me in such a short time. iv TABLE OF CONTENTS List of Figures………………………………………………………………………………….vii 1. Introduction………………………………………………………………………………1 2. Background………………………………………………………………………..……..4 2.1. The DC-DC Boost Converter……………………………………….….………..4 2.2. The Isolated Full-Bridge DC-DC Converter Family…………………...…...…7 2.3. The Flying Capacitor DC-DC Converter……………………………...…...…19 2.4. The Multilevel Modular Capacitor Clamped DC-DC Converter…...………41 3. The NX DC-DC Converter……………………………………………...……………...57 3.1. NX DC-DC Converter Details………………………………….………………57 3.2. NX Converter Analysis…………………………………………………………65 3.3. NX Converter ZCS Possibilities……………………………...………………..72 3.4. 8X Converter Experimental Results………………………………………..…79 4. Conclusions…………………………………………………………...…………………88 Bibliography v LIST OF FIGURES Figure 1. System Configuration………………………………………………………………...3 Figure 2. DC-DC Boost Converter……………………………………………………………..4 Figure 3. Boost Converter Switching Waveforms…………………………………………….5 Figure 4. Simulation Results for Boost Converter………………………………….…………6 Figure 5. Isolated Full-Bridge DC-DC Buck Converter………………………………………8 Figure 6. Isolated Buck Converter Model………………………………………………..……9 Figure 7. Isolated Full-Bridge Buck Converter Waveforms………………………..……….10 Figure 8. Simulated Full-Bridge Buck Converter………………………………………..…..12 Figure 9. Isolated Full-Bridge DC-DC Boost Converter………………………….........……13 Figure 10. Isolated Boost Converter Model…………………………………………………..14 Figure 11. Isolated Full-Bridge Boost Converter Waveforms……………………..…….….15 Figure 12. Simulated Full-Bridge Boost Converter……………………………………...…..17 Figure 13. Flying Capacitor DC-DC Converter………………………………….........……..20 Figure 14. Flying Capacitor Switching State I…………………………………….……..…..21 Figure 15. Flying Capacitor Switching State II…………………………………….………...22 Figure 16. Flying Capacitor Switching State III…………………………………….……….23 Figure 17. Flying Capacitor Switching State IV………………………………….………….24 Figure 18. Simulated Flying Capacitor DC-DC Converter Waveforms…………..………..26 Figure 19. The 3X DC-DC Converter…………………………………………………….…..29 Figure 20. 3X Converter 1X Switching State……………………………………….………..31 Figure 21. 3X Converter 2X Switching State I……………………………………………….32 vi Figure 22. 3X Converter 2X Switching State II……………………………………..……….33 Figure 23. 3X Converter 3X Switching State I……………………………………………….34 Figure 24. 3X Converter 3X Switching State II……………………………………………...35 Figure 25. 3X Converter 3X Switching State III………………………………………….….36 Figure 26. 1X to 2X Transition with 10 nH Stray Inductance…………………………..…..37 Figure 27. 1X to 2X Transition with 100 nH Stray Inductance……………………………..38 Figure 28. 2X to 3X Transition with 10 nH Stray Inductance……………………………....40 Figure 29. 2X to 3X Transition with 100 nH Stray Inductance……………………………..40 Figure 30. The MMCCC………………………………………………………………..……..42 Figure 31. The MMCCC Module…………………………………………………………..…42 Figure 32. MMCCC Switching State I………………………………………………………..43 Figure 33. MMCCC Switching State II……………………………………………..…….….44 Figure 34. MMCCC Simulation Results……………………………………………….……..47 Figure 35. ZCS-MMCCC……………………………………………………………….……..48 Figure 36. ZCS-MMCCC Switching State I…………………………………………….……49 Figure 37. ZCS-MMCCC Switching State II…………………………………..…….………50 Figure 38. ZCS-MMCCC C4 Voltage Example………………………………….…….…….53 Figure 39. ZCS-MMCCC Simulation Output Voltage and Inductor Currents………...…54 Figure 40. ZCS-MMCCC Simulation Capacitor Voltages…………………………….....…54 Figure 41. ZCS-MMCCC Input Current…………………………………………………….55 Figure 42. ZCS-MMCCC 4-Phase Input Current………………………………….………..56 Figure 43. The 6X DC-DC Converter………………………………………………..…….…58 Figure 44. The Modular 6X DC-DC Converter……………………………………………...58 vii Figure 45. The NX Converter Module………………………………………………………..59 Figure 46. 6X Switching State I……………………………………………………...………..60 Figure 47. 6X Switching State II…………………………………………..…………………..61 Figure 48. 6X Simulation Input and Output Voltages……………………………...……….63 Figure 49. 6X Simulation Capacitor Voltages…………………………………………..……64 Figure 50. The ZCS 6X Converter……………………….…………………………………...73 Figure 51. ZCS 6X Converter Switching State I……………………………...……….……..74 Figure 52. CS 6X Converter Switching State II……………………………….……….…….74 Figure 53. ZCS 6X Simulation Input and Output Voltages………………….………..…….77 Figure 54. ZCS 6X Simulation Inductor Currents………………………...………….……..77 Figure 55. ZCS 6X Simulation Capacitor Voltages………………………………...………..78 Figure 56. The 1-kW 8X prototype (For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this thesis)………...…..81 Figure 57. 8X Prototype Converter Waveforms with Vin=10 V (Channel 1: Vgs, 20 V/div; Channel 2: Vds, 10 V/div; Channel 3: Vds, 10 V/div; Channel 4: IIs, 5 A/div)…………...….82 Figure 58. Simulated 8X IIs Current Waveform………………………………………..…83 Figure 59. 8X Prototype Converter Waveforms with Vin=10 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 25 V/div; Channel 3: Vin, 5 V/div; Channel 4: IIs, 5 A/div)………………84 Figure 60. 8X Prototype Converter Waveforms with Vin=18 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 50 V/div; Channel 3: Vin, 10 V/div; Channel 4: IIs, 10 A/div)…………....85 Figure 61. 8X Prototype Converter Waveforms with Vin=20.4 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 50 V/div; Channel 3: Vin, 10 V/div; Channel 4: IIs, 10 A/div)…………....85 Figure 62. 8X Prototype Converter Waveforms with Vin=35 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 100 V/div; Channel 3: Vds, 25 V/div; Channel 4: IIs, 10 A/div)…………..86 viii Figure 63. 8X Prototype Converter Waveforms with Vin=37 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 100 V/div; Channel 3: Vds, 25 V/div; Channel 4: IIs, 10 A/div)…………..86 Figure 64. Efficiency vs. Output Power for the 8X Prototype………………………...