Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

DOE Award #: DE-EE0006849

Brushless and Permanent Magnet Free Wound Field


Synchronous Motors for EV Traction

Daniel C. Ludois
University of Wisconsin -Madison
1415 Engineering Dr.
Madison, WI 53706
Email: ludois@wisc.edu
Phone: 608-262-8211

Team Organizations:

University of Wisconsin – Madison


Illinois Institute of Technology

Final Technical Report, 3/31/2017


Report Coverage: 10/1/2014 – 12/31/2016
1) Abstract/Executive Summary

 Rare-earth permanent magnets (PMs) have been subject to market volatility and are
largely single sourced from a foreign power. Permanent magnets in EV traction interior
permanent magnet synchronous motors (IPMSMs) are a significant fraction of the cost
and impose temperature limitations.
 Additionally, the PMs provide a fixed flux level which is always, "on", leading to safety
concerns during inverter faults and requiring additional current to be injected into the
machine during field weakening to buck the magnetic flux. This additional current
lowers the power factor of the machine, requires that the traction inverter be oversized to
supply the reactive current, and leads to increased ohmic losses in the stator and inverter.
Induction machines (IMs) must also draw reactive current from the traction inverter to
magnetize the machine leading to a lower power factor.
 Wound field synchronous motors (WFSM) are a potentially advantageous alternative to
commercially dominant IPMSMs and IMs for electric vehicle traction due to the
following properties.
 Permanent magnet free
 Complete control of the field excitation with potential for optimal field weakening
and loss minimization over driving cycles.
 Field magnetization can be drawn from the rotor side increasing the power factor
of the machine, lowering volt-amp (VA) burden on the traction inverter.
 EV traction applications require extremely high reliability and power density reducing
the viable use of brushes and other classical field power transfer technologies such as
brushless exciters. Capacitive power transfer (CPT) technology offers an attractive
means of providing brushless power transfer to the rotor field windings.

2) Accomplishments & Outcomes

 A rapid WFSM magnetic and thermal design optimization environment has been created.
 Two capacitive power coupler (CPC) systems were designed, constructed, and
demonstrated; one for low speed (oil film journal bearing style) and one for high speed
operation (air-gapped parallel plates).
 High performance WFSM prototypes were designed, prototyped, and experimentally
characterized with brushes and slip rings and brushless capacitive power couplers (CPC).
 The peak volumetric and specific power density exceeded USDRIVE 2020
targets.
 The prototype machine exhibited a broad band of very high efficiency (>95%)
and high power factor.
 Operation with the air-gapped paralleled plates CPC was demonstrated to 55 kW
power output at 4000 RPM.

1
 A kilowatt scale, 2 MHz switching CPC power electronics driver was demonstrated.
 Voltage and field current regulation during generator operation was demonstrated with a
journal style bearing CPC.
 A new prototype rotor has been designed which further increases the torque and power
densities while lowering field winding power transfer requirements.

TABLE 1: HIGHEST MEASURED PERFORMANCE ACHIEVED FROM A WFSM PROTOTYPE IN THE PROJECT
Machine Parameters Experimental Value
Peak Motoring Torque / Power* 190.55 Nm / 79.6 kW * @ 4000 RPM
Peak Motoring Efficiency / Torque 95 % @ 87.5 Nm @ 4000 RPM
Mass** 40.64 kg
Volume*** 11.06 ℓ
Volumetric Peak Torque Density*** 17.22 Nmℓ-1
Specific Peak Torque Density** 4.69 Nm kg-1
Volumetric Peak Power Density*** 7.19 kW ℓ-1
Specific Peak Power Density** 1.95 kW kg-1
* Continuous, held for over 30 seconds
** Includes stator, rotor and shaft mass (shaft is oversized)
*** Cilindrical volume of active materials plus ATF spray cooling rings

The content of this project has been published in peer reviewed IEEE publications:
1. J. Dai; S. Hagen; D. Ludois; I. Brown, "Synchronous Generator Brushless Field
Excitation and Voltage Regulation Via Capacitive Coupling Through Journal Bearings,"
in IEEE Transactions on Industry Applications, early access online, to appear in
July/August edition 2017.
2. Di Gioia, A., Brown I.P., Knippel R., Hagen S., Ludois D.C., Nie Y., Dai J.J., Alteheld,
C., "Design of a wound field synchronous machine for electric vehicle traction with
brushless capacitive field excitation," 2016 IEEE Energy Conversion Congress and
Exposition (ECCE), Milwaukee, WI, 2016, pp. 1-8.
3. J. Dai, S. Hagen, D. C. Ludois and I. P. Brown, "Synchronous generator field excitation
via capacitive coupling through a journal bearing," 2016 IEEE Energy Conversion
Congress and Exposition (ECCE), Milwaukee, WI, 2016, pp. 1-8.
4. J. Dai and D. C. Ludois, "Capacitive Power Transfer Through a Conformal Bumper for
Electric Vehicle Charging," in IEEE Journal of Emerging and Selected Topics in Power
Electronics, vol. 4, no. 3, pp. 1015-1025, Sept. 2016.
5. J. Dai and D. C. Ludois, "A Survey of Wireless Power Transfer and a Critical
Comparison of Inductive and Capacitive Coupling for Small Gap Applications," in IEEE
Transactions on Power Electronics, vol. 30, no. 11, pp. 6017-6029, Nov. 2015.
6. S. Hagen, R. Knippel, J. Dai and D. C. Ludois, "Capacitive coupling through a
hydrodynamic journal bearing to power rotating electrical loads without contact," 2015
IEEE Wireless Power Transfer Conference (WPTC), Boulder, CO, 2015, pp. 1-4.

2
3) Introduction

This project designed, prototyped, and demonstrated a brushless and permanent magnet free
wound field synchronous motor (WFSM) for EV traction. The rotor field winding was excited
using non-contact capacitive power transfer (CPT) technology. In an EV traction application, a
brushless and permanent magnet free WFSM would have several potentially significant benefits
and advantages over commercially dominant interior permanent magnet synchronous motors
(IPMSMs) and induction motors (IMs) including reduced cost through the removal of rare earth
permanent magnets (IPMSMs), higher system and machine efficiencies though power factor
improvements (IPMSMs and IMs), loss minimization field control, improved safety through
field control during inverter fault conditions, and the possibility for power take-off and microgrid
support. The unique enabling technology for this project is the CPT technology which allows
brushless operation of the WFSM in alternative form factors.
Field weakening for a large constant power speed range (CPSR) can be achieved with the proper
design of the IPMSM, i.e. the magnet flux can be bucked by the d-axis armature flux. This is a
non-ideal solution however as the flux produced by expensive rare-earth permanent magnets is
being bucked by injecting a large reactive current, -id, into the machine, also reducing the power
factor. To achieve this, the traction inverter kVA rating must be oversized increasing the cost
and size of the entire EV electric drive. The DOE has estimated that the power factor of IPMSM
increases the cost of the traction inverter by 15% [1]. Positive reactive current, id, is also drawn
by IMs to magnetize the machine reducing the machines power factor.
The main defining feature of WFSM is the complete control of the field excitation from the rotor
side. Near unity power factor and optimal field weakening can be achieved over the entire
torque-speed range because of the direct handle on the field excitation. This control reduces
stress on the inverter and other electronic components because additional reactive power for field
weakening is not necessarily cycled through the inverter. The efficiency of the inverter increases
and the inverter cost and volume can be minimized. The complete control of the field excitation
in WFSM also brings other advantages including an extra control input for loss minimization at
both a machine and system level. Field excitation provided by windings instead of magnets also
removes the risk of demagnetization at high temperatures which occurs in IPMSMs which
require expensive heavy rare earth elements to be added to increase the magnet temperature
grade.
Given the potential advantages of the WFSM in traction applications, why has it not seen wide
spread use? Historically, power has been provided to the WFSM field winding using three
methods: slip rings, brushless exciters, and rotating transformers. Recently three other brushless
methods of transferring power to the rotor field winding have been developed: harmonic,
inductive and capacitive power transfer. Slip rings, while being a well understood technology,
are not well suited to an automotive environment because of reliability and maintenance
concerns (brushes need to be changed and introduce brush dust). Brushless exciters and rotating
transformers add significant weight and shaft length to the overall machine. Power transfer of
brushless exciters and rotating transformers is also speed dependent and introduces extra system
dynamics (L/R time constant). Circuit representations of inductive and capacitive power transfer
systems are shown in Fig. 1. Compared to inductive power transfer, CPT technology can operate

3
at higher speeds (simple rotating structure) and efficiency (no windings). CPT has similar system
dynamics to slip rings. Another CPT advantage is accurate rotor current sensing from the stator
side (since there is no operating point dependent magnetizing branch as would exist in a
transformer). The key enabling technology for leveraging the attractive attributes of WFSM as
EV traction motors is the CPT.

Fig. 1. Circuit representation of field coupling for brushless (a) inductive and (b) capacitive couplers.

In this project an incremental development approach has been taken with only one component of
the overall motor and excitation system changed at one time, Fig. 2. Two prototype WFSM
stators and rotors were designed and prototyped along with two sets of excitation systems, i.e.
two sets of slip rings and brushes and capacitive power couplers. Using slip rings and brushes
decoupled the development of the WFSM from the capacitive power coupler development
allowing each to be tested individually before uniting them.

