Next Article in Journal
Spacecraft Medium Voltage Direct-Current (MVDC) Power and Propulsion System
Previous Article in Journal
JEEMRC: Joint Event Detection and Extraction via an End-to-End Machine Reading Comprehension Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Parasitic-Based Model for Characterizing False Turn-On and Switching-Based Voltage Oscillation in Hybrid T-Type Converter

1
Centre for Industrial Electronics (CIE), Institute of Mechanical and Electrical Engineering, University of Southern Denmark, 6400 Sonderborg, Denmark
2
Danfoss Drives A/S, 6300 Grasten, Denmark
3
Semikron Danfoss, Husumer Straße 251, 24941 Flensburg, Germany
4
FuE-Zentrum FH Kiel GmbH, Schwentinestr. 24, 24149 Kiel, Germany
*
Author to whom correspondence should be addressed.
Electronics 2024, 13(10), 1808; https://doi.org/10.3390/electronics13101808
Submission received: 21 March 2024 / Revised: 25 April 2024 / Accepted: 30 April 2024 / Published: 7 May 2024
(This article belongs to the Topic Power Electronics Converters)

Abstract

:
High frequency and high voltage switching converters utilizing wide bandgap semiconductors are gaining popularity thanks to their compactness and improved efficiency. However, the faster switching requirements gives rise to new challenges. A key issue is the increased oscillation of the drain–source voltage caused by the switching action of the complementary switch in the same phase or change of state of the other phase switches. The voltage stress caused by these oscillations can damage the switch. Furthermore, the high dv/dt during turning-on of one switch might result in false turn-on of the complementary switch due to the miller effect. In this paper, these issues are investigated in a T-type converter through analytical and experimental analysis. Based on the proposed analytical approach, simple and cost-wise solutions utilizing an optimum design of gate driver circuits and circuit layout modifications can be developed to cope with the aforementioned issues. A comprehensive analytical model of the converter with consideration of parasitic capacitances and inductances is developed. By performing sensitivity analysis on the model, the effect of the parasitic parameters on the drain–source voltage oscillation and gate–source voltage amplitude in case of false turn-on is studied. The validity of the model is then assessed through numerical simulations and experimental results.

1. Introduction

Silicon carbide (SiC) semiconductors favor faster switching, lower switching loss and lower turn-on voltage compared with the traditional silicon technology [1,2,3]. The fast-switching capability of the SiC switches provide the opportunity for developing power electronic converters with high switching frequency and high voltage ratings. Through boosting the switching frequency, the passive components of the converters can be shrunk, hence offering compact and high power density converter design [4,5]. Thanks to the high switching capability of SiC, they can be used in resonant converters, which are widely employed in various applications like electric vehicle battery chargers [6], drive systems [7], and V2X power converters [8]. Furthermore, SiC semiconductors are able to maintain higher junction temperature, which makes them suitable candidates for the compact power electronic converters in integrated motor drive applications [9]. Electro-thermal modeling of SiC-based converters for different power range has been studied in the literature [10,11,12] in order to improve the thermal performance of such converters.
Despite the aforementioned superior characteristics, the application of SiC semiconductors in high voltage/high frequency converters creates new issues [13]. On the one hand, the voltage overshoot across the switch’s drain–source terminals is higher due to the ultrafast switching (high dV/dt) [14]. The voltage overshoot stresses the switch and can damage the switch if it exceeds the switch voltage rating. On the other hand, unintentional voltage oscillations across the gate–source of a switch during turning-on of another switch might give rise to false turn-on [15,16]. SiC switches are more prone to such false turn-on due to the lower turn-on threshold voltage [17]. False turn-on contributes to cross conduction across two switches in a leg, which increases the losses and might even result in switch failure due to overcurrent [18]. Both drain–source voltage overshoot and false turn-on issues are dependent on the parasitic capacitances and inductances in the circuit layout.
In addition to the abovementioned issues, another challenge is associated with three-phase compact converters which are used in high-power, high-frequency application [19]. In this case, a change in the switching state of switches in one phase gives rise to drain–source voltage oscillations in the switches of the other phases [20,21]. This may have a negative impact on the switch voltage blocking and load current quality. This issue elevates in the case of three-phase PMs, which increase the power density through compact design [22]. In such a compact design, the mutual inductance/capacitance between each two phases of the three-phase converter are intensified, hence elevating the cross-phase voltage oscillations [23]. These oscillations can be mitigated using an active/passive snubber circuit [24,25].
The former issue, false turn-on, can be mitigated using an active Miller clamp technique in the gate driver [26,27]. However, this solution requires a gate driver design specific to the switch characteristics, which increases the cost [28]. Additionally, the reliability of this approach may be compromised depending on the circuit layout [29,30]. An extra isolated voltage source can be connected to the gate–source capacitor during the switch turn-off to maintain the gate–source voltage at less than the turn-on threshold [31]. Despite the improved reliability, the mentioned protection circuit is not cost effective [31]. The drain–source voltage overshoot can be limited within the safe range using a snubber circuit [32,33,34]. However, the conventional snubber configurations cannot be used in high-frequency applications because of the limitation in power loss and switching speed [35,36,37]. In practice, the issues associated with the voltage overshoot may be resolved slightly in the converter design stage by employing suitable SiC switches as well as a low parasitic power module (PM) and a PCB layout design [38,39].
Prior to introducing costly solutions such as snubbers, modified gate driver designs, and special circuit layouts, it is important to clarify the necessity of such solutions. A compact electrothermal model has been presented in [40] to characterize the SiC switch, where the whole circuit parasitic effects was not fully considered. Experimental characterization of the SiC switch and its body diode is also presented in [12] using a double pulse test setup, as well as a temperature-dependent physical thermal model in [41]. Although, the mentioned approaches are usable to precisely switch characteristic detection, proposing a general circuit model to determine the impact of the non-optimal layout parameters on the voltage stress and oscillation across the pre-characterized SiC switch is of interest. This paper proposes an analytical study of the mentioned issues to help quantify the drain–source voltage overshoot and gate–source voltage oscillations. The analytical results provide a baseline for assessment of the safe converter operation despite the presence of the drain–source voltage overshoot and the possibility of false turn-on. Based on the proposed analytical approach, simple and cost-wise solutions utilizing an optimum design of gate driver circuits and circuit layout modifications can be developed. This way, the cost of the converter can be reduced without compromising the performance.
The specific contributions of the paper are as follows.
  • A general switching model of three-phase hybrid T-Type converter, consisting of SiC MOSFETs and Si IGBTs, is developed, which has not been studied before. This type of front-end converter is chosen due to the comparison results yielded from [42]. All possible effective parasitic components, which are modeled as stray inductances and switch intrinsic capacitors, are considered in the model. This makes a comprehensive model to be applied to other power electronic converter types.
  • Unlike the conventional analytical model of the converter, sensitivity analysis is carried out here to investigate the impact of parasitic capacitances and inductances on the damping and natural oscillation frequency of the switch drain–source and gate–source voltages, simultaneously. This way, the most critical parasitic parameters affecting voltage overshoot and false turn-on are determined using the same analytical model.
  • Regarding the proposed built compact converter, a new issue has been raised which is affecting the voltage oscillation characteristics across the switch. Accordingly, the effect of the switching in one phase on the drain–source voltage of the switches in the other phases is investigated in detail by simulating the equivalent circuit model, which was not completely focused in the literature.
  • Simulation and experimental results have been carried out to verify the analytical model.
The rest of the paper is organized as follows. The possibility of false turn-on and drain–source voltage oscillation in a T-type converter with consideration of different modulation scenarios is investigated in Section 2. In Section 3, a detailed analytical model of one of the phases of the T-type converter is derived. In Section 4, the analytical model is analyzed to identify the sensitivity of damping and natural oscillation frequency on the parasitic parameters. In Section 5, the single-phase model is extended to three phases to analyze the impact of switching in one phase on the performance of the other phases. Experimental results based on a hybrid T-type converter are presented in Section 6 to validate the analytical and numerical results. Section 7 concludes the paper.