…….87 ix 1. Introduction Today, power electronics are making an ever greater impact on all areas of industry. As energy costs climb and greater control is desired, the drive for greater efficiency fuels the growth of the power electronics field. In the automotive industry, vehicles are increasingly reliant on power electronics for supplying their drive-trains and other systems; in the aerospace industry, the need to maximize fuel economy has lead to a desire for both maximized efficiency and minimized weight. The energy sector is also seeing rapid growth in the use of power electronics. The proliferation of solar and wind power has lead to a need for more power electronics circuits operating on the grid than ever before. Solar power, in particular, raises issues in common with those in the other aforementioned industries. Like other areas, great importance is placed on efficiency, reliability, and environmental tolerance. Photovoltaic (PV) modules present an interesting set of design challenges. In many situations, the voltage supplied by photovoltaic modules is too low to be of much use by itself. Furthermore, the module only supplies direct current electricity, so it needs to be inverted for many systems. This state of affairs makes power electronics vital for the use of PV modules. PV modules, such as the Mitsubishi Electric PV-MJT250GB, have maximum output voltages of roughly 40 V [1]. Although it may be tempting to suggest placing multiple panels in series to achieve a higher voltage, the reliability would be adversely affected and the current capabilities would be low. From this point of view, it is necessary to place multiple panels in parallel to maximize reliability and current (power) capability. This necessitates the use of a high voltage gain dc-dc converter to boost the voltage to a more useful level. Considering this, several options are available; either one high power converter can be used to 1 convert the voltage from all the cells together, or multiple smaller converters can be used with each module to boost the voltage. Furthermore, the task of regulating the voltage can fall to either the dc-dc converter, or to the inverter. These choices make for a high degree of freedom when designing power electronics systems for PV sources. This means some assumptions must be made in order to focus on a more manageable set of options. One aspect of design that engineers should always consider is cost. In situations involving mass production or the installation of a large number of facilities, the cost of the electronics used must be minimized. With this in mind, the cost of employing one high power converter could easily outweigh the price of using multiple low power converters. Furthermore, the use of small dc-dc converters coupled to each PV module would increase the system’s overall reliability; even if one converter was to fail, the dc-bus voltage would still be useable. Cost and reliability make an excellent case for the use of low power dc-dc micro-converters fixed to each PV panel. However, this still leaves many options available. From the standpoint of cost and reliability, the traditional topologies are easily obtained on the commercial market and have been used for many years. Furthermore, most of these designs are capable of regulating the voltage of the dc bus. Nevertheless, as will be shown, they may not be the best choices for this situation, where high voltage gain and high temperature capabilities are a must. More recent topologies, such as the many switched-capacitor designs out there, are capable of high voltage gain and have designs well-suited for operation at high temperatures. However, the number of components can easily be larger, and the option of voltage regulation is not always available. Before getting into the finer details, it is important to examine the overall situation. 2 Figure 1. System Configuration Consider the general situation described in Figure 1. The PV module is connected to a dc-dc converter, which is then connected to the load through an inverter. Here, due to the nature of solar cells, bi-directional power flow is unnecessary. Ignoring the inverter, the options for the dc-dc converter have already been made clear. To make a clear progression to the latest converter technologies, it is important to start by introducing the older circuits. In this fashion, the case for each design can be more easily made. With that in mind, the oldest and most basic design to consider is obviously the traditional boost converter. From this, further exploration will be done by examining the traditional isolated topologies, such as the isolated full-bridge converter, which is likely one of the most popular designs currently in use. Having looked at the older designs, the switched capacitor designs can be explored; this study can begin with the flying capacitor topology and can work its way up to the multilevel modular capacitor-clamped converter (MMCCC). This provides an excellent base from which the merits of the new topology can be expounded upon. As alluded to before, a new switched-capacitor dc-dc converter design has been developed, and its positive and negative features will be detailed in later sections. The design seeks to achieve both the high voltage gain and high temperature operation capabilities without many of the drawbacks suffered by other designs. 3 2. Background The situation as described in Figure 1 leaves many options available. Focusing on the dcdc converter, voltage regulation, isolation, and the use of magnetic components are factors to consider. Nevertheless, the inverter can, to a degree, regulate the voltage output of the system, and the necessity of galvanic isolation is debatable. Before a complete discussion of a new dc-dc converter topology is possible, a thorough understanding of the other options is necessary. To this end, it makes sense to begin with one of the oldest and most basic of the dc-dc converters: the boost converter. 2.1. The DC-DC Boost Converter Consisting of only two switching devices, an inductor and a capacitor, the dc-dc boost converter is the most basic converter to consider for this application. Figure 2. DC-DC Boost Converter In Figure 2 the dc-dc boost converter circuit can be seen. Although depicted using MOSFETs, any suitable switch would work. This converter also lacks isolation, although that is only a problem if isolation is required. As suggested by the name, the converter boosts the input 4 voltage to any higher voltage. In theory, the voltage gain could even be infinity, but the truth is much less phenomenal. In this converter, Q1 and Q2 function as a complementary pair; while Q1 is on, Q2 is off and vice-versa. The time that Q1 is on verses the total switching period is referred to as the duty cycle, D. It is this D that controls the output voltage. The switching behavior along with the inductor waveforms, can be seen in Figure 3. Figure 3. Boost Converter Switching Waveforms When S1 is equal to 1, Q1 is on and Q2 is off; likewise, when S2 equals 1, Q2 is on and Q1 is off. So the voltage across the inductor is equal to the input voltage while Q1 conducts, and is equal to the input voltage minus the output voltage while Q2 is conducting. Since the voltage across the inductor must average to zero, the following situation occurs. 