4
Incremental Development Approach
WFSM Prototype 1 WFSM Prototype 1 WFSM Prototype 2 WFSM Prototype 2 WFSM Prototype 2
Stator Stator Stator Stator Stator

WFSM Prototype 1 WFSM Prototype 1 WFSM Prototype 1 WFSM Prototype 2 WFSM Prototype 2
Rotor Rotor Rotor Rotor Rotor

Slip Rings and Slip Rings and


CPC Prototype 1 CPC Prototype 1 CPC Prototype 2
Brushes 1 Brushes 2

Fig. 2. Incremental development of WFSM prototype stators, rotors, and capactive power couplers.

4) WFSM Multi-Objective Optimization Environment

In order to design an optimized WFSM which rivals or exceeds the performance of state of the
art IPMSMs and IMs using a CPC for rotor field excitation, a flexible design environment has
been developed. A mechanical, thermal and electromagnetic parametric model of the WFSM has
been employed to explore the design space, and a differential evolution optimization algorithm to
find the best candidate for realization.
Electromagnetic and thermal evaluation of WFSM designs has been carried out, centered on a
parametric geometry engine implemented in MATLAB and interfaced to external FEA and
thermal equivalent circuit software (FEMM, Infolytica MagNet and MotorCAD) via an ActiveX
interface, conceptually represented in Fig.3. The centralized interface allows for parametric rotor
and stator geometry generation, material assignment, current loading and selection between static
and transient magnetic simulations. After the simulations are carried out with external packages,
the data conditioning and performance evaluation routines are executed in the MATLAB
environment as part of the full population based optimization using the differential evolution
algorithm. The resulting geometry and losses can be exported to the Motor-CAD package for
thermal modeling. A considerable portion of the development effort was devoted to
parallelization, with multiple simulation instances, to speed up the optimization process.
Attention and care was paid to ensure that the electromagnetic simulation packages (FEMM and
MagNet) have comparable results through control of the meshing and material properties.
Critical to the full exploration of the WFSM design space is the use of a parametric geometry
engine using dimensional and non-dimensional geometric quantities (stator and rotor) which
does not allow non-physical geometries to occur when using a population based design
optimization algorithm.

5
Fig. 3. Flexible WFSM design environment.

The WFSM parametric geometry is characterized by only four dimensional parameters: the stack
length, the outer stator radius, the minimum airgap length and the shaft radius (see Fig. 4, Fig. 5,
Table 1 for reference,), all other dimensions are defined by the 4 dimensional and 24 non-
dimensional ratios of geometric entities. Once the machine pole pairs, p, and number of stator
slots, z, are defined, the pole and slot pitch are fixed, and all other parameters are non-
dimensional. A parameter hierarchy has been established in order to generate a unique geometry
case. The non-dimensional ratios, k, are defined in the nomenclature.
In the stator template of Fig. 4, the stator inner radius, Rsi, and the stator yoke depth, dSy, are
calculated first from two split ratios, kSi and kSy, of the stator outer radius, Rso

𝑅𝑆𝑖 = 𝑘𝑆𝑖 𝑅𝑆𝑜 (1.)

(1−𝑘𝑆𝑖 )
𝑑𝑆𝑦 = 𝑘𝑆𝑦 𝑅𝑆𝑜 (2.)
(1+𝑘𝑆𝑦 )

Then a stator tooth thickness, wSt, and stator slot opening, wSo, are imposed with two additional
ratios, kst and kSo, and half stator slot angular span, αs. The values of these geometric entities are
limited by the intersection of inner stator and parallel tooth surfaces.

𝑤𝑆𝑡 = 𝑅𝑆𝑖 sin(𝑘𝑆𝑡 𝛼𝑠 ) (3.)

𝑆𝑡 𝑤
𝑤𝑆𝑜 = ( 𝑅𝑆𝑖 tan(𝛼𝑠 ) − cos(𝛼 )
) 𝑘𝑆𝑜 (4.)
𝑠

The stator slot depth is also calculated from these parameters, since the tooth dimensions are
needed to calculate surfaces and volumes used to derive the iron and winding window surfaces.
(1−𝑘 )
𝑑𝑆𝑡 = 𝑅𝑆𝑜 (1+𝑘 𝑆𝑖 ) (5.)
𝑆𝑦

After the main magnetic path is defined (via tooth and yoke sizing) the slot bottom radius is
derived (intersecting a 90 degrees spanning arc, that starts at the yoke inner surface, with the
parallel tooth), finally a tooth tip thickness and angle are imposed to link the inner stator to the
straight tooth.

6
The rotor template has been designed with a similar concept, but additional flexibility has been
added, allowing points to merge and collapse. Fig. 6 shows the four possible types of rotor shape
for each pole shoe surface, allowing the exploration of different electromagnetic and mechanical
tradeoffs, since the airgap length is the same for different cross sectional areas of the pole shoe,
the stress concentration will also be affected. The rotor geometry type is obtained following the
flowchart in 7. The common feature of the geometry is the pole shape, defined by a minimum
airgap thickness,𝛿0 , and an inverse cosine pole function.
𝛿
0
𝑅𝑅𝑜 (𝜃𝑝 ) = ( 𝑅𝑆𝑖 − cos(𝜃 )
) (6.)
𝑝

The rotor tip point will be characterized by the maximum airgap size (𝛿𝑀𝐴𝑋 in Fig. 5), by a pole
thickness and a distance from the magnetic q-axis (respectively 𝑙𝐶 and
𝑑𝑅𝑞,𝐴 in Fig. 5).
𝛿0
𝑅𝑅𝑡 = ( 𝑅𝑆𝑖 − ) (7.)
cos(𝜃𝑀𝐴𝑋 )
𝑅𝑅𝑡
𝑑𝑅𝑞,𝐴 = sin(𝛼𝑝 )𝑘𝑅𝑞 (8.)
tan(𝜃𝑀𝐴𝑋 )
In order to meet contrasting requirements of mechanical stiffness (bigger pole body) and
reduction of rotor leakage flux (smaller pole body), the geometry can morph into one of the four
cases according to two parameters: the distance from the rotor interpolar axis (magnetic q-axis
leading rotor d-axis by 90 elec. degrees) and the pole thickness. If the initial pole shape meets
both constraints, a type A rotor is finalized, imposing a neck thickness parameter (defined as a
fraction of the pole rotor tip height) and a yoke thickness (defined as a fraction of the pole rotor
tip radial distance from the shaft center).

ℎ𝑅𝑛 = 𝑅𝑅𝑡 sin(𝜃𝑀𝐴𝑋 )𝑘𝑅𝑛 (9.)

𝑅𝑅𝑦 = 𝑅𝑠ℎ + 𝑘𝑅𝑦 (𝑅𝑅𝑡 − 𝑅𝑠ℎ ) (10.)


If the rotor pole thickness is smaller than the desired value, 𝑙𝑡ℎ , types B and C are tested until the

interpolar axis distance is simultaneously met (𝑑𝑅𝑞 ≥ 𝑑𝑅𝑞 ). If the thickness of type A doesn’t
meet the mechanical requirement, a horizontal segment is added to the pole body, if the case B
thickness, 𝑙𝐵 , is not enough, another segment parallel to the magnetic q-axis, is added. This
process allows the requirements to be met simultaneously if a solution is possible. Case D is
conceptually similar to case C, where the q-axis distance is met by definition, substituting the

rotor pole tip in case A with one that meets the desired distance 𝑑𝑅𝑞 from the magnetic q-axis.

The desired thickness 𝑙𝑡ℎ , has been derived from mechanical modeling of the rotor structure
using a design of experiments approach and is internally evaluated in MATLAB.
After the dimensional geometry is generated from the input parameters, MATLAB draws it in
the linked simulation software and assigns the material properties in the solvers. Since the
maximum torque for a given excitation depends on the current angle, a series of five static
simulations identifies the maximum torque current angle before a full time stepping transient
(MagNet) or equivalent static reconstructed (FEMM) simulations are carried out to evaluate the
speed-dependent variables (torque ripple, flux linkage, voltage per turn, core losses etc.). Core
losses in the stator and rotor pole face are estimated using modified Steinmetz and CAL2 loss
models [2]. Care has been taken to ensure that essentially similar results are obtained between

7
simulation programs and techniques. For example, in MagNet the iron loss is estimated in each
element in the FEA simulation, while in FEMM sampling by regions has been implemented in
order to closely approximate the same iron loss estimation. After post-processing, all the output
data is nested in a MATLAB data structure. The main FEMM advantage is that the
parallelization is relatively easy and is not affected by licensing issues.
After electromagnetic simulations complete and the total losses are estimated, they are
transmitted to MotorCAD along with the WFSM geometric design. Because of the high power
density required, ATF spray of the end windings was selected for cooling. To properly model the
cooling system implementation of the rotor winding end turns, the thermal equivalent circuit
within MotorCAD was modified to have direct heat transfer from the rotor winding end turn
nodes to the ATF fluid.