2. Topology-Based Miller Plateau Analysis

To illustrate the mechanism of false turn-on, an active T-type rectifier is considered in this study. The reason for choosing this converter topology is its acceptability in the industry as it has the best compromise between high power density and efficiency, as well as the possibility of reaching the pure sinusoidal supply current with a smaller input filter [36]. The topology of a single-phase active T-Type converter is shown in Figure 1. Switches Q1 and Q2 should maintain a high breakdown voltage that is equal to the DC link voltage (2VDC) while the breakdown voltage cuts in half (VDC) for the zero-level switches of Q3 and Q4. Therefore, SiC MOSFETs are used for Q1 and Q2 switches, while Q3 and Q4 are Si IGBT.
The zero-voltage level is provided using two back-to-back IGBTs with the same source and different command gate signals. The T-type converter is generally composed of two half-bridges (see Figure 1). The upper side half-bridge loop (red color) includes complimentary switches Q1 and Q3 while the lower side half-bridge (blue color) includes Q2 and Q4 as the complimentary set of switches. The modulation algorithm depicted in Figure 2 shows that in the positive half cycle of the reference waveform (voltage or current depending on the application), the lower side switch of the leg (Q2) is always off. Accordingly, during the switching of the upper half-bridge at the positive cycle, switch Q4 is on. Thus, the current can be conducted in both directions through Q4. The same situation is considered for the lower half-bridge. According to the control algorithm, switch Q1 is off during the whole negative half cycle of the reference waveform. Hence, the Q3 is always on at this period and can conduct the bi-directional current through itself. Also, the dead time is defined at switching state changes.
As mentioned, to protect both half-bridges from shoot-through issues, the dead time is necessary to be considered for switching transient between Q1 and Q3, as well as switching states for Q2 and Q4. Nevertheless, each switch suffers from false turn-on occurrence after turning-on of the complimentary switch, which also causes unwanted shoot-through. In this paper, the Miller effect is investigated in three cases presented below:
(1)
For the upper side half-bridge, where turning-on of SiC MOSFET (Q1) affects false turn-on of Si IGBT (Q3);
(2)
For the lower side half-bridge, where turning-on of Si IGBT (Q4) affects false turn-on of SiC MOSFET (Q2);
(3)
For leg switches, where turning-on of SiC MOSFET (Q1) affects false turn-on of SiC MOSFET (Q2).
Accordingly, for cases 1 and 2, the gate–source voltage oscillations of switches Q3 and Q2 are investigated during the transient turning-on of switches Q1 and Q4, respectively. In addition, for case 3, the gate–source voltage oscillations of switch Q2 should be monitored during the transient turning-on of switch Q1. During the transient, in all cases, the drain voltage variation of the target switch equals 0.5 Vdc. This is found by referring to the switching modulation of the T-type converter. Although the electric model and circuit equations of case 3 are the same as the others, this case is the most extreme one. The reason is that case 3 has fewer switches in the loop, which results in less loop resistance and a lower damping ratio. This means the oscillation amplitude and the possibility of false turning-on magnifies in case 3. In addition, the turn-on threshold voltage of SiC MOSFET is lower than IGBT, which leads case 3 to be more susceptible to a false turn-on. Therefore, as the worst one, case 3 was chosen to be investigated in this paper.
Switch Q2 is supposed to be off in theory in the transient time of Q1 turning on. However, the following analysis shows that the induced voltage across the gate–source capacitor of Switch Q2 may even exceed the turn-on threshold voltage, which results in cross conduction. Repeating the problem in thousands of cycles will damage and even burn the switches. For demonstrative purposes, the parasitic capacitances of switches Q2, Q3, and Q4 are illustrated in Figure 3. Just before a false turn-on occurs, Q2 is in its off state while Vdrv,2 is considered to be grounded. When Q1 turns on, the drain terminal of Q2 faces a high dv/dt, while the source of Q2 is directly connected to the ground through a leakage inductance. High dv/dt at drain terminal of Q2 leads to a current passing through gate–drain capacitance (Cgd2). The reason is that before Q1 turns on, Q3 and Q4 conducted the current. Accordingly, the drain voltage was Vdc when Q3 and Q4 were conducting. After turning on Q1, the drain voltage jumps to 2Vdc, which results in current flow through Q2 gate resistance, shown as RG2 in Figure 3. Exceeding gate resistance voltage from the turn-on threshold voltage of Q2, this switch is pushed to cross conducting, mistakenly causing a significant switching loss. More detailed analysis of false turn-on in T-type converters is still challenging in order to improve the reliable performance of the next generation ultra-high frequency converters which will use WBG switches.

3. Developed Model-Based Study

This paper aims to reveal the effective switch’s parameters or PCB layout leakages on false turn-on and voltage oscillation characteristics across the drain–source of the switch. Here, this is carried out by assessing the dependency of the natural frequency, ωn, and damping ratio, ζ, on the switch’s intrinsic parameters (Cds, Cgs, Cgd) and circuit leakage inductances. The simple and developed models considering all parasitics are analyzed mathematically in the following.

3.1. Simple Circuit Model

Based on the abovementioned, an abrupt turn-on of Q1 results in high positive voltage derivative at the drain terminal of Q2. Q1 turn-on time depends on the turn-on gate resistance the switch’s intrinsic capacitances. In addition, the gate–source voltage of Q2 during turning-on of Q1 is affected by the gate resistance and Cgd dv/dt, which is considered as induced current. The well-known first-order circuit at this transient time (Q1 turn-on) is illustrated in Figure 4, to study parasitic negative effects on induced gate–source voltage amplitude. The transfer function of the mentioned model using the Laplace form is as follows:
V g s 2 s V d s 2 s = s R G 2 C g d 2 1 + s R G 2 C g d 2 + C g s 2
Vds2(s) and Vgs2(s) are the drain–source voltage and gate–source voltage of switch Q2, respectively. Variable s in Equation (1) can be replaced by to achieve the frequency response function as:
V g s 2 j ω V d s 2 j ω = j ω R G 2 C g d 2 1 + j ω R G 2 C g d 2 + C g s 2 .
In order to decrease false turn-on possibility, the absolute gain value of Equation (2) needs to be as low as possible. This guarantees Vgs2 to be less than the turn-on threshold voltage of Q2. Referring to Equation (2), it is obvious that false turn-on possibility may decrease by lower gate–drain capacitor, Cgd2 while it increases by low gate–source capacitor (Cgs2) values. The experimental and analytical results in next sections will verify the findings here.
It should be mentioned that the model illustrated in Figure 4 is not complete due to lack of drain–source capacitor, Cds2, and also the circuit leakage inductances. The latter has a considerable effect on Vds oscillation in some circuit topologies [43]. The following developed model is investigated in the circuit using SiC switches to comprehensively evaluate which condition is more prone to the occurrence of the false turn-on and voltage oscillation.