5 VL = 0 = Vin D + (Vin − Vout )(1 − D) ⇒ 0 = Vin D + Vin − Vout − Vin D + Vout D ⇒ 0 = Vin − Vout + Vout D = Vin + ( D − 1)Vout ⇒ −Vin = ( D − 1)Vout ⇒ (2.1.1) Vin = (1 − D)Vout ⇒ Vout 1 = Vin 1 − D As seen in (2.1.1), the voltage gain theoretically increases as D goes from 0 to 1; (2.1.1) approaches infinite as D approaches 1. To better illustrate an example of this, a simulation was performed using Saber Simulator. The switching frequency was 100 kHz, using a 50 µH inductor, a 100 µF output capacitor, and a 20 load resistor. The input voltage was 20 V, with a duty cycle of 40%. The results can be seen in Figure 4. Figure 4. Simulation Results for Boost Converter The results were slightly off from their ideal values, due to losses in the switches on-resistances. Nevertheless, the results match up very nicely to expectations for these conditions. However, as this suggests, losses in the switches and inductor would greatly limit the actual capabilities of the converter [2]. All analysis will be based of the assumption that the converter is in continuous- 6 conduction mode (CCM). Consider the total device stress ratio of the boost converter, a relative measure of the cost of the semiconductor devices. First, assume the converter is lossless (Pin=Pout), and set Pout, Vout and Iout to 1 p.u. (per unit). The result is that: P 1 p.u. 2 RS = 2 × Vout × Iin = 2 × out = 2 × 1 p.u. × = p.u. 1− D 1− D 1− D (2.1.2) In (2.1.2), RS is the total device stress ratio using the average values of the circuit’s voltages and currents. There are two switches, and each must be able to block the output voltage and pass the input current. This led to the above result, which shows a major flaw in this converter’s use in this application; as the voltage gain increases, the device stresses also increase, asymptotically approaching infinity as D goes to 1. Even for a relatively low duty cycle, like 0.5, the stress ratio would already be 4 p.u. This converter would not be practical for applications requiring very high voltage gains, such as photovoltaic arrays. Furthermore, the switch utilization factor of the boost converter, defined as the output voltage divided by the switch power rating, is: Pout 1 p.u. = = (1 − D) 1 p.u. PS 1− D (2.1.3) This is the same for each switch, and it is clear that the switch utilization becomes increasingly poor as the voltage gain is increased. This again suggests that the boost converter, despite its initially apparent high voltage gain capabilities, is, in fact, not suited to this application. 2.2. The Isolated Full-Bridge DC-DC Converter Family The isolated full-bridge dc-dc converter family consists of some of the most commonly employed topologies. These converters are simply slightly evolved versions of their older buck and boost converter cousins. As the name would suggest, these converters are isolated, through 7 the use of a transformer, and use a full-bridge configuration of four active switches. Whether or not the system’s inductor is before the active switches or after the rectifier determines the type of converter. Thus, there are two configurations to consider in this topology family: the voltage-fed buck converter and the current-fed boost converter. Of these two circuits, the isolated full-bridge buck converter is the most commonly employed. However, either circuit could be viewed as applicable for this PV application. The converter has galvanic isolation via the transformer, which also offers the possibility of high gain through the use of a large turn ratio. The operations of these converters vary slightly from their simpler predecessors. Starting with the more common of the two converters, the isolated full-bridge buck converter can be seen in Figure 5 below. This is basically just an extended version of the buck converter. Figure 5. Isolated Full-Bridge DC-DC Buck Converter The switching pattern of this converter is slightly different than that of its non-isolated cousin, however. This converter switches [Q1, Q4] and [Q2, Q3] in pairs, but it does not do so complementarily; instead, this circuit makes extensive use of the dead-time between the two 8 switching states. The topology’s operation can be better understood through the use of a roughly equivalent model. Seen in Figure 6, it is assumed that the transformer is nearly ideal; in other words, leakage inductances are not considered and the magnetizing inductance is very large. Figure 6. Isolated Buck Converter Model Here the magnetizing inductance, LM, is visible before the ideal transformer with its turn ratio of n. The voltage across the magnetizing inductance and the transformer itself is called vT, while the voltage seen across the rectifier output is called vs. It is this voltage, vs, that works to determine the inductor current, iL. During this analysis, it is assumed that the inductor current is always continuous, and that the diodes are ideal; no forward voltage drops across the diodes are considered. The switches are also considered ideal, and transformer characteristics such as hysteresis and saturation are never modeled. This greatly simplifies the analysis without much distortion from the actual function. To begin, note the eight waveforms graphed in Figure 7 below; these describe the action of the ideal converter in detail [3]. 9 Figure 7. Isolated Full-Bridge Buck Converter Waveforms Drawn with a 50 percent duty cycle in mind, these ideal waveforms show how the isolated buck converter functions. Note that the duty cycle itself is defined slightly differently from before, however; now the duty cycle, D, refers to the amount of time one pair of switches is turned on during one switching cycle, with this pair alternating each cycle. For instance, during the period 0>RC. Before continuing, notice the behavior of C3A. It seems to be mostly linear in both sections of the switching period, aside from its charging; this is because the load current is always running through the output stage capacitors, and this is what makes it behave differently. 