8
TABLE I. PARTIAL LIST OF DIMENSIONAL AND NON-DIMENSIONAL GEOMETRIC PARAMETERS.
Symbol Definition

𝑚 Phases
𝑝 Pole pairs
𝑞 Slots per pole per phase
𝑧 Stator slots (2𝑚𝑝𝑞)
𝛼𝑝 Half pole angular span [Rad.]; 𝛼𝑝 = 𝜋/(4𝑝)
𝛼𝑠 Half stator slot angular span [Rad.]; 𝛼𝑠 =
𝜋/(4𝑚𝑝𝑞)
𝑅𝑆𝑜 Stator outer radius
𝑅𝑆𝑖 Stator inner radius
𝑘𝑆𝑖 Stator-rotor split ratio
𝑑𝑆𝑦 Stator yoke depth
𝑘𝑆𝑦 Stator yoke ratio
𝑤𝑆𝑡 Stator tooth thickness
𝑑𝑆𝑡 Stator slot depth
𝜃𝑆𝑡 Stator tooth tip angle
𝑑𝑆𝑜 Stator opening depth
𝑘𝑆𝑡 Stator tooth thickness ratio
𝑤𝑆𝑜 Stator slot opening
𝑘𝑆𝑜 Stator slot opening ratio
𝑅𝑆𝑏 Stator slot bottom radius
𝛿0 Minimum airgap
𝛿𝑀𝐴𝑋 Maximum airgap (at pole tip)
𝑅𝑅𝑜 Rotor outer radius (at minimum airgap)
𝑅𝑠ℎ Shaft radius
𝑅𝑅𝑡 Rotor radius at pole tip
𝑤𝑅𝑞 Rotor tip distance from q-axis
𝑘𝑅𝑞 Rotor tip distance ratio
ℎ𝑅𝑛 Rotor pole neck thickness

9
Fig. 4. Partial stator dimensional parameters.

Fig. 5. Partial rotor calculated output dimensions.

10
Fig. 6. Example rotor pole morphing

Fig. 7. Flow chart of rotor pole generation.

11
5) Capacitive Power Coupler Design

Capacitive power transfer (CPT) utilizes an AC electric field to push power across a gap as
shown in Fig. 8 [3]. To push current through the coupling capacitance safely, the system must
stay below the gap medium’s electric field breakdown strength. To keep the field below the
critical value (for air, 3kV/mm) the surface should be large in area, A, and the power transfer
frequency  should be very high (~MHz). Much like transformers, capacitive couplers have a
flux limitation dual, in Amps/Hertz. Throughput power is governed by the impedance
magnitude of the load, with higher impedance yielding higher power, as CPT system is
inherently current driven.

Fig. 8. Basic concepts of small gap capacitive power transfer

The capacitive power coupler (CPC) to be used with the WFSM is an axial flux hydrodynamic
configuration [4], since it has very large surface area per unit volume. Journal bearing style
couplers are also possible but are more suitable for lower speeds. Axially stacked dielectric
coated aluminum rotor and stator plates (100 mm diameter and 0.4064 mm thick) form the
coupling capacitors. A spiral groove in the stator plates forms a hydrodynamic thrust bearing
maintaining a 115 micron gap with air as the working fluid, shown in Fig. 9, resulting in 2.8 nF
of coupling capacitance per plate set. This mechanical design is a direct extension of [4,5]. A
class E amplifier and rotating rectifier have been designed for a switching frequency range of
500 kHz to 2.0 MHz with 1200 V SiC switches, shown in Fig. 10. The CPC was designed for a
peak 2 kW transfer with a mass of 600 grams. Detailed electronics analysis for component sizing
and circuit operation for the CPCs is documented in [6-8]. The CPC is 1/3 the axial length of the
traditional brushless exciter for this machine rating and potentially smaller and less expensive
than inductive transfer approaches [4]. Other CPT advantages include rotor current sensing on
the stator side that is insensitive to saturation and temperature dependencies [6]. In addition to
the CPC imposed constraint on the design of the WFSM of a maximum rotor field power of 2.0
kW, rotor geometric design variables were constrained to ensure the inner diameter of the rotor
winding end turns was greater than 135 mm, such that the CPC and ATF-compatible casing can
be placed immediately next to the rotor back-iron and minimize the machine axial length.

12
(a) (b) (c)
Fig. 9. Axial flux hydrodynamic bearing capacitive power coupler, (a) exploded model of coupler stack, (b)
cutaway solid model of the interleaved coupler plates, (c) stator coupler plate, (d) rotor coupler plate.

13
(a)

(b) (c)
Fig. 10. Field driver and rotating rectifier circuit schematic: Lp1 Lp2 provides constant current, Lr1 Lr2 are resonant
inductors, capacitors C1 and C2 form the CPC assembly, D1-D4 represent the rotating rectifier and Lf, Rf are
the field winding inductance and capacitance. (b) primary side class E amplifier, (c) secondary side rotating
rectifier.

6) WFSM Design Optimization

WFSM specifications for peak output power, continuous output power, specific power density,
and volumetric power density were developed in consultation with DOE USDRIVE targets: 55
kW peak power output, volumetric peak power density of 5.7 kW/l, and specific peak power
density of 1.6 kW/kg. A base speed of 4000 RPM was selected along with a target constant
power speed range of three to one (corresponding to 12,000 RPM). This base speed is a
reasonable hybrid between base speeds typically seen in automotive traction motors used in
hybrid vehicle drivetrains and those used in pure electric vehicle drivetrains. The three to one
constant power speed range (CPSR) also a reasonable CPSR for a main automotive traction
motor.

14
The stator outer diameter was constrained based on typical packaging constraints in an
automotive application and to fit the dynamometer available for experimental characterization.
An additional minimum stator inner diameter constraint was imposed to allow the CPT to nest
inside the stator end turns if desired. The CPT also imposed several constraints on the rotor field
winding including a maximum rotor field winding power transfer of < 2.0 kW, field winding
terminal current of 7 A, and minimum field winding terminal voltage of 200 V. After extensive
test runs of the optimizer to identify the most feasible initial topology a 48 slot 8 pole single
layer winding design was selected compatible with automotive mass production. Using the
differential evolution optimization method with extensive test runs showed that hard constraints
act as additional objectives as long as a suitable population is found, and in general, it has been
observed that adding hard constraints speed up convergence more than increasing the number of
optimization objectives.
The final optimization runs for the prototype 1 and 2 geometries had 14 input parameters
including dimensional geometric, non-dimensional geometric and physical variables. Two
optimization objectives were used to maximize the torque density and "goodness" (average
electromagnetic torque divided by the square root of the losses), Table II. Additionally 7 hard
constraints were imposed: torque ripple <5%, minimum average electromagnetic torque > 140
Nm, maximum average electric torque < 150 Nm, maximum rotor ohmic losses <2.5 kW,
maximum stator total losses < 6 kW, stator current density 15 - 30 A/mm^2, and rotor current
density of 15 - 25 A/mm^2, Table II. A graphical projection of the results on an inverse torque
density versus "goodness" plane of the full population represented by black dots (7,500 total
members) and only the designs that meet ripple and rotor loss constraints are shown in Fig. 11(a)
and Fig. 11(b) respectively. The yellow dot represents the design selected for prototyping and
experimental characterization.
TABLE II. FINAL OPTIMIZATION OBJECTIVES AND CONSTRAINTS
Objectives Goal
Torque Density (Average Maximize
Electromagnetic Torque/Volume)
”Goodness” (Average Maximize
Electromagnetic Torque/ Plosses)
Hard Constraints Value Units
Torque Ripple < 5% Per Unit
Average Electromagnetic Torque > 140 Nm
(Minimum) [55 kW target output]
Average Electromagnetic Torque < 150 Nm
(Maximum)
Rotor Ohmic Losses (Maximum) < 2500 W
Stator Total Losses (Maximum) < 6000 W
Stator Current Density 15 – 30 A/mm2
Rotor Current Density 15 – 25 A/mm2

15
(a) (b)

Fig. 11. Optimization results for the design of prototypes 1 and 2 shown for the (a) total 7,500 member population
and (b) designs which meet torque ripple and rotor loss constraints. The yellow dot indicates the design
selected for prototypes 1 and 2.

A. Design of Experiments for Rotor Structural Limits


A central composite design of experiments technique was used to predict the maximum range of
five rotor dimensional variables at the top hierarchical level (stator outer radius, rotor outer
radius, minimum airgap distance, rotor neck thickness and rotor pole thickness), shown in Fig. 5,
to avoid mechanical failures. The rotor laminations were initially modeled as an anisotropic
combination of Voigt and Reuss composites in the radial and axial directions respectively [9].
When compared to a solid isotropic block of silicon steel (M250-35A), this assumption lowers
the Young’s Modulus of the rotor by 8-12%, corresponding to an 8-12% increase in structural
deflection under rotational load. This model better captures the elastic behavior of the laminated
structure with respect to a solid silicon-steel model, but does assume a bonded lamination
structure. All other material properties were assumed to be that of plain M250-35A silicon steel.
The rotor field winding was modeled as a single entity with a density equal to 50% of solid
copper to account for the fill factor to determine the rotor core load, and separately a single wire
stress model has been evaluated for the winding design. Stress analysis was carried out in
SolidWorks. The Von Mises stress, safety factors, and displacement were computed at 12,000
RPM (targeted maximum rotor speed) and 15,000 RPM. Representative Von Mises stress and
equivalent strain results are shown in Fig. 12(a) and Fig. 12(b).
In addition to the determination of maximum ranges for the five rotor dimensional variables, the
response surface calculated as a least squares linear regression was added as rotor stress and
safety factor estimates to the geometry generation routine. The response surface sets the
constraints that allow any final optimization results to meet a stress safety factor of 2.0 at the
maximum design speed.

16
(a) (b)

Fig. 12. Rotor structural limitation analysis, (a) example von Mises stress, (b) example equivalent strain.