3.2. Higher Order False Turn-On Model Considering Parasitic Parameters

The Cgd dV/dt circuit shown in Figure 5 is used to analytically and experimentally investigation of SiC MOSFET’s false turn-on in one leg. In theory and simulation, the ON and OFF states of Q1 are modeled by a pulsating voltage source series to ON resistance whose value can be extracted from the switch datasheet document, in practice. The value of R1 is adjusted during the analysis such that the in-mind dv/dt is applied to the drain terminal of Q2. The gate terminal of the Q2 is connected to the ground through a gate resistor, RG2. The induced gate–source voltage (Vgs2) of Q2 is studied by changing dv/dt at the drain terminal of Q2. Using this technique, the maximum eligible dv/dt across the switch can be obtained in which the switch is still free of cross conduction.
The circuit shown in Figure 5a is the complete model of the structure illustrated in Figure 3, which includes all the parasitic capacitors of the switch as a result of using DOL technology for the die switch planting process, as well as the parasitic inductance because of the single layer conducting track. Figure 5b illustrates a rearrangement of the elements in Figure 5a in which Q2 is off. Now, the step responses of both Vgs2 and Vds2 in terms of the input step voltage are derived. By applying KCL to nodes, the following equations can be extracted:
V p V D 2 s L t o t a l + R t o t a l = s V d s 2 C d s 2 + s V g d 2 C g d 2 ,
s V g d 2 C g d 2 = V g s 2 s C g s 2 + V G 2 s L G 2 + R G 2 ,
V S 2 s L l + R l = s V d s 2 C d s 2 + s V g s 2 C g s 2 ,
V G 2 V G S , t r a n . S L G 2 = V G S , t r a n . R G 2 ,
L t o t a l = L l + + L m i d + L + + L ,
R t o t a l = R l + + R m i d + R + + R ,
where L l ± , L m i d , and L ± are the leakage inductances of the upper and lower side of the leg, leakage inductance at the middle of the leg between two switches and stray inductance of the conductor from DC link to the leg, respectively. The same index is also used for the parasitic resistances. The whole system parameters which are modeled in the simulations and used in analysis are depicted in Table 1.
The values of parasitic capacitances are extracted from the datasheet of switches that are used in the experimental setup. Furthermore, the data regarding the leakage inductances is obtained through simulations of the entire PM using Ansys Q3D, employing the same size and material parameters, as depicted in Figure 6. This PM contains die switches for both T-Type converter and inverter.
Because of using Kelvin connection at source point of each die switch, the value of Ll− is considered near zero in analysis. Based on Equations (3)–(6) and by considering Vp as the input step function, the step response of Vds2 and Vgs2 is calculated. Then, the results are plotted as shown in Figure 7a,b. It is noteworthy that the initial value of the step voltage is 400 V as IGBTs Q3 and Q4 conduct. In addition, the final value of Q2 drain voltage (VD2) is 800 V instantaneously after Q1 turns on. Shown in Figure 7a, Vds oscillates about 175 ns and then it reaches 800 V in steady state. Also, Vgs in Figure 7b jumps to approximately 2.5 V at transient time of false turn-on. Once after reaching the peak amplitude, Vgs falls below the threshold voltage for turn-on (vth = 2.5 V). The results of Figure 7 are in a good agreement with the experimental results obtained in Section 6. The proposed model is also discussed with more details in the following section.

4. Model Discussion

4.1. Sensitivity Analysis of ωn and ζ

The general transfer function of Vds, based on the proposed model and its corresponding Equations (3)–(8), can be obtained as follows
V ds s = s + z 1 s + z 2 s 2 + 2 α s + β s s + p 1 s + p 2 s 2 + 2 ζ ω n s + ω n 2
where z1, z2 and p1, p2 are the zeros and poles, respectively, and α and β are two constants in the transfer function. Also, in Equation (9), ωn and ζ are the natural frequency and damping ratio, respectively, which determine the frequency of oscillations as well as the convergence speed of the system response. These two parameters are affected by the SiC MOSFET’s parasitic components. To study the sensitivity of the natural frequency and damping ratio on the parasitic inductances and capacitances of the device, variation of each parasitic element is examined on Equation (9).
Figure 8 depicts the impact of the switch’s parasitic capacitances on the damping ratio. Due to the sensitivity analysis, more focus has been carried out on the curve slope rather than its amplitude. Parameters of switch Q2 have been considered in this study. In Figure 8, the slopes for Cds2 and Cgd2 are negative up to certain values and will be increasingly positive beyond them. This means that increasing Cgd2 and Cds2 first increases and then decreases the Vds2 oscillations. However, the slope for Cgs2 is relatively constant, which shows that Cgs2 has no considerable impact on damping ratio. To ensure the same curve trend for Cgs2 due to the possibility of bifurcation, more values above the nominal Cgs2 (1.3 nF) are considered in the plot. Nevertheless, the same result has been concluded, which shows negligible impact of Cgs2 on damping ratio.
On the other hand, referring to Figure 9, Cds2 and Cgd2 have a sensible impact on the natural frequency. Among them, Cgd2 is more effective. The negative slope in these two curves shows that the oscillation frequency of Vds2 across Q2 will be reduced by increasing Cds2 and Cgd2. However, Cgs2 has a little impact on the oscillation’s frequency.
The result in Figure 10 clarifies these findings. In this figure, Vds2 is plotted for Cgd2 = 2 nF which is higher than the one in nominal Cgd2 (Figure 7a), as mentioned in Table 1. Comparing Figure 10 to Figure 7a, it can be seen that the settling time is less in Figure 10. This means bigger Cgd2 leads to higher damping ratio with less oscillatory frequency. In addition, the peak–peak voltage oscillation (Vp-p) decreases to 250 V in Figure 10, while the value is 400 V in Figure 7a. Since the overshoot amplitude has an inverse relation with the damping ratio, this finding in Figure 10 is in good agreement with the findings extracted from Figure 8 and Figure 9, which show that increasing Cgd2 will result in an increment of the damping ratio and a decrement of the natural frequency.
The dependencies of the natural frequency and the damping ratio of Vds on the parasitic inductors are presented in Figure 11 and Figure 12, respectively. Figure 11 shows a decrease in the damping ratio when the major loop inductance, Ltotal, is increased. On the other hand, the increment of the source inductance, Ll, is followed by a damping ratio increase. Based on this finding, source inductance may have a positive effect against false turn-on of switch in some cases. In return, the effect of gate inductance, LG, on the damping ratio of Vds can be neglected. Regarding Figure 12, it is obvious that the natural frequency of Vds decreases by increasing Ltotal and Ll. On the other hand, LG has no significant effect on the natural frequency. Comparison of Figure 8 and Figure 9 with Figure 11 and Figure 12 demonstrates both natural frequency and damping ratio are more influenced by the parasitic capacitances, as the slope variations are bigger in Figure 8 and Figure 9.

4.2. Dependency of False Turn-On on Parasitics

In this part, the dependency of false turn-on due to the induced Vgs on the parasitic capacitances is analyzed. Both simulation and theoretical results are illustrated in Figure 13 in the form of the gate–source peak amplitude. As seen, there exists an appropriate agreement among them. In addition, Figure 13 verifies that during false turn-on, Cgd has a prominent impact on the induced gate–source voltage. In addition, during false turn-on, a roughly piecewise linear relationship exists between Cgd and the gate–source voltage. This observation agrees with the studies previously presented in Section 3. It should be mentioned that 10 pF is the nominal value of Cgd. Therefore, in this analysis, the nominal value is considered by extrapolation which is impractical for SiC. This analysis tries to provide the insight of how substantial the effect of Cgd could be even with a decreased value. Further analysis of Figure 13 determines the increased value of Cgs is followed by a decrement of the magnitude of the induced gate–source voltage during false turn-on. This observation again is in appropriate compliance with the studies previously presented in Section 3.
Vgs waveform is analytically-drawn for the increased Cgs to 13 nF as shown in Figure 14. Comparing with Figure 7b, it is found that the peak value of Vgs is roughly 0.5 V, which is only 20% of peak voltage at nominal Cgs (1.3 nF). This finding can also be verified by referring to Figure 13c.
In a similar way, Figure 15 demonstrates the impact of Cds2 on Vgs for the increased Cds2 to 5 nF compared to nominal Cds2, which is mentioned in Table 1. Figure 15 shows that increasing Cds2 will result in the decrement of the induced gate–source voltage. This could not manifest based on the response function of the simple model presented in Equation (2). The effect of Cds for false turn-on is not as prominent as Cgs, but it is still required to be considered. It is considerable that increasing capacitance generally results in an increment of switching losses during the switches turn-off/on. Therefore, in high dv/dt circuits, the values of Cgs and Cds are required to be optimized to minimizing the total switching losses. Accordingly, their beneficial aspects during false turn-on are not completely offset by their parasitic effects during intended turn-off and turn-on. Experimental studies are presented in the next section to validate the reported analytical results.