3.2. NX Converter Analysis In order to get a practical analytical understanding of this converter, a closer look at Figure 46 and Figure 47 is needed. Because the circuits are symmetric, solutions for only one switching state can easily be applied to the other. Using Figure 46, it can be seen that it breaks down into three separate current loops. Making the safe assumption that the resistances and capacitors in each module are identical and that the output current is constant, some general equations describing the system can be written. Vin = R1i1(t ) + vC1B (t ) i1(t ) = C1 dvC1B dt Vin = R2 i2 (t ) − vC1A (t ) + vC 2 A (t ) i2 (t ) = C2 dvC 2 A dv = −C1 C1A dt dt Vin = R3 i3 (t ) − vC 2 B (t ) + vC 3B (t ) i3 (t ) = −C2 (3.2.1) (3.2.2) (3.2.3) (3.2.4) (3.2.5) dvC 2 B dv = C3 C 3B + I out dt dt (3.2.6) dvC 3 A dt (3.2.7) I out = −C3 65 In these equations, the subscripts in the resistances and the capacitances refer to the number of the loop in question. From these equations, it is easy to recognize that the currents and voltages will mostly follow the standard forms for RC circuits, but (3.2.5)-(3.2.7) throw off the simplicity of the standard form. In order to solve these equations, the boundary conditions have to be specified; however, in order to have the correct boundary conditions, the general solutions must be used in conjunction with things already known about the circuit, especially the average current through the capacitors during a given half-period. In order to shorten the equations, the boundary conditions are given the following designation: VCXA1 refers to the capacitor voltage of CXA at the beginning of the first switching state. VCXB2 refers to the voltage of CXB at the beginning of state II, where X refers to the module number. These values will be valuable for designing the converter. The generalized solutions for the previous equations are as follows: τ1 = R1C1 i1(t ) = (Vin − VC1B1 ) e−t /τ1 R1 vC1B (t ) = Vin (1 − e−t /τ1 ) + VC1B1 × e−t /τ1 (3.2.9) (3.2.10) C1C2 , τ 2 = R2C2 X C1 + C2 (3.2.11) (Vin + VC1A1 − VC 2 A1 ) e−t /τ 2 (3.2.12) C2 X = i2 (t ) = (3.2.8) R2 −t /τ 2 − 1) + V −t /τ 2 )     C2Vin (e 1 C1A1(C1 + C2e   vC1A (t ) =    −t /τ 2 ) C1 + C2   +C V   2 C 2 A1(1 − e  66 (3.2.13) −t /τ 2 ) + C V −t /τ 2 )     C1Vin (1 − e 1 1 C1A1(1 − e   vC1A (t ) =    −t /τ 2 + C ) C1 + C2   +V  2  C 2 A1(C1e  C3 X = C2C3 , τ 3 = R3C3 X C2 + C3 −t /τ 3   (C + C )(V + V 3 in C 2 B1 − VC 3B1)e  2  i3 (t ) =  R3 (C2 + C3 )  + I R C (1 − e−t /τ 3 )  out 3 2  1 (3.2.14) (3.2.15) (3.2.16) 2  C 2V  + C3 (Vin + VC 2 B1 − VC 3B1)e−t /τ 3  2 C 2 B1   −C ( I t + C V − C V  3 in 3 C 3B1) − C2 I out t 1  3 out  (3.2.17) vC 2 B (t ) = 2  +C C I R (1 − e−t /τ 3 ) + C C V (e−t /τ 3 − 1)  (C2 + C3 ) 2 3 in  2 3 out 3   −t /τ 3 + 1) + C C V −t /τ 3 )  2 3 C 3B1(1 − e  +C2C3VC 2 B1(e   C 2 ((1 − e−t /τ 3 )(V + V  in C 2 B1 − I out R3 )   2   −t /τ 3 ) + C 2V 1  +VC 3B1e 3 C 3B1 − C3 I out t  vC 3B (t ) = (C2 + C3 )2  +C2 (C3 (Vin + VC 2 B1 + VC 3B1) − I out t )     −t /τ 3   −C C (V + V   2 3 in C 2 B1 − VC 3B1)e  (3.2.18) I vC 3 A (t ) = VC 3 A1 − out t C3 (3.2.19) Due to the effect that the load has on the last stage’s capacitors, the equations became somewhat ugly. However, for the most part the boundary conditions are much simpler and, perhaps, are more useful. Using the fact that the averages of (3.2.9), (3.2.12), and (3.2.16) are all 2Iout, plus the assumption that the exponential terms die off after a half-period, the boundary conditions were accurately found to be: VC1A1 = VC1B 2 = Vin 67 (3.2.20) I T VC1B1 = VC1A2 = Vin − out sw C1 (3.2.21) I T VC 2 A1 = VC 2 B 2 = 2Vin − out sw CX 2 (3.2.22) I T VC 2 B1 = VC 2 A2 = 2Vin − out sw C1 (3.2.23) I T VC 3B1 = VC 3 A2 = 3Vin − out sw C1 2 C2 C3 Tsw 1 − + I out R3 ( I out Tsw (C2 + C3 ) − C2 Iout ) 2 C2C3 C2 + C3 (3.2.24)  2   C2 (Vin + VC 2 B1 − I out R3 )    Tsw 1  +C 2V  (3.2.25) VC 3B 2 = VC 3 A1 = −C I 2  3 C 3B1 3 out 2  (C2 + C3 )   T  +C2 (C3 (Vin + VC 2 B1 + VC 3B1) − I out sw )  2   Using MATLAB to compare these equations to the values derived from simulation, the voltages were accurate to within one-thousandth of a volt. Furthermore, this allows for the easy calculation of the capacitor voltage ripple. I T ∆VC1 = VC1B 2 − VC1B1 = out sw C1 (3.2.26) I T ∆VC 2 = VC 2 A2 − VC 2 A1 = out sw C2 (3.2.27) These voltage ripples are peak-to-peak values. However, the voltage ripple for the output stage capacitors is not so easily calculated, as their maximum voltage occurs during the charging phase just slightly after the beginning of the half-period. This is a value best calculated numerically, if it is needed. As a rule of thumb, however, it may be better to go with: 68 I T ∆VC 3 ≈ out sw 2C3 (3.2.28) This is an approximation of the peak-to-peak voltage ripple on the final module capacitors, but it may be more valuable for designing the converter from the standpoint of being easy to use. Nevertheless, equations representing the currents and voltages at work within the 6X converter, and extending to any NX converter, have been derived. However, that leaves the converter’s efficiency to be analyzed. The power losses within the NX dc-dc converter can be broken down into four parts: the capacitor charging loss, the conduction loss, the switching loss, and the gate drive loss. Since the gate drive loss is likely the smallest component, and is often powered separately from the converter’s main supply, it can be neglected for now. That still leaves the other three components. In regards to multilevel converter circuits, this subject was already covered in [6] and [15], but its work can easily be extended to the NX converter. When discussing the subject of losses, it is best to begin by discussing that which is most familiar first. The switching loss refers to the power losses incurred while the semiconductor devices go through their transition from on to off and vice versa. This happens because the switch, considered to be a MOSFET in this case, does not turn on or off instantaneously. In fact, as soon as the gate threshold voltage is reached, the MOSFET will begin to conduct current, but the drain-source voltage will not begin to drop until the gate-source capacitor is completely charged [16]. This sort of behavior can be somewhat difficult to model, and a less accurate but more easily utilized method to model it can be found in [17]. According to this paper, the commonly used formula for the estimation of a MOSFET’s switching losses is: 1 1 2 Psw = VDS I D (ton + toff ) f sw + COSS VDS f sw 2 2 69 (3.2.29) This equation assumes a linear transition of the drain-source voltage, VDS, and the drain current, ID, during the turn-on and turn-off times, ton and toff. The first term represents the energy from the changing voltage and current multiplied by the switching frequency, fsw, giving the power. The second term accounts for the energy stored in the MOSFET’s output capacitance, COSS. Multiplying this by the frequency again gives the power losses from this effect. (3.2.29) estimates the switching losses for one MOSFET, so the inputs to the equation have to be changed for each device with different switching conditions or component parameters and power lost in each switch should be added up. Another thing to note in (3.2.29) is the direct dependence on the switching frequency; as the switching frequency increases, so does the associated switching loss. This would put an upper bound on the switching frequency of the converter. However, the actual behavior