B. WFSM Optimization Constraints and Objectives


A parallel stochastic differential evolution optimization algorithm was used for the multi-
objective optimization. For each generation, a group of designs are evaluated in parallel as a
crossover of the parameters of previous generation best members, and at the end of each
generation the best members are selected by the algorithm. The implementation allows the user
to set two different types of objectives and constraints on the optimization run: the objectives
maximize the resulting output parameter and select the best members in the population, while the
hard constraints directly eliminate the member from the population when not met. The effect on
the overall simulation is that the hard constraints are met early on, but once met have little effect
on the following generations, while the objectives force the Pareto front to move further.
In order to obtain high torque density without sacrificing the motor efficiency, the two objectives
selected for maximization are torque density and “Goodness” (ratio of the torque and the square
root of machine total losses), that gives a better compromise with respect to the simple loss
minimization. Some additional hard constraints were added to enforce secondary targets: a
maximum torque ripple of 5% for realistic operation as an automotive traction motor, an average
torque range between 140 and 150 Nm to target the design peak power of 55 kW while avoiding
oversized motors, maximum rotor field losses to be able to operate the CPC in the whole
predicted temperature range of the windings (up to 190 C), and maximum stator total losses
ensure proper operation given the ATF spray cooling system’s estimated capability. The
objectives and hard constraints are listed in Table II.
The 14 input parameters to the optimization are divided into three groups: dimensional
geometric, non-dimensional geometric, and physical variables. The 3 dimensional geometric
variables are the minimum airgap distance (0.5 to 1 mm), the stator outer radius (80 to 220 mm)
and the machine axial length (50 to 200 mm). For the non-dimensional parameters, 4 apply to the
stator: the split ratio between stator inner and outer radius, kSi, tooth thickness ratio, kSt, yoke
thickness ratio, ksy, and slot opening ratio, kSo, and the remaining 5 to the rotor: neck thickness
ratio, kRn, pole shoe thickness, 𝑙𝑡ℎ , distance from the interpolar axis ratio, 𝑑𝑅𝑞 , rotor yoke
thickness ratio, 𝑘𝑅𝑦 , and winding shape ratio, 𝑘𝑅𝑤 . Physical variables are the stator and rotor net
current density, assuming reasonable slot fill factors of 0.4 in the stator and 0.5 in the rotor
windings.

17
C. Optimization Results and Down Selection
The final optimization results are shown in Fig. 11. Each WFSM design (7,500 total population
members) is shown as a black dot. WFSM designs meeting the torque ripple constraint of less
than 5% are indicated by a blue dot. The Pareto front of best tradeoff designs are colored red and
the final design down-selected for prototyping and experimental characterization is the yellow
dot. Enforcing the CPC imposed rotor copper loss limit of 2.0 kW in the previous distribution,
the population of Fig. 11(b) is obtained. The selected design for prototyping is the one with the
largest predicted torque density on the Pareto front of the resulting optimization designs.
D. Final Structural Design and Rotor Stress Reduction
To improve the structural material modeling, test coupons cut from an early revision M250-35A
lamination stack were subjected to ultrasonic material property measurements [9]. The Young’s
modulus was measured in both the axial and radial directions (through the lamination stack and
across the lamination stack). Lack of shear ultrasonic transducers precluded the evaluation of the
Poisson’s ratios of the laminations. Comparison of the measured material properties with those
of earlier material models showed a 7.3% decrease in the radial Young’s modulus.
Structural FEA was carried out on the down selected lamination model using the updated
material model obtained through ultrasonic measurement. Fillets added to the most stressed
sections of the rotor produced a minimum factor of safety of 1.5 at 15,000 RPM (12,000 RPM is
maximum operational design speed). A maximum rotor deflection of 0.0478 mm was predicted,
closing the air gap by 4.78%. Additional magnetic FEA found that the addition of stress relief
filets did not adversely affect the electromagnetic power conversion performance of the motor.
The selected lamination design for prototyping with rotor stress relief fillets is shown in Fig. 13
and the machine’s parameters and predicted performance in Table III.

Fig. 13. Final lamination design selected for prototyping.

18
TABLE III. PROTOTYPE MACHINE PREDICTED PARAMETERS

Machine Parameters Simulated Value


Stator Outer Diameter / Airgap Length 254 mm / 0.95 mm
Rotor Outer Diameter / Stack Length 176 mm / 91 mm
Base Speed / Base Frequency 4000 RPM/ 267 Hz
Maximum Speed / CPSR 12 kRPM / 3:1
Peak Power / Peak Torque at Base Speed 59 kW / 141 Nm
Volume (Gross* / Net**) 11.06 / 4.66 ℓ
Weight (Gross* / Net**) 40.64 / 32.65 kg
Torque Density (Gross* / Net**) 12.79 / 30.29 Nmℓ-1
Vol. Power Density (Gross* / Net**) 5.36 / 12.69 kW ℓ-1
Specific Power Density (Gross* / Net**) 1.45 / 1.81 kW kg-1
Total Losses / Efficiency at Peak Power 4.13 kW / 93.5 %
Rated Stator / Rotor Current Density 18 / 17 A mm-2
*Gross refers to the full system (including shaft and spray cooling rings),
**Net refers to the active materials of the electric machine (CPC, core and windings, including the end windings).

7) Simulated WFSM Prototype Performance

To predict the WFSM performance and design the stator and rotor windings, a mapping of the
prototype’s predicted torque, losses, and single turn voltage has been carried out using the same
software developed for the design optimization. The parameters swept in the mappings are the
stator and rotor current density, torque angle, and rotational speed. However, the results are
presented in the terminal current scale, to allow for direct comparison to the experimental results.
Both predicted and experimental results are presented in this section to compare the predicted
values with those measured during dynamometer testing. Additional details about the
dynamometer mapping are included in following sections.
E. Torque and Efficiency Maps
The predicted shaft torque mapping is plotted alongside the experimental results (red dots) at the
base speed of 4,000 RPM for the current angle for maximum torque, Fig. 14(a). Also, the current
angle for maximum torque has been plotted in Fig. 14(b) and Fig. 14(c) where a positive current
angle is measured counter clockwise from the q-axis. Both the predicted and experimentally
measured results are in good accordance. At high stator current magnitudes, above 200 Apeak, the
predicted and measured torque begins to deviate slightly. This is most likely due to the lack of
detailed information of the BH curve of the stator core material for very high saturation. It
should be noted that the experimental torque data for the region with rotor current larger than 4
A and stator current larger than 300 A have been measured from a test of 30 seconds or longer.
The predicted efficiency maps of Fig. 15(a) and 15(b) are two possible representations of the
data. Fig. 15(a) plots the maximum predicted efficiency for a given torque level at a given speed.
The high speed operation is limited by the terminal voltage. Fig 15(b) is the efficiency map for
the maximum torque at the speed of 4000 RPM, referred to the torque data of Fig. 14(a). The
predicted values therefore refer to different assumptions: maximum efficiency operation on a
speed range for Fig. 15(a) and efficiency at maximum torque operation at a given speed (4000
RPM) for Fig. 15(b).

19
(a)

(b)

Fig. 14. Predicted and experimental results, (a) shaft torque (surface simulated, red dots experimental) for the
maximum torque current angle, (b) current angle of the maximum torque for a given field and stator current
(left simulated, right experimental) at 4000 RPM.

20
(a) (b)

Fig. 15. Predicted prototypes 1 and 2 efficiency with a winding temperature of 70 C versus (a) speed and torque
for peak efficiency and (b) as a function of field and stator currents at the maximum torque angle at 4000
RPM.

F. Voltage Maps and Winding Design


The results of the voltage mapping are displayed in Fig. 16(a) for the base speed overlaid with
the experimental data, which is in good accordance for field currents larger than 2 A (less than
5% discrepancy), with the highest difference of 15% for 200 A stator current and 2 A rotor
current. The stator and rotor windings were optimized using the data provided in the torque and
voltage maps as a function of stator current angle, stator current density and rotor current density
at 4,000 RPM (base speed) and 12,000 RPM (maximum design speed).
The stator winding design is a compromise between a maximum voltage at 3:1 constant power
speed ratio of 350 V (limited by the inverter peak phase voltage) and a maximum stator terminal
current of 300 A at the base speed due to the drive limitation. The resulting stator winding is 4
turns per coil, 10 AWG 16 strands in hand per turn, 2 slot per pole per phase single layer
distributed winding. The wire gauge and strands in hand were selected to maximize the
achievable slot fill when using hand insertion.
The CPC resonant class E inverter required the field resistance to be in the range of 20 to 45 Ω at
the nominal voltage of 300 V for the entire possible operating temperature range of the WFSM,
20 to 190 °C. In addition, structural modeling had shown that a wire size of at least AWG 24
must be used to guarantee structural integrity at high speed and temperature. Given these
constraints the rotor winding with 239 turns per pole of AWG 21.5, for which the rotor resistance
remains in the range of 26 to 45 Ω over the operating temperature range, was designed.
With the winding defined, the predicted machine equivalent circuit parameters were calculated
(Table IV has the predicted and measured values at 20 °C). Since the inductances are affected by
the machine saturation, Fig. 17 shows the predicted values and the some experimental results for
field and mutual inductances, and Fig. 18, and d-axis and q-axis stator inductances.

21
(a) (b)

Fig. 16. (a) Voltage map at 4000 RPM, 20 current angle (green points are experimental results), (b) voltage map at
12,000 RPM, 70 current angle. All voltages are peak values per turn per pole including all harmonic
content.