5. Model Extension to Three-Phase T-Type Converter

The single-phase converter had been investigated in previous sections. The same challenges are assumable for each phase of three-phases system which is shown in Figure 16. In addition, in inverter mode, the voltage of each phase (VphX = VphxVM) will be affected by the switching of two other phases. This issue is reflected in form of voltage oscillation across VphX. At each time, the low-frequency reference voltage (Vref) applying to one phase (leg 1, 2, or 3) is in opposite polarity to the others depending on the phase angle. Table 2 shows the sign of reference voltages of the three phases over the positive half cycle of phase 1. The same situation will occur for the negative half cycle.
For 0° to 60° degrees of Vref in phase 1 (Vref,1), the effect of switching in legs 2 and 3 on Vph1 is investigated as a scenario. For simplification, ON and OFF state of the switch can be modeled by Ron and charged capacitor of Cds, respectively. Accordingly, the circuit schematics before and after changes in the switching state of both legs 2 and 3 are shown in Figure 17. Ron, Cds, Ron’, and Cds’ illustrate MOSFET and IGBT ON and OFF state, respectively. Regarding Table 2, SiC MOSFET Q21 turns off and Si IGBT Q23 turns on for leg 2 only. In similar way, SiC MOSFET Q32 turns off while Si IGBT Q34 turns on in leg 3 only.
Like the circuit analysis in Section 3, Figure 5, a simplified circuit schematic by equating the switching state changes to pulse voltage across Vph2,3 and midpoint (M) is shown in Figure 18 for θ1 = 0°. Here, the effect of switching state changes of leg 1 on Vph1 is not considered. The resulted voltage oscillation in this situation is obvious due to the leakage inductance of the route. Table 3 shows the amount of the parasitic components in Figure 18 and switches’ parameters including Cds and ON-resistance, which are employed for model simulation. It is noteworthy that the duty cycles of VP,ph2 and VP,ph3 depend on the exact phase degree of the related reference voltages, Vref,2 and Vref,3, which are 120° different in phase. As an example, for θ1 = 0°, 30° and 60°, the corresponding duty cycles of VP,ph2 and VP,ph3 are 0.86, 0.5, 0 and 0.86, 1, 0.86, respectively.
Figure 19 depicts the voltage oscillation of Vph1 with respect to switching state changes for two other phases (ph2 and ph3) when θ1 = 0°, 30°, and 60°. Based on the superposition principle, Vph1 contains both oscillations of each Vph2 and Vph3 which occur in different moments depending on θ1. For better demonstration, the amount of θ1 is kept constant for 6 high frequency cycles (6Ts). This is acceptable due to the slow changes of θ1 comparing the switching speed.
For each amount of θ1, the equivalent pulse voltage of phases 2 and 3 is illustrated, as well as the oscillation across phase 1 voltage (Vph1) resulting by switching state changes in two other phases. The total results are also shown in black color which is the integration of two distinct oscillation. There is an appropriate agreement with the superposition principle which can be applied for such simplified linear circuit. It is obvious that by any change in the switching state of phases 2 and 3, Vph1 oscillates around VDC. Without using a suitable filter at the output, this oscillation may be transferred to the load in stand-alone inverter mode, which is undesirable. Thus, this matter is taken into consideration in filter design for three-phase applications.

6. Experimental Evaluation

6.1. Double Pulse Setup Test

To authenticate the analytical and simulation-based investigations, the circuit depicted in Figure 5 underwent experimental scrutiny. This examination employed a SiC MOSFET for the legs and back-to-back Si IGBTs to attain a zero-voltage level, as illustrated in Figure 20. The three-phase T-Type converter utilized in this configuration served as the rectifier section of the entire Front-End converter, renowned for its high power density.
For characterization, the Tektronix AFG 31000 double pulse generator (Tektronix company, Portland, OR, USA) was employed to generate two pulses with adjustable duty cycles based on the switch current rate. This facilitated the characterization of both the gate driving circuit and the under-test switch. The current behavior of inductor L during the double pulse test (DPT) is depicted in Figure 21, where the inductance current exhibited a ramp-like increase upon switch activation and decreased slightly during the off time due to the voltage drop across the freewheeling diode. Subsequently, detailed investigation of the obtained Vgs in DPT was conducted to verify the theoretical findings.
Figure 22 considers all potential scenarios for false turn-on in the experimental testing circuits, with theoretical analysis and simulations focusing on case 3, regarded as the most critical scenario, wherein the false turn-on of switch Q2 is investigated during the activation of switch Q1.

6.2. False Turn-On Evaluation

To address all three potential scenarios outlined previously, identical circuits were subjected to experimental testing. Figure 23 presents the experimental outcomes of the test system, featuring VDC = 400 V and DPT inductance L = 220 uH. This figure depicts the gate–source voltage of the two switches with the highest likelihood of activation, alongside the gate–source and drain–source voltages of the switch undergoing the double pulse application. The results encompass all three previously introduced cases in Section 2. Consequently, switches Q1 and Q2 were observed in case 1, where switch Q3 underwent DPT testing. Additionally, switches Q1 and Q4 were monitored in case 2, where the double pulse was applied to switch Q2. Similarly, switches Q1 and Q3 were scrutinized in case 3, involving the application of the double pulse to the gate–source of switch Q1.
Across all three cases within the experimental setup (Figure 23a–c) and in case 3 of the analysis (Figure 7b), the peak value of the gate–source voltage was approximately 2.5 V. To facilitate better comparison, subsequent analysis exclusively concentrated on the experimental result of case 3.
Figure 24 provides a magnified perspective of the Vgs in Figure 23c, highlighting two turning-on transients (number 1: first turn-on; number 2: second turn-on) for clarity. Comparing the analytical/simulation model with experimental results reveals reasonable agreement. The peak value of the gate–source voltage in both experimental and analytical/simulation results is approximately 2.4 V. Notably, in the analytical model, the gate–source voltage reaches its first minimum value at 100 ns, whereas experimentally, this occurs slightly later, suggesting a slower experimental result attributed to larger circuit parasitics due to the new layout design. Nonetheless, the trends and shapes of the experimental waveforms reasonably align with the analytical/simulation results. It is worth noting that false turn-on during the falling edge of the applied gate–source pulses is less significant due to the deadtime between switching state changes, making monitoring of the rising edge of the pulses (indicated by dash loops) essential. In practical settings, induced gate–source voltage amplitude and voltage oscillation across switches are halved compared to DPT tests, as no inductance is placed in parallel to the switch, thus eliminating current conduction through the corresponding freewheeling diode during turning-off of the complementary switch. Consequently, the sensed step voltage across the switch in case 3 in real operation was VDC (400 V) rather than 2VDC (800 V), as depicted in Figure 7a.
In all cases, it was observed that the damping ratio of the drain–source voltage was higher during the second turn-on. This phenomenon occurs because before the second turn-on, the freewheeling diode of the switch, operating in parallel with inductance L, conducts the circulating current. Consequently, this diode effectively clamps the phase voltage to 2VDC in case 1, VDC in case 2, and 0 V in case 3. Prior to the first turn-on, as no current flows through the freewheeling diode, the drain–source voltage cannot be clamped by it.