of the voltages and the currents during the switching transition cannot really be expected to behave in this fashion. To that end, (3.2.29) can be used only as a very rough estimation. Luckily, switching loss is not the main source of loss in most switched capacitor converters. The conduction loss and capacitor charging loss provide most of the losses. Conduction loss, in the case of switched capacitor converters, refers to the losses suffered due to the load current passing through the resistance within the circuit. These currents, described in 3.1 near (3.1.1), are based off of the average currents that must be passing through the various components of the system during each switching state. The resistance comes primarily from the semiconductors’ on-resistances, but the capacitors’ ESR and the trace resistance also contribute somewhat to the losses. For any analysis of these losses, the trace resistance will be ignored, and the ESR is expected to be low, as each individual capacitor’s ESR 70 is small and multiple capacitors in parallel are used for each stage capacitance. However, thanks to the symmetry of the circuit, the conduction loss should be the same during both switching states. Using state I of the 6X as an example, the conduction loss could be expressed as: Pcond = (4 I out )2 ( RS1P + RS 2 N ) + (2 I out )2 ( RS1B + RC1A + RC1B + RS 2 A (3.2.30) + RC 2 A + RC 2 B + RS 3P + RS 3B ) + (1I out )2 ( RC 3 A + RC 3B ) This should give an estimation of the conduction loss in the 6X converter, and is easily extendable to any NX converter. As seen in (3.2.30), it is purely dependant on the output current and the parasitic resistance. Lowering the circuit’s resistances should minimize this type of loss. Finally, only the capacitor charging loss remains. The capacitor charging loss is likely the largest source of power loss in a converter of this type. Whenever two capacitors are placed in series and one discharges into the other, a certain amount of energy is lost. The amount of energy stored in a capacitor is given by: 1 Ecap = CV 2 2 (3.2.31) At first, the two capacitors are at completely different voltages, but they reach equilibrium by the end of the half period. However, an energy loss is associated with this process. For example, consider two capacitors, both with capacitance C, separated by a resistor. One capacitor has voltage V+∆V and the other only V. The total energy in this setup is initially: 1 1 Estart = C (V 2 + ∆V 2 + 2V ∆V ) + CV 2 2 2 (3.2.32) Assuming that the RC time constant is very small compared to the switching period, the total energy at the end becomes: 1 1 Eend = C (2V 2 + 2V ∆V + ∆V 2 ) 2 2 71 (3.2.33) Thus, the energy lost in the transfer is equal to the difference between the two energies. In this example, it becomes: 1 Estart − Eend = C ∆V 2 4 (3.2.34) An easy way of applying this to any other system, like the 6X, is just calculating the energy lost by looking at the starting and ending energy of a two capacitor system and subtracting. The voltages given in (3.2.20)-(3.2.25) are useful for accomplishing this task. The power loss from this can be derived simply by taking the energy lost and multiplying by the switching frequency. Overall, the use of the RMS currents to calculate the overall resistive losses might provide a more direct means of calculating the efficiency. The NX converter still has other possibilities, however. 3.3. NX Converter ZCS Possibilities The similarities of the NX converter and the MMCCC suggest the possibility of achieving zero-current-switching to decrease the switching losses. It was already mentioned that ZCS reduces switching losses virtually to zero by forcing the switch current to zero during both the turn-on and turn-off transitions. Using [12] and [13] as a starting point, [19] suggested using ZCS in the NX converter. However, several of the results found in [19] have flaws capable of quick correction using the results shown in section 2.4 for the ZCS MMCCC. It can be easily shown that the ZCS version of the NX converter breaks down into circuits identical to those found in the MMCCC. To begin, it should be noted that all circuits have some amount of parasitic inductance in them. Wherever current loops exist in an electric circuit, and all circuits must possess them, stray inductance can be found. This is simply a result of Maxwell’s laws, and, although the 72 geometry can be carefully designed so as to minimize the stray inductance, it can never be reduced to zero. With that in mind, this stray inductance can be utilized to form the resonant circuits needed to achieve ZCS. Consider the slightly modified circuit in Figure 50 below. Figure 50. The ZCS 6X Converter This circuit is the same as that shown in Figure 44 before, but now stray inductances are included in each module. Once again, the aim is to design and tune the circuit so that the resonant frequency of each series resonant loop equals the converter’s switching frequency. To analyze the behavior of the circuit, however, the circuit’s loop equations must be derived. To this end, it is important to examine the converter’s two switching states. Switching states I and II can be seen in Figure 51 and Figure 52 below. It is easy to see that each state can be broken down into three individual circuits that can be solved for and made to fit the circuit’s general behavior using the average currents and voltages. 73 Figure 51. ZCS 6X Converter Switching State I Figure 52. CS 6X Converter Switching State II The individual circuits end up being identical in behavior to those shown in Figure 36 and Figure 37 from earlier. Since states I and II are symmetric, the solutions for the inductor currents and capacitor voltages will be the same, but with a time shift equal to half of the switching period. Making note of that, consider the equations governing the behavior of switching state I. diL1B + vC1B dt (3.3.1) dvC1B dt (3.3.2) diL 2 A − vC1A + vC 2 A dt (3.3.3) Vin = L1B iL1B = C1B Vin = L2 A iL 2 A = −C1A dvC1A dv = C2 A C 2 A dt dt 74 (3.3.4) Vin = L3B iL3 = −C2 B diL3B − vC 2 B + vC 3B dt (3.3.5) dvC 2 B dv = C3B C 3B + Iout dt dt (3.3.6) dvC 3 A dt (3.3.7) I out = −C3 A The similarities between (3.3.1)-(3.3.7) and (2.4.4)-(2.4.13) are striking. Since the average voltage across each individual capacitor in each module is the same as in the MMCCC, the results will be the same. Once again, it is assumed that all capacitors equal C and the inductors follow the rule where L1A=L1B=L, L2A=L2B=2L and L3A=L3B=2xL. With this, the solutions for the inductor currents and capacitor voltages in state I are as follows. ω= 1 2π = LC Tsw (3.3.8) iL1B (t ) = π