TABLE IV. PREDICTED AND EXPERIMENTALLY MEASURED WFSM PROTOTYPE EQUIVALENT CIRCUIT PARAMETERS.

Parameters* FEMM MagNet Experimental


Lf [mH] 2330 2386 2278
Ld [mH] 0.969 0.998 0.98
Lq [mH] 0.565 0.597 0.60
Lm [mH] 32.36 32.94 33.46
Rf [Ω] 26.952 26.952 26.5
Rs [Ω] 0.0153 0.0153 0.02
* Unsaturated inductance values and resistances at 20 °C

Fig. 17. Predicted and experimentally measured field, Lf, and mutual, Lm, inductances.

22
Fig. 18. Predicted and experimentally measured stator inductances, Ld, and Lq.

8) WFSM Prototype Construction

Two wound field synchronous machine prototype designs have been constructed. The first
prototype was found to have a distorted stator inner diameter resulting in a stationary saliency.
The unintended saliency made regulation of the stator currents very difficult. The second
prototype stator and rotor lamination stacks were laser cut, annealed, and unitized using laser
welding from M15 29 Ga (stator) and M250-35A steel (rotor), Fig. 19.
Winding and insulation materials were selected for ATF compatibility and class H, or higher,
thermal performance. Twenty type T thermocouples were embedded in the stator for thermal
characterization. To retain the rotor windings at 12,000 RPM with potentially high rotor winding
temperatures and avoid ATF degradation, rotor endcaps were machined from polyether ether
ketone (PEEK). To further ensure rotor winding retention, Kevlar fiber banding was used as an
overlay on a rotor endcap shelf, though this is not necessary according to structural finite element
analysis.
The shaft of the motor was machined from AISI 1144 steel, chosen for it high machinability and
increased strength over the more common AISI 1045 steel. The shaft uses a hub system allowing
for rotors to be exchanged and also passes the field winding leads internally through the shaft
past the bearings to slip rings and CPC terminals. The shaft is oversized compared to a typical
automotive environment, allowing for iteration of the electromagnetic components without loss
of structural integrity. The combined rotor and shaft of the prototype was dynamically balanced
(ISO 1940 /1 G1.0 at 9,600 RPM) to suppress mechanical vibration modes in the operational
speed range of the motor. The geometry of the rotor (length: diameter > 0.5) is such that a dual
plane balance was required.

23
The stator is clamped in a tubular aluminum housing and aligned via an array of radially acting
clamp bolts; a design which facilitates quick exchange of the stator, Fig. 20. Brass shims on the
surface of the stator distribute the clamping force, thereby preventing accidental delamination of
the stator. For the current prototype, the brushes and CPC are located external to the motor
housing to allow for easy access during testing. The CPC housing is designed, however, to allow
it to operate in an ATF spray environment inside the machine shell. ATF spray cooling of the
stator and rotor end turns is implemented using a relatively simple system, Figs. 21-22. Coiled
copper piping with one closed end is connected to the pump system. The spray holes consist of
two sets of 16 holes (1.15 mm diameter) offset by 30 degrees with one set cooling the rotor end
winding and another cooling the stator end windings.

Fig. 19. Wound field synchronous motor prototype.

Fig. 20. Stator housing with clamping system and non-drive end plate.

24
ATF Inlet

Fig. 21. Drive end ATF spray cooling system.

CPC
Slip Rings

Fig. 22. Non-drive end ATF spray cooling system and field winding terminal wire passthrough.

25
Fig. 23. Spray ATF cooling ring test.

26
9) CPC Prototype Construction

Several sets of hydroflex coupling capacitors were constructed in house at UW-Madison. Here,
the stator and rotor plates were water jet cut from 0.016” thick 6061 aluminum stock. These
plates are pictured in Fig. 24. The rotor plates were anodized to ensure galvanic isolation at zero
speed. These plates were then alternately stacked, 5 stators and 4 rotors, 0.003” apart, onto a
dynamometer test stand to verify the capacitance over the intended speed range of 0 – 10k RPM.
Tests verified that the capacitance value was sufficient to power the rotor (capacitance was > 1.5
nF) over all conditions.

(a) (b)

(c)
Fig. 24. Photos of CPC hydroflex (a) stator plate, (b) rotor plate, (c) small dyne stand to test fully
assembled CPCs to up to 10k RPM.

27
After evaluation on the small benchtop dynamometer, the CPC hydroflex stack could be installed
on the shaft of the WFSM for test on the traction dynamometer. Two stacks are installed on the
WFSM along with a rectifier board to realize a complete CPC as shown in Fig. 25. The CPC
was mounted external to the case and connects to the rotor via a hollow shaft. This was done for
convenience in testing, especially troubleshooting.

Fig. 25. CPC prototype number 2 mounted on WFSM for testing.

10) Dynamometer Testing of Prototypes

The WFSM prototype 2 performance was measured with both slip rings/brushes and CPC on
a 180 kW Anderson Electric Controls dynamometer with a 1000 Nm HBM T40 torque flange,
Fig. 26(a). The prototype WFSM stator was driven by a Semikron Semikube IGBT Module
Stack IGD-20424-P1N6-DH-FA inverter, Fig. 26(b) and Fig. 27(a). For testing with slip
rings/brushes, the field winding was connected to a Magna-Power PQ500-20 DC power supply
through an added snubber circuit and diode to protect the power supply in the case of excessive
field transients, Fig. 27(a). The ATF spray cooling pump, reservoir, and filter were also
contained in the same blue cabinet, Fig. 27(b). A custom DSP board, based around the Spectrum
Digital TI F28335 DSP, was used to control the inverter, regulate currents in the WFSM, and
perform field oriented control, Fig. 28. A block diagram of the overall control system for
WFSM dynamometer testing is shown in Fig. 29. An interface to the DSP control code was
created in MATLAB with signals transmitted over a serial link to avoid corruption by EMI, Fig.
30. To regulate the stator currents, a discrete-time complex vector current regulator has been
implemented to decouple the stator d-q axis cross-coupling, Fig. 31. The WFSM and drive
efficiency were measured using Yokogawa WT1800 and PX8000 power analyzers when tested
with slip rings/brushes, Fig. 26(b). Stator phase currents were measured using Yokogawa 96031
current probes and a LEM Ultrastab system connected to the WT1800 and PX8000, respectively.
Field current and voltage were measured both by the power analyzers and captured on a LeCroy
HDO6034-MS. WFSM efficiency was measured during testing with the field excited by the CPC
using a WT1800 due to the unavailability of the Yokogawa PX8000. CPC converter DC input

28
voltage and current were measured with the power analyzer, and the CPC primary side switching
current and voltage waveforms were captured on the LeCroy HDO6034-MS. To record the
various stator, case exterior, and ambient temperatures a 20 channel Agilent thermocouple reader
was used.

(a) (b)

Fig. 26. Dynamometer testing setup, (a) 180 kW Anderson Electric Controls dynamometer with prototype WFSM
mounted to the left with slip rings, (b) drive cabinet housing stator inverter, DC power supply, oil cooling
setup, and measurement equipment including Yokogawa WT1800 and PX8000 power analyzers, and
LeCroy HDO6034-MS oscilloscope.

(a) (b)
Fig. 27. Drive cabinet contents, (a) Semikron Semikube IGBT Module Stack IGD-20424-P1N6-DH-FA inverter
(left) and Magna-Power PQ500-20 DC power supply (right), (b) ATF spray cooling pump, reservoir, and
filter.

29
Fig. 28. Custom Texas Instruments F28335 DSP signal conditioning board used for WFSM controls development.

Dyne. PIU Semikron Drive WFSM Dyne


A
or U UDC + Motor U
iA A B V
Stator Thermal Torque Dyne
V iB B Rotor Field
UDC- Winding Couples Meter Incremental Drive W
W iC C Winding C
if Uf Encoder
DC-Link
TempUDC iA iB iC Uf if
Agilent Analog External
Measurement Module Sensor CPC Magna-
PC -
& PC With Power
Analog(±10V) Data
F 28335 DC
RS485 Recording
HB1± HB2± HB3± Semikron Drive Translation Board Power Isolated
PWM Invertering Scaling Scaling Rev B is compatible with internal Scaling Source
Buffer
Voltage Translation Card1 Card 2 phase current sensors or external Card3 brushless brushed
Semikron Drive Translation Board current sensors ABZ
RS Handbox
Analog (0V-3V)
485
1-2 3-4 5-6 1_P 2_P 1_P 2_P 3_P 55 63 1_P 2_P CAN-A 3_P 4_P eQEP1
PWM1-6 ADC_ Card_1 ADC_ Card_2 GPIO ADC_ Card_5 ADC_ Card_5
Machine Control Panel
Halt Signal
based on Matlab GUI
ADCINA0-1 ADCINA 3-5 ADCINA6-7 ADCINB6-7
SCI-A
F 28335 I/O Board A PC

Fig. 29. Overall WFSM control system structure for dynamometer testing.

Fig. 30. MATLAB Interface to DSP controller over a serial link.

30
Discrete sampling
delay compensation
+ *
- 1 vdqs (1  e  ( R s / L s ) T s )
i *dqs K e(Rs / Ls )Ts e jeTs Latch
1  z 1 R s ( z  e  ( R s / Ls )Ts )
+ -
i dqs jeTs Discrete
e
machine model
Complex vector
K (e jeTs  z 1e( Rs / Ls )Ts ) controller in direct Back-EMF
(1  z 1 ) discrete design decoupling look-
up table
e i f

Fig. 31. Direct discrete time complex vector current regulator.