6.3. Voltage Stress Evaluation

In this subsection, we focus on cases 2 and 3, examining the corresponding circuits depicted in Figure 22b,c, respectively. Theoretical analysis is conducted specifically for case 3, investigating the drain–source voltage (Vds) of SiC switch Q2 during the activation of Q1. To assess the impact of intrinsic capacitors, as depicted in Figure 8 and Figure 9, another switch with different capacitances—Cds, Cgs, and Cgd—compared to Q2, requires investigation. Hence, case 2 was explored alongside case 3 to gather experimental results for Si switch Q4, facilitating a comparison with the results of SiC switch Q2. Notably, the differences in drain–source characteristics between Q2 and Q4 are primarily attributed to the intrinsic capacitors, as both switches received identical gating pulses with matching turn-on/off times and dv/dt rates. Subsequent analysis compared the results with analytical findings, focusing on Cds as the sole varying parameter. Figure 25 illustrates the Vgs and Vds of switches Q2 and Q4 in case 2, along with switches Q1 and Q2 in case 3, highlighting the Vds at two rising edges of the gate–source pulses (indicated by dash loops) in Figure 26. Notably, Vds overshoot for Q2 in case 3 was 16% and 8% during the first and second rising edges, respectively, while these figures stood at 28% and 17% for Q4 in case 2. This indicates a higher damping ratio of Vds for Q2, attributed to its smaller Cds compared to Q4. Such observations align well with analytical results, as showcased in Figure 8a, where an increase in Cds up to 1.5 nF results in a decrease in damping ratio. Additionally, the natural frequency of Vds oscillations remains similar for both switches, owing to the closely matched values of Cgd, which significantly influences ωn. This correlation between experimental and analytical findings underscores the higher voltage overshoot during the first turn-on in both cases due to the absence of voltage clamping, a feature provided before the second turn-on by the switch’s body diode in parallel to the inductance, L. Overall, the observed overshoot amplitude and natural frequency in practice are generally lower than those predicted by theory, primarily due to anticipated parasitic effects associated with the PM layout and technology.
The analytical model exhibits a higher overshoot in the drain–source voltage compared to the experimental measurements, with the time needed to reach a steady state slightly prolonged in the experimental scenario. This suggests that the damping ratio in the experimental data surpasses that of the analytical model, while the natural frequency in the experiment was lower than predicted by the analytical model. This discrepancy is likely attributed to unaccounted-for experimental circuit parasitics in the analytical model.

6.4. Effects of dv/dt

Examining the effect of dv/dt involves altering the turn-on time of the switch, which is directly influenced by the turn-on gate resistance (RG,on). By reducing RG,on, an increase in dv/dt can be achieved. While previous plots were based on RG,on = 15.6 Ω, Figure 27 presents Vds oscillation overshoot for varying RG,on values. Specifically, the Vds2 waveform for a new RG,on = 11.5 Ω was compared with RG,on = 15.6 Ω during the first and second Q1 turn-on transients in case 3. It is evident that lower RG,on values result in higher voltage overshoot across the switch, indicating a correlation with increased dv/dt. Additionally, the overshoot amplitude is lower during the second turn-on transient for both turn-on gate resistance values due to this effect. Figure 28 illustrates the gate–source induced voltage for switch Q2 in case 3 for RG,on = 15.6 Ω and RG,on = 11.5 Ω. Notably, decreasing turn-on gate resistance leads to an increase in induced Vgs, thereby escalating the risk of false turn-on with a higher voltage change rate. Consequently, the peak value of Vgs is 2.4 V for RG,on = 15.6 Ω, whereas it rises to 3.2 V for RG,on = 11.5 Ω, exceeding the safe range. This experiment aims to determine the maximum allowable dv/dt considering the induced voltage that might trigger false turn-on, ensuring it remains below the switch’s turn-on threshold voltage. Finally, Figure 29 showcases the DPT test waveform in a comprehensive double pulse testing scenario with RG,on = 11.5 Ω.

7. Conclusions

An analytical methodology for evaluating drain–source voltage oscillation and false turn-on issue in T-Type converter with hybrid structure (SiC MOSFET, Si IGBT) has been presented. In this method, the equivalent circuit of the converter with consideration of parasitic capacitances and inductances is used to derive a mathematical model of the converter. The voltage stress and the possibility of false turn-on depend on the overshoot and duration of oscillations, which are a function of the damping ratio and natural frequency. Sensitivity analysis based on the developed model is carried out to investigate the impact of parasitic parameters on the drain–source and gate–source voltages’ damping ratio and natural frequency. The sensitivity analysis results show that the impact of parasitic capacitances on the voltage oscillations across both gate–source and drain–source is greater than the parasitic inductances. Furthermore, the most critical parameter affecting the possibility of false turn-on is the gain drain capacitance of the switch. The analytical results have been validated quantitatively with the simulation and experimental data. Experimental results show that the voltage overshoot across the switch reduces with decreasing of the drain–source capacitance. Also, increasing dv/dt by using a smaller gate turn-on resistor leads to higher induced gate–source voltage, which increases the possibility of false turn-on. The presented method can be beneficial in designing further high-voltage power converters to operate in the safe region using wide band gap switches which are suitable in high-frequency applications.

Author Contributions

The individual contribution of the authors have been stated as follows: Conceptualization, A.B., M.S.G. and N.C.; methodology, A.B. and N.C.; software, A.B., N.C., S.B. and J.B.; validation, A.B. and M.S.G.; formal analysis, A.B.; investigation, A.B., N.C. and M.B.; resources, T.E., N.C., S.B. and J.B.; data curation, A.B. and T.E.; writing—original draft preparation, A.B. and M.S.G.; writing—review and editing, M.B. and M.S.G.; supervision, A.B. and T.E.; project administration, T.E.; funding acquisition, T.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by funds from the Energy Technology Development and Demonstration Program (EUDP), via the project with ref.: 64020-2075. Find further information on EUDP here: https://www.eudp.dk/en/om-eudp (accessed on 1 April 2024).

Data Availability Statement

The authors confirm that no new data have been created in this work.