I out sin(ωt ) (3.3.9) vC1B (t ) = Vin − π I out cos(ωt ) Cω (3.3.10) vC1A (t ) = Vin + π I out cos(ωt ) Cω (3.3.11) vC 2 A (t ) = 2Vin − π I out cos(ωt ) Cω I πI vC 2 B (t ) = − out t + 3Vin + out 2C 4 I x + out 2 L 3π I out + C 4 (3.3.12) L ω cos( t) C x (3.3.13) ω L sin( t) C x I I 3π I out ω ω iL3B (t ) = out − out cos( t) + sin( t) 2 2 4 x x x 75 (3.3.14) vC 3 A (t ) = 4Vin + π Iout 2 I πI vC 3B (t ) = − out t + 4Vin + out 2C 4 I x − out 2 L Iout − t C C L 3π I out − C 4 (3.3.15) L ω cos( t) C x (3.3.16) ω L sin( t) C x x ≈ 0.79821687 (3.3.17) These solutions are just like the MMCCC, but the each switching state is symmetric. In other words, the solutions for state II are identical to (3.3.8)-(3.3.17) except that the module capacitor and inductor to which they apply are reversed and a time shift is used. As for using these equations, several points can quickly be made. For the first and second modules, the capacitor voltage ripple is seen to be: ∆vC1,2 = 2π I out Cω (3.3.18) This value is a peak-to-peak voltage, and it is easily shown to be the same as (3.2.26) or (3.2.27) for a fixed set of capacitance, output power, and switching frequency. Furthermore, the voltage behavior of the output stage capacitors is not different enough from the others to warrant the use of a totally different equation for calculating the voltage ripple. The third module’s capacitor’s voltage ripples can be safely estimated as: ∆vC 3 ≈ π Iout Cω (3.3.19) This takes into account the effect of the two capacitors also being in series, reducing the ripple in half. In order to prove that this concept is possible, a simulation was performed. Using Saber Simulator once again, the ZCS 6X converter was simulated using conditions that should already be familiar. The switching frequency was 100 kHz, the input voltage was 20 76 V, the load was 20 , and the stage capacitors were all 100 µF. The inductors for the first, second and third modules were 25.33 nH, 50.66 nH and 40.4377 nH respectively. Switch resistances and diode voltages were kept very low to estimate ideal conditions, and no dead-time was utilized. The results of the simulation are shown in the following figures. Figure 53. ZCS 6X Simulation Input and Output Voltages Figure 54. ZCS 6X Simulation Inductor Currents 77 Figure 55. ZCS 6X Simulation Capacitor Voltages In Figure 53 it can be seen that the output voltage is, on average, very close to eight times the input. Furthermore, the voltage ripple now behaves very much like a cosine function. The inductor currents depicted in Figure 54 illustrate the zero-current-switching that is achieved with this setup. Finally, Figure 55 shows the validity of (3.3.11), (3.3.12), (3.3.15) and (3.3.16), and it gives a graphic example of what behavior to expect. Although all of this seems to suggest that the ZCS NX converter is a viable option, there are drawbacks. During all the preceding analysis, the circuits were considered to be either composed entirely of resistances and capacitances or capacitances and inductances. In reality, all three coexist in the physical design, and the parasitic circuit elements are distributed throughout the layout. This means not only that the voltage and current behavior should deviate substantially from the previously predicted cases, but that tuning the circuit for ZCS would be extremely difficult. One area in particular to note on the NX converter topology is the node connecting the half-bridges’ center points to the capacitors, as seen in Figure 44. The current through this region is alternating current, so it can be referred to as the ac node. Currents from two loops that 78 were, up to now, considered separate flow through this area. If inductance is located here, it will not only have an effect on the resonant behavior of the circuit, but it will also couple the current loops together. This has the effect of making the circuit nearly impossible to tune without adjusting everything all at once, as well as making the solutions for the currents and voltages significantly more complicated. While it may very well be that this effect cannot be ignored in a real circuit, it is perhaps not terribly significant in the larger scheme of things. However, with the general idea of the stray inductance having an effect on the converter’s behavior, more sense can now be made about the experimental results. 3.4. 8X Converter Experimental Results In order to test the actual capabilities of the NX converter topology, an 8X 1-kW converter was designed, built and tested. The operating input voltage range was between 20 V and 40 V, and it was decided to use the hard-switching design rather than spend so much time trying to tune a circuit that could, in all likelihood, be incapable of being properly tuned. In order to make this circuit, the following choices were made. Since the input voltage can be as high as 40 V, MOSFETs capable of blocking at least one or two times that were needed. Using a multiplication factor of 1.5 for safety, half of the MOSFETs needed to be rated for 60 V and the other half for 120 V. This limited the selection available considerably, and the decision was made based on a combination of low on-resistance and low gate charge. The result was the choice to use the IPB017N06N3 G and the IPB036N12N3 G from Infineon Technologies, which were rated at 180 A and 60 V and 120 V respectively. As for the stage capacitors, choices were even more limited. Due to their small size, high-temperature capabilities, low ESR and ESL, multilayer ceramic capacitors (MLCC) were selected as the optimal choice. However, finding MLCCs with both voltage ratings and 79 capacitances that were high enough made the range of options smaller yet. In the end, two types of capacitors were purchased: the 100-V 15 µF C5750X7S2A156M, and the 250-V 2.2 µF C5750X7T2E225K. Both capacitors are manufactured by TDK Corporation. Initially, the C5750X7S2A156M was going to be used in the first module, but the C5750X7T2E225K ended up being used for the stage capacitors in all the modules; the C5750X7S2A156M was relegated to supporting the dc input voltage. Six capacitors were used in place of each stage capacitor illustrated in Figure 44 due to size constraints. Before continuing, however, an interesting feature of MLCCs should be mentioned. The materials used to produce MLCCs possess properties that cause the device’s capacitance to change with the dc voltage across it. While the ratings may state 2.2 µF as the capacitance, it is important to note that this is the case only at 0 V. As the voltage increases from zero, the capacitance of the device decreases rapidly. This makes the stage capacitances