G. Dynamometer Testing of WFSM with Brushes


The WFSM no-load bearing and windage losses were measured (~120 W at 4,000 RPM) with
and without the ATF spray cooling system engaged. No significant difference was observed. The
open circuit core losses and open circuit line to line voltage as a function of the rotor field
current were measured and compared with simulation predictions, Fig. 32(a) and 32(b), showing
very close agreement.

(a) (b)

Fig. 32. Open circuit, (a) core losses and (b) line to line voltage as a function of field excitation current and
rotational speed.

WFSM prototype dynamometer torque, efficiency, and power factor mapping was also
carried out for a range of field currents, stator current magnitudes, and current angles (angle, ,
with respect to the q-axis with CCW taken as positive) at 1000, 2000, 3000, and 4000 RPM. The
mapping was carried out within a stator temperature range of ~35 to ~100 C, Fig. 33. A nominal
temperature of ~50 C was targeted for measurements. The torque, current angle for maximum
torque, machine efficiency (at 4000 RPM and maximum torque current angle), and power factor
(at 4000 RPM and maximum torque current angle) mapping results for operation with slip rings
are shown in Fig. 14, Fig. 34, and Fig. 35. The predicted and measured torque maps agree
closely. The higher experimental versus predicted torques at high field and stator currents is
most likely due to lack of detailed BH curve data at high saturation levels. The measured
efficiency map shows a very broad high efficiency band with >95% efficiency. The measured
efficiency map corresponds well with the predicted efficiency map in Fig. 15. Both power

31
analyzers were used to estimate the WFSM efficiency with the calibrated current sensors and
Yokogawa PX8000 reporting approximately 1 % higher efficiency than the Yokogawa WT1800
with essentially identical contours. The power factor map at 4000 RPM, Fig. 35, shows the
power factor at the current angle for maximum torque which is not necessarily the current angle
for maximum power factor. The power factor is quite high over the lower half of the map. Small
reductions in the torque capability allow for increased or unity power factor (reduction of
reluctance torque). A screen capture of the Yokogawa WT1800 power analyzer at the peak
torque operating point is shown in Fig. 36 demonstrating operation at ~ 80 kW output at the base
speed of 4000 RPM.

Fig. 33. Thermal data of peak power testing, detail of stator currents of 200 A (test points 500 to 700), 100 A (test
points 700 to 750) and 50 A (test points 775 to 850), rotor current from 7 to 3 A with intervals of 1 A, 4000
RPM speed. The trace colors correspond to the following temperature measurements: red = buried in coil
A, purple = coil A slot side, yellow = coil A slot bottom, green = coil A end turn, blue = ambient.

(a) (b)

Fig. 34. Measured WFSM efficiency results at 4000 RPM using (a) Yokogawa PX8000 and (b) Yokogawa WT
1800. The temperature of the stator windings ranged between 50 to 100 C. Efficiencies are for maximum
torque current angles.

32
Fig. 35. Measured WFSM prototype 2 power factor at the maximum torque current angle at 4000 RPM.

Fig. 36. Yokogawa WT1800 power analyzer screen Capture showing ~80 kW WFSM output with slip rings and
brushes. Elements 1, 2, 3 - machine phases, element 4 – field power, element 5 – inverter DC link

33
H. Dynamometer Testing of Wound Field Synchronous Machine Prototype with Capacitive
Power Coupler
The torque output of the WFSM operated with the CPC was equivalent to operation with brushes
and slip rings at the same stator current magnitude and rotor field current. Operation with the
CPC was tested up to the target power output of 55 kW. A screen capture from the Yokogawa
WT 1800 power analyzer in Fig. 37 provides terminal data for the 55kW test. This demonstrates
that a CPC is possible in a traction application. Output power levels vs stator current and CPC
tank current are in Fig. 38.

Fig. 37. Power Analyzer Screen Capture showing 55kW WFSM output with CPC excitation at 4000 RPM.
Elements 1, 2, 3 - machine phases, element 4 – field power, element 5 – inverter DC link.

Output Power vs. CPC & Stator Currents at 4000RPM


60

50

40
Output Power [kW]

30
Is = 300 A
Is = 250 A
20 Is = 200 A
Is = 150 A
10 Is = 100 A

0
0 1 2 3 4 5 6 7
CPC Tank Current, [A rms]

Fig. 38. WFSM output power at 4000 RPM as a function of CPC tank current and stator current magnitude.

34
Relevant CPC voltages and current were measured during the 55kW load test. Measured
voltages and current are plotted in Fig. 39 and correspond to quantities labeled in the circuit
diagram in Fig. 10(a). From these waveforms, it is seen that the converter operates at 2MHz and
soft switches (zero voltage switching) for low losses at high frequency. The CPC tank current is
6.5 Arms, which corresponds to a DC field current of about 5 amps when losses and parasitics
are accounted for.

Fig. 39. Measured CPC circuit waveforms during 55kW WFSM test. 2MHz switching and 6.5 A rms CPC tank
current. *note: labels correspond to circuit diagram in Fig. 10(a).

11) Comparison of WFSM with CPC Performance with DOE Targets

The WFSM performance metrics using slip rings and brushes are listed in Table V and
compared to USDRIVE 2020 and predicted design values. The measured performance
significantly exceeds USDRIVE 2020 targets without the use of permanent magnets. The peak
power output can also be held for more than 30 seconds, essentially continuously as the stator
current magnitude is now limited by the Semikron drive. The higher measured than predicted
performance is because of the thermal properties of the machine. The predicted performance
limited the stator and rotor current densities to 18 and 17 A/mm^2 respectively. Due to the
effectiveness of the spray cooling system, significantly higher values of current density and
terminal currents were achievable resulting in very favorable peak volumetric and specific torque
and power densities that compare favorably with commercial IPMSMs and IMs (17.22 Nmℓ-1,
4.69 Nm kg-1, 7.19 kW ℓ-1, 1.95 kW kg-1 ). The specific power density metrics would be further
improved by the use of a shaft more appropriately sized for the application. Compared to the
USDRIVE target of 55 kW peak power output, a peak measured power output of 79.6 kW was
achieved with slip rings and brushes. Operation at 55 kW with the CPC was verified. The
performance of the WFSM with the CPC met the project go/no go and milestone requirements.

35
TABLE V: COMPARISON OF WFSM PERFORMANCE WITH USDRIVE 2020 TARGETS.

USDRIVE Prototype Prototype 2 Stator/


Machine Parameters 2020 Target Prototype 2 Rotor
Motor Type Any WFSM WFSM

Base Speed / Max Speed [RPM] N/A 4,000/12,000 4000/TBD

Mass [kg] N/A < 42.30 40.64**

Volume [ℓ] N/A < 9.6 11.06***

Peak Motoring Power [kW] 55 59 79.6*


Continuous Motoring Power [kW] 30 30 Held peak continuously

Peak Torque [Nm] N/A 141 190.55*


Specific Power Density [kW/kg]* 1.6 1.4** 1.95**

Volumetric Power Density [kW/ℓ]** 5.7 5.4*** 7.19***

Specific Peak Torque Density [Nm/kg] N/A 3.46** 4.69**

Volumetric Peak Torque Density [Nm/l] N/A 12.79*** 17.22***


* Continuous, held for over 30 seconds during testing.
** Includes stator, rotor, and shaft mass (shaft is oversized)
*** Cylindrical volume of active materials plus ATF spray cooling rings
**** Cylindrical volume of active materials

12) Integration of WFSM with CPC Excitation into a Microgrid

In certain scenarios, it is conceivable that a vehicle with a WFSM hybrid traction drive could be
used for mobile power take off. Controller designs for controlling real and reactive power export
from a WFSM powered from an internal combustion (I.C.) engine prime mover have been
carried out for isolated load and grid connected operation, Fig. 40. Using a simplified model of
the I.C. engine consisting of inertia and a delay term between fuel input and torque output a
speed controller with predictor compensator has been developed to cancel the time delay, Fig.
41. An additional torque disturbance input decoupling (DID) as also been investigated. An
output phase voltage regulator has also been designed using a simple PI controller for use with a
brushless exciter, Fig. 42. In grid connected scenarios the output voltage magnitude and phase of
the WFSM generator must be matched to the grid before closing contactors at the point of
common coupling. A three phase PLL has been designed to force the convergence to the correct
phase, Fig. 43, and integrates with a droop controller.

36
Fig. 40. Proposed power take-off control scheme for using the traction WFSM as a generator when used in a hybrid
electic vehicle where the IC engine can serve as a prime mover.

(a)

(b)
Fig. 41. Speed controller with predictor compensator for the I.C. engine fuel input to torque output delay. (a)
proposed predictor controller, and (b) equivalent block diagram with the delay term eliminated.

37
(a)

(b)
Fig. 42. Output phase voltage regulator controller; (a) WFSM dynamic model with brushless exciter, and (b) PI
output phase voltage regulator.

(a)

(b)
Fig. 43. Three phase, phase locked loop (PLL) for synchronization to grid; (a) three phase, PLL and (b) I.C. engine
speed reference generator and droop controller.

A block diagram of the control scheme for stator voltage regulation (at the heart of reactive
power flow control) with a CPC is shown below. Controller designs for controlling real and
reactive power export from a WFSM powered from an internal combustion (I.C.) engine prime
mover have been carried out for isolated load and grid connected operation. The block diagram
below documents the synchronous generator control scheme for stator voltage regulation using a
CPC as the field driver. The details of the controller are covered extensively in [6].