Conflicts of Interest

Author Nicklas Christensen was employed by the company Danfoss Drives A/S. Author Stefan Behrendt was employed by the company Semikron Danfoss. Author Jesco Beyer was employed by the company FuE-Zentrum FH Kiel GmbH. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Bhattacharya, S.; Narwal, R.; Shah, S.S.; Baliga, B.J.; Agarwal, A.; Kanale, A.; Han, K.; Hopkins, D.C.; Cheng, T.-H. Power Conversion Systems Enabled by SiC BiDFET Device. IEEE Power Electron. Mag. 2023, 10, 39–43. [Google Scholar] [CrossRef]
  2. Kumar, A.; Moradpour, M.; Losito, M.; Franke, W.-T.; Ramasamy, S.; Baccoli, R.; Gatto, G. Wide band gap devices and their application in power electronics. Energies 2022, 15, 9172. [Google Scholar] [CrossRef]
  3. Yadlapalli, R.T.; Kotapati, A.; Kandipati, R.; Koritala, C.S. A review on energy efficient technologies for electric vehicle applications. J. Energy Storage 2022, 50, 104212. [Google Scholar] [CrossRef]
  4. Mocevic, S.; Yu, J.; Fan, B.; Sun, K.; Xu, Y.; Stewart, J.; Rong, Y.; Song, H.; Mitrovic, V.; Yan, N. Design of a 10 kV SiC MOSFET-based high-density, high-efficiency, modular medium-voltage power converter. IEnergy 2022, 1, 100–113. [Google Scholar] [CrossRef]
  5. Cha, K.-H.; Ju, C.-T.; Kim, R.-Y. Analysis and evaluation of WBG power device in high frequency induction heating application. Energies 2020, 13, 5351. [Google Scholar] [CrossRef]
  6. Dini, P.; Ariaudo, G.; Botto, G.; Greca, F.L.; Saponara, S. Real-time electro-thermal modelling and predictive control design of resonant power converter in full electric vehicle applications. IET Power Electron. 2023, 16, 2045–2064. [Google Scholar] [CrossRef]
  7. Dini, P.; Basso, G.; Saponara, S.; Romano, C. Real-time monitoring and ageing detection algorithm design with application on SiC-based automotive power drive system. IET Power Electron. 2024. [Google Scholar] [CrossRef]
  8. Verbrugge, B.; Rasool, H.; Hasan, M.M.; Chakraborty, S.; Geury, T.; El Baghdadi, M.; Hegazy, O. Reliability Assessment of SiC-Based Depot Charging Infrastructure with Smart and Bidirectional (V2X) Charging Strategies for Electric Buses. Energies 2022, 16, 153. [Google Scholar] [CrossRef]
  9. Buffolo, M.; Favero, D.; Marcuzzi, A.; De Santi, C.; Meneghesso, G.; Zanoni, E.; Meneghini, M. Review and Outlook on GaN and SiC Power Devices: Industrial State-of-the-Art, Applications, and Perspectives. IEEE Trans. Electron Devices 2024, 71, 1344–1355. [Google Scholar] [CrossRef]
  10. Rasool, H.; Verbrugge, B.; Zhaksylyk, A.; Tran, T.M.; El Baghdadi, M.; Geury, T.; Hegazy, O. Design optimization and electro-thermal modeling of an off-board charging system for electric bus applications. IEEE Access 2021, 9, 84501–84519. [Google Scholar] [CrossRef]
  11. Dini, P.; Saponara, S. Electro-thermal model-based design of bidirectional on-board chargers in hybrid and full electric vehicles. Electronics 2021, 11, 112. [Google Scholar] [CrossRef]
  12. Dini, P.; Saponara, S.; Chakraborty, S.; Hosseinabadi, F.; Hegazy, O. Experimental Characterization & Electro-Thermal Modeling of Double Side Cooled SiC MOSFETs for Accurate and Rapid Power Converter Simulations. IEEE Access 2023, 11, 79120–79143. [Google Scholar]
  13. Tan, J.; Zhou, Z. An Optimized Switching Strategy Based on Gate Drivers with Variable Voltage to Improve the Switching Performance of SiC MOSFET Modules. Energies 2023, 16, 5984. [Google Scholar] [CrossRef]
  14. Wei, S.; Wen, W. High-Frequency Oscillation of the Active-Bridge-Transformer-Based DC/DC Converter. Energies 2022, 15, 3311. [Google Scholar] [CrossRef]
  15. Khanna, R.; Amrhein, A.; Stanchina, W.; Reed, G.; Mao, Z.-H. An analytical model for evaluating the influence of device parasitics on Cdv/dt induced false turn-on in SiC MOSFETs. In Proceedings of the 2013 Twenty-Eighth Annual IEEE Applied Power Electronics Conference and Exposition (APEC), Long Beach, CA, USA, 17–21 March 2013. [Google Scholar]
  16. Zhao, Q.; Stojcic, G. Characterization of Cdv/dt induced power loss in synchronous buck DC-DC converters. In Proceedings of the Nineteenth Annual IEEE Applied Power Electronics Conference and Exposition, 2004. APEC’04, Anaheim, CA, USA, 22–26 February 2004. [Google Scholar]
  17. Peters, D.; Aichinger, T.; Basler, T.; Rescher, G.; Puschkarsky, K.; Reisinger, H. Investigation of threshold voltage stability of SiC MOSFETs. In Proceedings of the 2018 IEEE 30th International Symposium on Power Semiconductor Devices and ICs (ISPSD), Chicago, IL, USA, 13–17 May 2018. [Google Scholar]
  18. Li, R.; Zhu, Q.; Xie, M. A new analytical model for predicting dv/dt-induced low-side MOSFET false turn-ON in synchronous buck converters. IEEE Trans. Power Electron. 2018, 34, 5500–5512. [Google Scholar] [CrossRef]
  19. Dalal, D.N.; Christensen, N.; Jørgensen, A.B.; Jørgensen, J.K.; Bęczkowski, S.; Munk-Nielsen, S.; Uhrenfeldt, C. Impact of power module parasitic capacitances on medium-voltage SiC MOSFETs switching transients. IEEE J. Emerg. Sel. Top. Power Electron. 2019, 8, 298–310. [Google Scholar] [CrossRef]
  20. Liu, B.; Ren, R.; Jones, E.A.; Gui, H.; Zhang, Z.; Chen, R.; Wang, F.; Costinett, D. Effects of junction capacitances and commutation loops associated with line-frequency devices in three-level ac/dc converters. IEEE Trans. Power Electron. 2018, 34, 6155–6170. [Google Scholar] [CrossRef]
  21. Xie, W.; Ji, S.; Zhao, Z.; Mo, X. SiC MOSFET Crosstalk Modelling with Suppression Considering Impacts of dv/dt and di/dt. In Proceedings of the 2023 11th International Conference on Power Electronics and ECCE Asia (ICPE 2023-ECCE Asia), Jeju Island, Republic of Korea, 22–25 May 2023. [Google Scholar]
  22. Khanna, R.; Stanchina, W.; Reed, G. Effects of parasitic capacitances on gallium nitride heterostructure power transistors. In Proceedings of the 2012 IEEE Energy Conversion Congress and Exposition (ECCE), Raleigh, NC, USA, 15–20 September 2012. [Google Scholar]
  23. Wei, S.; Zhao, Z.; Yuan, L.; Wen, W.; Chen, K. Voltage oscillation suppression for the high-frequency bus in modular-multiactive-bridge converter. IEEE Trans. Power Electron. 2021, 36, 9737–9742. [Google Scholar] [CrossRef]
  24. HIZARCI, H.; KALAYCI, K.; Demirel, O.; Arifoğlu, U. Reducing electromagnetic interference in three-level T-type isolated bidirectional DC-DC converter using a snubber circuit. Int. J. Appl. Math. Electron. Comput. 2021, 9, 26–34. [Google Scholar] [CrossRef]
  25. Zhang, M.; Ren, N.; Guo, Q.; Sheng, K. Understanding Turn-On Transients of SiC High-Power Modules: Drain-Source Voltage Plateau Characteristics. Energies 2020, 13, 3802. [Google Scholar] [CrossRef]
  26. Qin, H.; Ma, C.; Wang, D.; Xie, H.; Zhu, Z.; Xu, K.; Wang, S. An overview of SiC MOSFET gate drivers. In Proceedings of the 2017 12th IEEE Conference on Industrial Electronics and Applications (ICIEA), Siem Reap, Cambodia, 18–20 June 2017. [Google Scholar]