even smaller than their rated values, and makes ZCS applicable only at a single input voltage without specialized control. Since the stage capacitors experience one, two, three and four times the input voltage, the amount of the capacitance drop changes from stage to stage as well. This needs to be taken into account when performing simulations of any real circuits. The circuit was constructed on a four-layer printed circuit board (PCB) using the top layer for the power circuitry, the bottom for the control electronics, and the middle two layers as power and ground planes to the control layer. Due to the layout, wires were used when connecting the ac node together, so some inductance on the order of a few tens of nanohenries should be in that area. The circuit itself is depicted in Figure 56 below. Although not depicted, the circuit was controlled using a pulse width modulation (PWM) chip made by Texas Instruments called the UC3525. The external components connected to this chip determine the 80 switching frequency, duty cycle and dead-time produced by the IC, and the operating range of these variables is also determined by them. Figure 56. The 1-kW 8X prototype (For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this thesis) In Figure 56, the top layer of the 8X prototype can be seen. This is where all the components of the power circuit are located. The MOSFETs are the small black boxes with metal leads, and the four pairs of them on the board’s left-hand side are the circuit’s half-bridges. The white wires are what were previously called the AC nodes, and they are made of wires in order to allow the use of a current sensor. The light brown squares are the MLCCs; the right-hand sets of MLCCs comprise the stage capacitors. The dimensions of the board are about 3 inches wide by seven inches long, making the converter quite small in size. All of this had the effect of increasing the power density. With this new board at hand, tests were performed to both confirm its function as well as determining its efficiency. 81 Once debugging was concluded, consecutive tests were performed under conditions of increasing input voltage to confirm the circuit’s proper functioning and its durability. These tests were performed using a switching frequency of 164 kHz, a dead-time of 500 ns as determined by the PWM chip, and a load of 200 . The first results of this testing can be seen in Figure 57. Figure 57. 8X Prototype Converter Waveforms with Vin=10 V (Channel 1: Vgs, 20 V/div; Channel 2: Vds, 10 V/div; Channel 3: Vds, 10 V/div; Channel 4: IIs, 5 A/div) In Figure 57, channel 1 displays the gate-source control signal to half of the MOSFETs (Vgs), while channels 2 and 3 show two of the MOSFET’s switching voltages (Vds). Channel 4 shows a current waveform, IIs, which is the current passing through the so-called AC node between the half-bridges and the stage capacitors. As can be seen in the above figure, the overall behavior of the converter is, in reality, quite unlike the ideal cases. Instead, simulation has shown that a more complex situation exists where the effects of stray inductance cannot be ignored. Using 82 capacitances and resistances roughly equal to those within the circuit, along with estimations for the stray inductance, the following current can be simulated. Figure 58. Simulated 8X IIs Current Waveform To be more specific, the stage capacitors, C1, C2, C3, and C4, are 12.9 µF, 11.6 µF, 10.2 µF, and 8.8 µF, respectively; this assumes an input voltage of 20 V, but that will have little impact on the basic shape of IIs. Using the datasheets for estimations, the capacitors’ ESR was 730 µ , and the 60 V and 120 V MOSFET on-resistances were 1.7 m and 3.6 m , respectively. The stray inductance placed at the AC portion of the converter was 10 nH. In both Figure 57 and Figure 58, the effect of the dead-time can be seen each time the current is forced back to zero, breaking up the otherwise sinusoidal waveform. Despite these results, testing progressed successfully. At the same operating point as Figure 57, the 8X function was also confirmed, as can be seen in Figure 59 below. 83 Figure 59. 8X Prototype Converter Waveforms with Vin=10 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 25 V/div; Channel 3: Vin, 5 V/div; Channel 4: IIs, 5 A/div) As expected, the output voltage is very close to being eight times the input voltage. Continuing on, the input voltage was gradually increased in order to confirm the converter’s safe operation. In Figure 60, the input voltage was 18 V, while Figure 61 shows the same waveforms with the input increased to 20.4 V, finally entering the nominal operating range. No trouble was encountered, so the voltage was further increased. In Figure 62, the input was increased to 35 V, and the MOSFETs were more closely monitored to ensure that they were still functioning correctly. Continuing this trend, the voltage was raised to 37 V, as seen in Figure 63, where the DC supply’s limit was reached, limiting further voltage increases. Nevertheless, no problems were encountered, and the output voltage was measured to be 288 V with the input of 37 V, just slightly lower than the expected 296 V. This small voltage drop is representative of larger losses 84 experienced at this power level, but it is still good. Figure 60. 8X Prototype Converter Waveforms with Vin=18 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 50 V/div; Channel 3: Vin, 10 V/div; Channel 4: IIs, 10 A/div) Figure 61. 8X Prototype Converter Waveforms with Vin=20.4 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 50 V/div; Channel 3: Vin, 10 V/div; Channel 4: IIs, 10 A/div) 85 Figure 62. 8X Prototype Converter Waveforms with Vin=35 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 100 V/div; Channel 3: Vds, 25 V/div; Channel 4: IIs, 10 A/div) Figure 63. 