38
1 I qs qs
1
Rs  RL
Lqs s
Controller Physical system
e
Vs
I I
2
ds
2
qs
RL
e
df I ds ds
1 1
Rs  RL
Lds s

 fd
1.5 Lm

KI
s
Vs* V f* Vf 1 f 1 If df Vs
KP Field Driver Lm Machine
s Lf
1.5 Lm I ds
Rf  fd
e Lm Rf

 vin 
 f switch  Class E  I Cs1_1  2 Vres  1  I res  2 ires  1 v f 
V f* Lookup Inverter jCs1 j Lr 2
I res
  Cout
Table  Dswitch 
i f 

Vout _1  I res 4 v f 
I res  i f 

Field Controller Class E Resonant Inverter (Complex vectors) Diode Rectifier Output Filter

Fig. 44. CPC and WFSM excitation and phase voltage control pseudo block diagram

The “Go/No-Go 2: Demonstration of – “Document performance experimentally in CERTS microgrid”


decision point requires a demonstration of WFSM stator voltage regulation while paired with
CPC excitation. This was implemented on a 10 kVA generator set, rather than the traction drive,
to allow the simultaneous pursuit of milestones and mitigate technical risk. Using the journal
bearing style CPC depicted in Figures 45 and 46 on a generator set, and the controller previously
documented (diagram above), we demonstrated voltage regulation during an RL loadstep (0.55
pf). A capacitive power coupler (CPC) was formed from an off-the-shelf rotary style journal
bearing and provided 1.7nF of coupling. Regulation is plotted in Fig. 47 and the same Class E
converter circuit provides high frequency current. Voltage regulation of a WFSG is demonstrated
during steady state and 1 per unit load steps, yielding a NEMA-MG1 class G2 rating with
resistive load. It is seen that the CPC raises the duty ratio and lowers frequency to support the
increase in load. After the AC load on the generator is removed, the controller automatically
returns to its no load condition prior to the transient. See published work [6] for more detail.

39
C1
C2+
lC

C1+
C2
(a) Stator (left) & rotor (right) (b) Assembled

C1+ C1
From To
Inverter Rectifier

C 2 C 2+
(c) Ideal circuit model

Fig. 45. Journal bearing style CPC

(a) Journal bearing stator and rotor (b) Bearing rotor with rectifier (c) Bearing stator with assembly

(d) Bearing rotor mounted on WFSG (e) Assembled journal bearing setup (f) Push-pull-Class-E inverter prototype
rotor
Fig. 46. Photographs of experimental setup to test the CPC excitation system on a WFSG

40
Fig. 47. RL load step on CPC excited generator.

VOLTAGE REGULATION RATING DEMONSTRATED IN A 10 KVA GENERATOR SET APPLICATION


Steady-State Transient Voltage
Transient
Voltage Voltage Rise Recovery Time
Voltage Dip [%]
Regulation [%] [%] [s]
Test results

R loads 1.7 20 12.5 0.86

R+L
1.5 35 21 1.2
loads

G1 ≤5 ≤ 30 ≤ 35 ≤ 2.5
NEMA
rating

G2 ≤ 2.5 ≤ 24 ≤ 25 ≤ 1.5

G3 ≤1 ≤ 18 ≤ 20 ≤ 1.5

13) WFSM Prototype 3 Rotor Design

Because time and budget conditions allowed, a third prototype rotor was designed to gauge the
impact of rotor flux barriers on WFSM performance. New rotor geometric templates were
developed which included the addition of flux barriers in the rotor pole neck and the rotor pole
shoe. Additionally a new rotor pole shaping algorithm was developed. Extensive optimization
studies were carried out using the new rotor geometric templates and pole shaping algorithms for
a range of electrical steel grades. To design a third rotor prototype with reduced field power
requirements and increased torque density, a new optimization study was completed using the
new rotor pole surface and wider geometric parameter ranges. The design that was selected from
the Pareto front served as a base or seed design for the inclusion of flux barriers in a second
41
Monte Carlo study where only flux barrier geometric and current density parameters were varied.
This was termed an extended optimization. The new base speed design without rotor flux
barriers, is the yellow dot in Fig. 48. The cross-sections of the base design for the extended
optimization (yellow dot), and the design selected for prototype rotor 3 with a moderate size flux
barrier are shown in Fig. 49 with the full load flux density shaded and magnetic field lines
shown. The final lamination design for the prototype rotor 3 is shown on the right side of the
same figure. The full load torque produced by this design is substantially higher than the current
prototype 2 with substantially lower required field power by reducing the air gap from 0.9mm to
0.75mm and by reducing the back iron on the rotor to allow more copper in the slots, and the
addition of the flux barrier. The rotor prototype 3, Fig. 50 and 51 will be tested during a follow-
on DOE project.

(a) (b)

Fig. 48. Design optimization for selection of base design for inclusion of flux barriers; (a) full optimization
population with design selected as base design highlighted in yellow and (b) designs filtered which meet
all hard constraints.

(a) (b) (c)

Fig. 49. (a) Base rotor design for extended optimization with flux barriers, rotor design with center flux barrier
selected for prototype 3 rotor after adding stress relief filets (center), and lamination design (right).

42
Fig. 50. WFSM prototype 3 rotor stack with PEEK endcaps and rotor hub

(a) (b)
Fig. 51. Wound, varnished, and balanced prototype rotor 3 with flux barriers.

43
14) Project Task Schedule, Milestones, and Go-No-Go Points

The Gantt Chart below presents the project timeline and critical tasks:

Q4 14 Q1 15 Q2 15 Q3 15 Q4 15 Q1 16 Q2 16 Q3 16
ID Task Name Start Finish Duration % Complete
Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct

1 Budget Period 1: Design, Construction, and Initial Testing of WFSM 10/1/2014 9/30/2015 261d 100%

2 Task 1.1: Design of Wound Field Synchronous Machine EV Traction Motor 10/1/2014 1/30/2015 88d 100%

3 1.1.1 Finalize WFSM motor target specification document 10/1/2014 10/31/2014 23d 100%

4 1.1.2: Determination of thermal and structural limitations and the feasible design space 10/15/2014 1/30/2015 78d 100%

5 Task 1.2: Development of WFSM – Prototype 1 11/3/2014 4/30/2015 129d 100%

6 1.2.1: [M1.2] Development of WFSM multi-objective optimization code 11/3/2014 3/31/2015 107d 100%

7 1.2.2: [M1.3] Multi-objective optimization and selection of candidate designs 2/2/2015 4/30/2015 64d 100%

8 1.2.3: System level studies and design selection for prototyping 4/1/2015 4/30/2015 22d 100%

9 Task 1.3: Construction of WFSM Prototype 1 5/1/2015 9/25/2015 106d 100%

10 1.3.3: [M1.4] Construction of Prototype 1 5/1/2015 9/25/2015 106d 100%

11 Task 1.4: Design of Capacitive Coupler Drive Electronics 10/1/2014 1/1/2015 67d 100%

12 1.4.1: [M1.1] Initial electrostatic design 10/1/2014 11/14/2014 33d 100%

13 1.4.2: Power electronics topology selection and simulation 11/3/2014 1/1/2015 44d 100%

14 Task 1.5: Design of Capacitive Coupler 11/3/2014 6/1/2015 151d 100%

15 1.5.1: Hydrodynamic design of capacitive coupling plates 11/3/2014 2/27/2015 85d 100%

16 1.5.2: Integrated mechanical design of capacitive coupling 2/27/2015 5/1/2015 46d 100%

17 1.5.3: System metrics verification via solid modeling 5/1/2015 6/1/2015 22d 100%

18 Task 1.6: Initial Verification and Testing 10/1/2014 12/18/2015 318d 100%

19 1.6.1: Control code development for WFSM and tuning 10/1/2014 12/18/2015 318d 100%

20 1.6.2: Dynamometer modification and assembly 4/1/2015 10/30/2015 153d 100%

21 Task 1.7: Construction of Capacitive Coupler Prototype 5/20/2015 9/15/2015 85d 100%

22 1.7.1: Capacitive coupler prototype 1 construction and bench test 5/20/2015 9/1/2015 75d 100%

23 - Go/No Go Decision Point 1 9/15/2015 9/15/2015 1d 100%

24 Budget Period 2: Demonstration of WFSM 10/1/2015 9/30/2016 262d 100%

25 Task 2.1: Dynamometer Testing of Prototypes 9/25/2015 8/30/2016 243d 100%

26 2.1.1: [M2.1] Dynamometer testing of WFSM prototype 1 using brushes 9/25/2015 12/18/2015 61d 100%

27 2.1.2: [M2.2] Dynamometer testing of WFSM and capacitive coupler prototypes 1 12/21/2015 3/18/2016 65d 100%

28 2.1.3: [M2.4] Dynamometer testing of WFSM and capacitive coupler prototypes 2 6/1/2016 8/30/2016 65d 100%

29 Task 2.2: Design of WFSM Prototype 2 12/18/2015 3/15/2016 63d 100%

30 2.1.1: Redesign of WFSM prototype 2 12/18/2015 2/15/2016 42d 100%

31 2.2.2: Optimize WFSM Prototype 2 2/16/2016 3/15/2016 21d 100%

32 Task 2.3: Construction of WFSM Prototype 2 3/16/2016 6/1/2016 56d 100%

33 2.3.1: Construction of prototype 2 3/16/2016 6/1/2016 56d 100%

34 Task 2.4: Redesign of Capacitive Coupler Drive Electronics 10/1/2015 3/31/2016 131d 100%

35 2.4.1 Redesign of capacitive coupling prototype 2 10/1/2015 3/31/2016 131d 100%

36 Task 2.5: Construction of Capacitive Coupler Prototype 2 3/15/2016 6/1/2016 57d 100%

37 2.5.1: Construction and assembly of WFSM and capacitive coupler prototype 2 3/15/2016 6/1/2016 57d 100%

38 Task 2.6: Investigation of Power Take-Off Capability and Microgrid Support 1/1/2016 8/31/2016 174d 100%

39 2.6.1: [M2.3] Design control code for real and reactive power via voltage/frequency droop 1/1/2016 8/31/2016 174d 100%

40 2.6.2: Document performance experimentally in CERTS microgrid 8/1/2016 8/31/2016 23d 100%

41 - Go/No Go Decision Point 2 8/31/2016 8/31/2016 0d 100%

42 Task 2.7: Dissemination of Knowledge 10/1/2014 9/30/2016 523d 100%

*Note: The Gantt chart was updated for BP2 to reflect shifts in the project schedule. The PMP
and continuation documents reflect this change and it was accepted by program officers Steven
Boyd and Adrienne Riggi. Also a single quarter no cost extension (10/1/2016 – 12/31/2016) was
granted to finalize project outcomes.
In Summary: 1. ALL project go/no-go points were achieved. 2. The project objectives have been
achieved and milestones documented in the table below.

44
Budget Period 1 Milestones & Go/No-Go Decision 1

Comments (progress
toward achieving
Planned Actual
Milestone Title and milestone,
Completion Completion Verification Method
Description explanation of
Date Date
deviation from plan,
etc.)

Analytical and finite element


confirmation of capacitive
12/3/14 coupler transferring average
Initial Electrostatic 11/14/14,
field power [≥300 W] and peak Complete
Design Complete ongoing
complete field power [≥600W] with
limited electric fields [<1.5
MV/m]

Development of Optimization results for sample


combined thermal designs match detailed finite
and electromagnetic 3/31/15 3/31/2015 element modelling results within
WFSM multi- 15% for average torque, torque Complete
objective ongoing complete ripple, phase flux linkage and
optimization code within 20% for stator core
complete losses.

Candidate designs meet the


Multi-objective
following technical goals: 55
optimization and
kW peak power for 18 sec., 30
selection of
4/30/15 5/07/2015 kW power continuous, specific Complete
candidate designs for
power >1.3 kW/kg, power
prototyping
density >4.5 kW/l in
complete
optimization analysis.

Construction WFSM Selected design for prototype 1


Prototype 1 7/1/15 9/30/2015 constructed and ready for bench Complete
Complete testing.

Capacitive Coupling
Bench Test Experimentally confirm
Complete capacitive coupling transfers
9/15/15 8/11/2015 average field power [≥300 W] Complete
GO / No-Go 1 and peak field power [≥600W]
to dummy load.
(see below)

45
Budget Period 2 Milestones & Go/No-Go Decision 2

Comments
(progress
toward
Planned Actual
Milestone Title and achieving
Completion Completion Verification Method
Description milestone,
Date Date
explanation of
deviation from
plan, etc.)
Update Project Management Approved by program
10/31/2015 10/31/2015
Plan (PMP) officers
Design for prototyping meets
Now: the following technical goals The WFSM
WFSM Prototype 1 Initial 12/18/2015 55kW peak power for 18 achieved the
Dynamometer Testing with 12/18/2015 sec., 30 kW power metrics set in
Brushes Complete Originally: continuous, specific power the verification
10/15/2015 ≥1.3 kW/kg, power density ≥ column
4.5kW/l
The machine
The measured performance
was successfully
Now: of the WFSM Prototype 1
Dynamometer Testing of retrofitted with
3/18/2016 meets or exceeds the
WFSM and Capacitive CPC on the
3/18/2016 following specifications:
Coupler Prototypes 1 dyne, continued
Originally: specific power density [≥1.3
Complete testing in
1/14/2016 kW/kg], volumetric power
progress to
density [4.5≥kW/l]
achieve metrics
The simulation demonstrates
Simulations and
that the WFSM stator
reduced scale
Simulation Validation terminal voltage can be
6/30/2016 5/20/2016 lab demo show
Complete regulated with CPT without
a NEMA G2
the need for the main traction
regulation rating
drive.
The measured performance
Measurements
Now: of the WFSM meets or
of:
7/30/2016 exceeds the following
WFSM Performance – 1.95 kW/kg
7/12/2016 specifications: specific
Prototype 2 Achieved 7.19 kW/l
Originally: power density [≥ 1.6 kW/kg],
7/1/2016 volumetric power density [≥
At 4000 rpm
5 kW/l].
A reactive (RL)
Performance in CERTS load was
Micro-grid Achieved stepped in and
8/31/2016 The WFSM is able to
out of a
Now 9/30/2016 transfer real and reactive
GO / No-Go 2 common
10/31/2016 power to the micro-grid.
(see Section 12 of this coupling point
report) with regulated
line voltage

46
Demonstration of Go/No-Go metrics
Go/No-Go 1, 300W and 600W Capacitive Coupler Tests with External Resistive Load
This section is provided as reference to verify Go/No-Go 1 was satisfied in BP1 of the project,
and is an excerpt from quarterly report 4.
Initial tests of the capacitive power coupler were carried out with the WFSM rotor winding left
disconnected. Load was added to the dc output of the coupler assembly via a pair of slip rings
and brushes also mounted on the WFSM rotor. Initial tests used a 30 Ω resistive load at a
rotational speed of approximately 900 RPM.

Fig. 52. Driver and CPC circuit schematic

47
300W Test
 833kHz switching frequency.
 Soft switching
 Maximum voltage on switches:
384V/390V

a) Switching waveforms
 ~600V maximum voltage on the
coupling capacitors. (~300V for
each)

b) Coupling capacitor voltage


 94.94V output voltage
 3.27A output current
 310W output power
 Efficiency 80%

c) Output waveforms

Fig. 53. 300W test waveforms

48
600W Test
 833kHz switching frequency.
 Soft switching
 Maximum voltage on switches:
534V/541V

a) Switching waveforms
 ~870V maximum voltage on the
coupling capacitors. (~435V for
each)

b)
 132.54V output voltage
 4.54A output current
 602W output power
 Efficiency 83%

c) Output waveforms

Fig. 54. 600W test waveforms

Bench Test Conclusion: The circuit is able to deliver 300W to a 30  load with 80% efficiency
and 600 W with 83% efficiency. The winding on the test synchronize machine has 31.4 
resistance, and it is rated for 200W power dissipation. So the converter can be used on the field
winding for the traction machine.

Go/No-Go 2, Voltage regulation for a CERTS microgrid environment.


See section 12 of this report.

49
References

1. “DOE Energy Efficiency and Rewnewable Energy Vehicle Technologies Program Multi-Year
Program Plan 2011-2015.” Dec-2010.
2. D.M. Ionel, M. Popescu, M.I. McGilp, T.J.E. Miller, S.J. Dellinger, and R.J. Heideman,
“Computation of core losses in electrical machines using improved models for laminated
steel,” IEEE Transactions on Industry Applications vol. 43, no.6, pp. 1554-1564, 2007.
3. Dai J. and Ludois D., “A Survey of Wireless Power Transfer and a Critical Comparison of
Inductive and Capacitive Coupling for Small Gap Applications”, Power Electronics, IEEE
Transactions on. Vol. 30(11), pp. 6017-6029. IEEE, 2015.
4. Ludois DC, Erickson MJ and Reed JK, “Aerodynamic fluid bearings for translational and rotating
capacitors in noncontact capacitive power transfer systems”, IEEE Transactions on Industry
Applications. Vol. 50(2), pp. 1025-1033, 2014.
5. Di Gioia, A., Brown I.P., Knippel R., Hagen S., Ludois D.C., Nie Y., Dai J.J., Alteheld, C.,
"Design of a wound field synchronous machine for electric vehicle traction with brushless
capacitive field excitation," 2016 IEEE Energy Conversion Congress and Exposition (ECCE),
Milwaukee, WI, 2016, pp. 1-8.
6. J. Dai; S. Hagen; D. Ludois; I. Brown, "Synchronous Generator Brushless Field Excitation and
Voltage Regulation Via Capacitive Coupling Through Journal Bearings," in IEEE Transactions
on Industry Applications, early access online, to appear in July/August edition 2017.
7. Dai J and Ludois DC (2015), “Single active switch power electronics for kilowatt scale capacitive
power transfer”, IEEE Journal of Emerging and Selected Topics in Power Electronics. Vol. 3(1),
pp. 315-323, 2015.
8. Dai J and Ludois DC, “Capacitive Power Transfer Through a Conformal Bumper for Electric
Vehicle Charging,”, IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 4,
no. 3, pp. 1015-1025, Sept. 2016.
9. M. van der Giet, K. Kasper, R. De Doncker, and K. Hameyer, “Material parameters for the structural
dynamic simulation of electrical machines,” Electrical Machines (ICEM), 2012 20th International
Conference, pp. 2994–3000, Sept. 2012.

50

You might also like