  27. Ma, C.-T.; Gu, Z.-H. Review on driving circuits for wide-bandgap semiconductor switching devices for mid-to high-power applications. Micromachines 2021, 12, 65. [Google Scholar] [CrossRef] [PubMed]
  28. Gao, F.; Zhou, Q.; Wang, P.; Zhang, C. A gate driver of SiC MOSFET for suppressing the negative voltage spikes in a bridge circuit. IEEE Trans. Power Electron. 2017, 33, 2339–2353. [Google Scholar] [CrossRef]
  29. Babaki, A.; Vaez-Zadeh, S.; Jahanpour-Dehkordi, M.; Zakerian, A. Wireless motor drives with a single inverter in primary side of power transfer systems. In Proceedings of the 2019 IEEE PELS Workshop on Emerging Technologies: Wireless Power Transfer (WoW), London, UK, 18–21 June 2019. [Google Scholar]
  30. Okilly, A.H.; Baek, J. Design and Fabrication of an Isolated Two-Stage AC–DC Power Supply with a 99.50% PF and ZVS for High-Power Density Industrial Applications. Electronics 2022, 11, 1898. [Google Scholar] [CrossRef]
  31. Liu, S.; Lin, H.; Wang, T. Comparative study of three different passive snubber circuits for SiC power MOSFETs. In Proceedings of the 2019 IEEE Applied Power Electronics Conference and Exposition (APEC), Anaheim, CA, USA, 17–21 March 2019. [Google Scholar]
  32. Dzhunusbekov, E.; Orazbayev, S. A new passive lossless snubber. IEEE Trans. Power Electron. 2021, 36, 9263–9272. [Google Scholar] [CrossRef]
  33. Behera, S.; Dash, S.K.; Sahu, M.K.; Sahu, I.; Parida, S. Design and Development of a New Soft-Switching Buck Converter. In Proceedings of the 2023 International Conference on Power Electronics and Energy (ICPEE), Bhubaneswar, India, 3–5 January 2023. [Google Scholar]
  34. Kasasbeh, A.; Kelleci, B.; Ozturk, S.B.; Aksoz, A.; Hegazy, O. Sepic converter with an LC regenerative snubber for ev applications. Energies 2020, 13, 5765. [Google Scholar] [CrossRef]
  35. Pastor, M.; Dudrik, J.; Michal, R. High-Frequency Soft-Switching DC-DC Converter with Simple Secondary Turn-Off Snubber. In Proceedings of the 2022 IEEE 20th International Power Electronics and Motion Control Conference (PEMC), Brasov, Romania, 25–28 September 2022. [Google Scholar]
  36. Christensen, N. Demonstration of High Power Density kW Converters Utilizing Wide-Band Gap Devices. Ph.D. Thesis, Aalborg University, Aalborg, Denmark, 2019. [Google Scholar]
  37. Mukherjee, S.; Ruiz, J.M.; Barbosa, P. A high power density wide range DC–DC converter for universal electric vehicle charging. IEEE Trans. Power Electron. 2022, 38, 1998–2012. [Google Scholar] [CrossRef]
  38. Liu, H.; Wang, Y.; Wheeler, P.; Zhou, X.; Zhu, K. A voltage spike suppression strategy based on De-Re-coupling idea for the three-phase high-frequency isolated matrix-type inverter. IEEE Trans. Power Electron. 2022, 37, 9024–9034. [Google Scholar] [CrossRef]
  39. Athwer, A.; Darwish, A. A Review on Modular Converter Topologies Based on WBG Semiconductor Devices in Wind Energy Conversion Systems. Energies 2023, 16, 5324. [Google Scholar] [CrossRef]
  40. Bonyadi, R.; Alatise, O.; Jahdi, S.; Hu, J.; Gonzalez, J.A.O.; Ran, L.; Mawby, P.A. Compact electrothermal reliability modeling and experimental characterization of bipolar latchup in SiC and CoolMOS power MOSFETs. IEEE Trans. Power Electron. 2015, 30, 6978–6992. [Google Scholar] [CrossRef]
  41. Heng, K.; Yang, X.; Wu, X.; Ye, J.; Liu, G. A temperature-dependent physical thermal network model including thermal boundary conditions for SiC MOSFET module. IEEE Trans. Electron Devices 2022, 69, 4444–4452. [Google Scholar] [CrossRef]
  42. Dalal, D.N.; Zhao, H.; Jørgensen, J.K.; Christensen, N.; Jørgensen, A.B.; Bȩczkowski, S.; Uhrenfeldt, C.; Munk-Nielsen, S. Demonstration of a 10 kV SiC MOSFET based medium voltage power stack. In Proceedings of the 2020 IEEE Applied Power Electronics Conference and Exposition (APEC), New Orleans, LA, USA, 15–19 March 2020. [Google Scholar]
  43. Rodríguez, M.; Rodríguez, A.; Miaja, P.F.; Lamar, D.G.; Zúniga, J.S. An insight into the switching process of power MOSFETs: An improved analytical losses model. IEEE Trans. Power Electron. 2010, 25, 1626–1640. [Google Scholar] [CrossRef]
Figure 1. A schematic of a single-phase active T-Type converter.
Figure 1. A schematic of a single-phase active T-Type converter.
Electronics 13 01808 g001
Figure 2. Switching modulation of T-Type.
Figure 2. Switching modulation of T-Type.
Electronics 13 01808 g002
Figure 3. The equivalent circuit for Q2 false turn-on demonstration considering parasitic capacitances.
Figure 3. The equivalent circuit for Q2 false turn-on demonstration considering parasitic capacitances.
Electronics 13 01808 g003
Figure 4. Simple circuit model of Q2 in the OFF state.
Figure 4. Simple circuit model of Q2 in the OFF state.
Electronics 13 01808 g004
Figure 5. Cgd dV/dt test; (a) circuit model with all possible parasitic elements, (b) complete circuit model.
Figure 5. Cgd dV/dt test; (a) circuit model with all possible parasitic elements, (b) complete circuit model.
Electronics 13 01808 g005
Figure 6. (a) PM simulated in Ansys Q3D, (b) middle track, (c) upper side (+) track, and (d) lower side (−) track.
Figure 6. (a) PM simulated in Ansys Q3D, (b) middle track, (c) upper side (+) track, and (d) lower side (−) track.
Electronics 13 01808 g006aElectronics 13 01808 g006b
Figure 7. Step response of (a) Vds across Q2, and (b) Vgs of Q2 during Q1 turn-on.
Figure 7. Step response of (a) Vds across Q2, and (b) Vgs of Q2 during Q1 turn-on.
Electronics 13 01808 g007
Figure 8. Damping ratio of ringing across the switch in terms of parasitic capacitor values; (a) Cds2, (b) Cgd2, and (c) Cgs2.
Figure 8. Damping ratio of ringing across the switch in terms of parasitic capacitor values; (a) Cds2, (b) Cgd2, and (c) Cgs2.
Electronics 13 01808 g008
Figure 9. Natural frequency of oscillation across the switch in terms of parasitic capacitor values; (a) Cds2, (b) Cgd2, and (c) Cgs2.
Figure 9. Natural frequency of oscillation across the switch in terms of parasitic capacitor values; (a) Cds2, (b) Cgd2, and (c) Cgs2.
Electronics 13 01808 g009
Figure 10. Vds oscillation for Cgd = 2 nF.
Figure 10. Vds oscillation for Cgd = 2 nF.
Electronics 13 01808 g010
Figure 11. Damping ratio of ringing voltage across the switch in terms of the stray inductances in the circuit; (a) Ltotal, (b) LG, and (c) Ll−.
Figure 11. Damping ratio of ringing voltage across the switch in terms of the stray inductances in the circuit; (a) Ltotal, (b) LG, and (c) Ll−.
Electronics 13 01808 g011
Figure 12. Natural frequency of oscillation across the switch (Drain–Source) in terms of the stray inductances in the circuit; (a) Ltotal, (b) LG, and (c) Ll−.