8X Prototype Converter Waveforms with Vin=37 V (Channel 1: Vgs, 20 V/div; Channel 2: Vout, 100 V/div; Channel 3: Vds, 25 V/div; Channel 4: IIs, 10 A/div) 86 With the correct operation confirmed, the efficiency of the converter became a concern. In order to measure the efficiency, two tests were performed, each at a slightly different switching frequency. In this case, the input voltage was held constant at 20 V, while the load resistance was changed. The output power and efficiency were measured at each point using a power meter, and this test was performed at both 156 kHz and 165 kHz. These frequencies are close, but the operating frequency range of the converter was limited by the PWM chip’s passive components. The results of the test can be seen in Figure 64. Figure 64. Efficiency vs. Output Power for the 8X Prototype This curve matches the curve seen in many other similar converters using MOSFETs, starting out low, increasingly rapidly to a maximum, and finally slowly declining with increased power. As can be seen, the two frequencies tested were too close to clearly illustrate the effects of frequency on the efficiency, but it seems likely that the 165 kHz operation is marginally better. This would likely be the result of reduced voltage ripple on the capacitors, reducing the voltage differences between them and lowering the transient currents experienced in the circuit. All things considered, however, the efficiency is fairly high. Efficiencies greater than 98% were 87 measured at several points during testing, confirming the converter’s viability as a candidate for use in high-temperature environments requiring both high efficiency and small size. 4. Conclusions In the end, it is important to consider the system under consideration once again. When the power produced by a photovoltaic panel must be conditioned for use in other applications, such as the one described in Figure 1, several issues must be considered. First, the PV panel’s voltage must be boosted and inverted, but many options are available to accomplish this. The traditional boost and isolated full-bridge converters could boost and regulate the panel’s voltage for use by the inverter, but they cannot achieve a very high boost ratio and their magnetic components not only increase their size and weight but also limit their temperature operation range. In order to achieve the desired high voltage gain, efficiency and power density, switched capacitor topologies represent an attractive option. While not really capable of regulating the DC voltage fed to the inverter, the inverter itself could be adjusted via its modulation index to achieve the desired AC output. As was shown, the flying capacitor converter had a low number of active devices, but it was not modular, has efficiency that is inversely related to the boost factor, and has a relatively complicated control scheme. On the other hand, the MMCCC is modular, has a simple control scheme, and has an efficiency not affected by the boost factor, but it requires more devices for a given voltage gain and some of those switches need to sustain greater voltage stresses. In order to overcome these negative aspects, the NX dc-dc converter was developed. The NX converter has many of the advantages of its relatives, but only a few of the drawbacks. It is a modular design, making it easy to increase the boost factor. The control is also simple, just using a complementary 50% duty cycle to control all of the semiconductors 88 involved. While nearly half of the active devices do have to be able to sustain twice the input voltage, the overall switch count has been reduced to the same as the flying capacitor topology. Finally, the charge paths have a maximum of three active devices, limiting the resistive losses and eliminating any dependency of the efficiency on the boost factor like the flying capacitor converter. These features combine to make the NX converter well-suited for use as a microconverter in PV systems. 89 BIBLIOGRAPHY 90 BIBLIOGRAPHY [1] Mitsubishi Electric, “MJT Series,” L-175-1-B8778-C datasheet, Jul. 2011 [Revised Mar. 2012]. [2] N. Mohan, T. M. Undeland, and W. P. Robbins, Power Electronics: Converters, Applications, and Design. New York, NY: John Wiley & Sons, Inc., 1995, pp. 172-178. [3] R. W. Erickson and D. Maksimovic, Fundamentals of Power Electronics, 2nd ed. New York, NY: Springer Science + Business Media, LLC, 2001, pp. 149-167. [4] F. Z. Peng, F. Zhang and Z. Qian, “A novel compact dc-dc converter for 42 V systems,” in IEEE 34th Annual Power Electronics Specialist Conference, 2003, pp. 33-38. [5] F. Zhang, F. Z. Peng and Z. Qian, “Study of the multilevel converters in dc-dc applications,” in IEEE 35th Annual Power Electronics Specialist Conference, 2004, pp. 17021706. [6] Z. Pan, F. Zhang and F. Z. Peng, “Power losses and efficiency analysis of multilevel dc-dc converters,” in Twentieth Annual IEEE Applied Power Electronics Conference and Exposition, 2005, pp. 1393-1398. [7] W. Qian, F. Z. Peng, M. Shen and L. M. Tolbert, “3X dc-dc multiplier/divider for HEV systems,” in Twenty-Fourth Annual IEEE Applied Power Electronics Conference and Exposition, 2009, pp. 1109-1114. [8] W. Qian, H. Cha, F. Z. Peng and L. M. Tolbert, “55-kW variable 3X dc-dc converter for plug-in hybrid electric vehicles,” IEEE Transactions on Power Electronics, vol. 27, no. 4, Apr., pp. 1668-1678, 2012. [9] F. H. Khan and L. M. Tolbert, “A multilevel modular capacitor clamped dc-dc converter,” in Conference Record of the 2006 IEEE Industry Applications Conference, 2006, pp. 966-973. [10] F. H. Khan and L. M. Tolbert, “5 kW multilevel dc-dc converter for hybrid electric and fuel cell automotive applications,” in Conference Record of the 2007 IEEE Industry Applications Conference, 2007, pp. 628-635. [11] F. H. Khan and L. M. Tolbert, “A multilevel modular capacitor clamped dc-dc converter,” IEEE Transactions on Industry Applications, vol. 43, no. 6, Nov.-Dec., pp. 1628-1638, 2007. [12] D. Cao and F. Z. Peng, “Zero-current-switching multilevel modular switched-capacitor dcdc converter,” in IEEE Energy Conversion Congress and Exposition, 2009, pp. 3516-3522. 91 [13] D. Cao and F. Z. Peng, “A family of zero current switching switched-capacitor dc-dc converters,” in 2010 Twenty-Fifth Annual IEEE Applied Power Electronics Conference and Exposition, 2010, pp. 1365-1373. [14] D. Cao and F. Z. Peng, “Multiphase multilevel modular dc-dc converter for high current high gain TEG application,” in 2010 IEEE Energy Conversion Congress and Exposition, 2010, pp. 4230-4237. [15] F. Z. Peng, “A generalized multilevel inverter topology with self voltage balancing,” IEEE Transactions on Industry Applications, vol. 37, no. 2, Mar.-Apr., pp. 611-618, 2001. [16] International Rectifier, Appl. Note AN-1084, pp. 1-13. [17] Y. Xiong, S. Sun, H. Jia, P. Shea and Z. J. Shen, “New physical insights on power MOSFET switching losses,” IEEE Transactions on Power Electronics, vol. 24, no. 2, Feb., pp. 525-531, 2009. [18] W. Qian, D. Cao, J. Cintron-Rivera, M. Gebben, D. Wey and F. Z. Peng, “A switchedcapacitor dc-dc converter with high voltage gain and reduced component rating and count,” IEEE Transactions on Industry Applications, vol. 48, no. 4, Jul.-Aug., pp. 1397-1406, 2012. [19] M. Gebben, J. G. Cintron-Rivera, W. Qian, D. Cao, X. Pei and F. Z. Peng, “A zero-currentswitching multilevel switched capacitor dc-dc converter,” in 2011 IEEE Energy Conversion Congress and Exposition, 2011, pp. 1291-1295. 92