Figure 12. Natural frequency of oscillation across the switch (Drain–Source) in terms of the stray inductances in the circuit; (a) Ltotal, (b) LG, and (c) Ll−.
Electronics 13 01808 g012
Figure 13. Effect of each parasitic capacitance of the switch on the induced gate–source voltage; (a) Cds, (b) Cgd, and (c) Cgs.
Figure 13. Effect of each parasitic capacitance of the switch on the induced gate–source voltage; (a) Cds, (b) Cgd, and (c) Cgs.
Electronics 13 01808 g013
Figure 14. Vgs of the OFF switch in case that Cgs = 13 nF.
Figure 14. Vgs of the OFF switch in case that Cgs = 13 nF.
Electronics 13 01808 g014
Figure 15. Vgs of the OFF switch in case that Cds = 5 nF.
Figure 15. Vgs of the OFF switch in case that Cds = 5 nF.
Electronics 13 01808 g015
Figure 16. Three-phase T-Type converter.
Figure 16. Three-phase T-Type converter.
Electronics 13 01808 g016
Figure 17. Switching state changes and the resulted phase voltage.
Figure 17. Switching state changes and the resulted phase voltage.
Electronics 13 01808 g017
Figure 18. Simplified circuit considering switching state changes of legs 2 and 3 in form of pulse voltage.
Figure 18. Simplified circuit considering switching state changes of legs 2 and 3 in form of pulse voltage.
Electronics 13 01808 g018
Figure 19. Voltage oscillation of Vph1 because of switching state changes for other phases (Ph2 and Ph3) for; (a) θ1 = 0°; (b) θ1 = 30°; and (c) θ1 = 60°.
Figure 19. Voltage oscillation of Vph1 because of switching state changes for other phases (Ph2 and Ph3) for; (a) θ1 = 0°; (b) θ1 = 30°; and (c) θ1 = 60°.
Electronics 13 01808 g019
Figure 20. The experimental setup for result verification by DPT test.
Figure 20. The experimental setup for result verification by DPT test.
Electronics 13 01808 g020
Figure 21. Current flow of the system under DPT test in practice.
Figure 21. Current flow of the system under DPT test in practice.
Electronics 13 01808 g021
Figure 22. The equivalent circuit model under DPT test for (a) upper side half-bridge of T-type converter (case 1), (b) lower side half-bridge of T-type converter (case 2), and (c) T-type leg (case 3).
Figure 22. The equivalent circuit model under DPT test for (a) upper side half-bridge of T-type converter (case 1), (b) lower side half-bridge of T-type converter (case 2), and (c) T-type leg (case 3).
Electronics 13 01808 g022
Figure 23. Induced gate–source voltage of switches with possibility of false turning on in DPT test; (a) case 1, (b) case 2, (c) case 3.
Figure 23. Induced gate–source voltage of switches with possibility of false turning on in DPT test; (a) case 1, (b) case 2, (c) case 3.
Electronics 13 01808 g023
Figure 24. Enlarged view of the induced gate–source voltage of Q2 in DPT test of case 3; (a) first turn-on of Q1, (b) second turn-on of Q1.
Figure 24. Enlarged view of the induced gate–source voltage of Q2 in DPT test of case 3; (a) first turn-on of Q1, (b) second turn-on of Q1.
Electronics 13 01808 g024
Figure 25. (a) Vds and Vgs of the switches in case 2, and (b) Vds and Vgs of the switches in case 3.
Figure 25. (a) Vds and Vgs of the switches in case 2, and (b) Vds and Vgs of the switches in case 3.
Electronics 13 01808 g025
Figure 26. (a) Vds across switches Q2 and Q4 for case 2 and 3 during first turning-on, and (b) Vds across switches Q2 and Q4 for case 2 and 3 during second turning-on.
Figure 26. (a) Vds across switches Q2 and Q4 for case 2 and 3 during first turning-on, and (b) Vds across switches Q2 and Q4 for case 2 and 3 during second turning-on.
Electronics 13 01808 g026
Figure 27. Comparison results of different voltage changes rate using different RG,on; (a) first turn-on of switch Q1and (b) second turn-on of switch Q1.
Figure 27. Comparison results of different voltage changes rate using different RG,on; (a) first turn-on of switch Q1and (b) second turn-on of switch Q1.
Electronics 13 01808 g027
Figure 28. Induced Vgs of Q2 in case 3 for different RG,on; (a) RG,on = 11.5 Ω and (b) RG,on = 15.6 Ω.
Figure 28. Induced Vgs of Q2 in case 3 for different RG,on; (a) RG,on = 11.5 Ω and (b) RG,on = 15.6 Ω.
Electronics 13 01808 g028
Figure 29. The overall key waveforms of DPT test with replacing RG = 15.6 Ω by RG = 11.5 Ω.
Figure 29. The overall key waveforms of DPT test with replacing RG = 15.6 Ω by RG = 11.5 Ω.
Electronics 13 01808 g029
Table 1. System parameters of the system model.
Table 1. System parameters of the system model.
ParameterDefinitionValueParameterDefinitionValue
L+Upper side conductor leakage inductance12.5 nHLLower side conductor inductance12.5 nH
Rleak+Upper side conductor leakage resistance12.5 mΩRleak.−Lower side conductor resistance12.5 mΩ
LmidLeg inductance between the switches5 nHRmidLeg resistance between the switches5 mΩ
Ll+Leg inductance on top of the upper switch5 nHLlLeg inductance under the lower switch5 nH
Rl+Leg resistance on top of the upper switch5 mΩRlLeg resistance under the lower switch5 mΩ
LG2Gate on/off stray inductance0.5 nHRG2,onGate ON-resistance15.6 Ω
Cgd2Gate–Drain capacitor for SiC10 pFCgs2Gate–Source capacitor for SiC1300 pF
Cds2Drain–Source capacitor for SiC50 pFRon,Q1SiC MOSFET on resistance75 mΩ
Table 2. Polarity of reference voltage for each phase.
Table 2. Polarity of reference voltage for each phase.
Vref for:
Phase 1/SignPhase 2/SignPhase 3/Sign
θ1 = 0–60°+120°–180°+240°–300°-
θ1 = 60°–120°+180°–240°-300°–360°-
θ1 = 120°–180°+240°–300°-0–60°+
Table 3. Parameters of the system presented in Figure 18.
Table 3. Parameters of the system presented in Figure 18.
ParameterDefinitionValueParameterDefinitionValue
L+Upper side conductor leakage inductance 40 nHLConductor leakage inductance 40 nH
Ron,11SiC MOSFET ON-resistance75 mΩRon,14Si IGBT ON-resistance75 mΩ
Cds31, Cds22, Cds12SiC MOSFET
drain–source capacity
550 pFCds13Si IGBT
drain–source capacity
55 pF
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Babaki, A.; Golsorkhi, M.S.; Christensen, N.; Baharizadeh, M.; Behrendt, S.; Beyer, J.; Ebel, T. Parasitic-Based Model for Characterizing False Turn-On and Switching-Based Voltage Oscillation in Hybrid T-Type Converter. Electronics 2024, 13, 1808. https://doi.org/10.3390/electronics13101808

AMA Style

Babaki A, Golsorkhi MS, Christensen N, Baharizadeh M, Behrendt S, Beyer J, Ebel T. Parasitic-Based Model for Characterizing False Turn-On and Switching-Based Voltage Oscillation in Hybrid T-Type Converter. Electronics. 2024; 13(10):1808. https://doi.org/10.3390/electronics13101808

Chicago/Turabian Style

Babaki, Amir, Mohammad Sadegh Golsorkhi, Nicklas Christensen, Mehdi Baharizadeh, Stefan Behrendt, Jesco Beyer, and Thomas Ebel. 2024. "Parasitic-Based Model for Characterizing False Turn-On and Switching-Based Voltage Oscillation in Hybrid T-Type Converter" Electronics 13, no. 10: 1808. https://doi.org/10.3390/electronics13101808

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop