Design Concept of A Lightweight Electric Motor Casing With Support From Thermomechanical Simulations

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Int. J. Automotive Composites, Vol. 2, No.

2, 2016 167

Design concept of a lightweight electric motor casing


with support from thermomechanical simulations

Timo Müller*, Gustavo Blazek and


Frank Henning
Institute for Vehicle Systems Technology (FAST),
Karlsruhe Institute of Technology (KIT),
Rintheimer Querallee 2, Building 70.04, 76131 Karlsruhe, Germany
Email: timo.mueller@kit.edu
Email: gustavo.blazek@gmail.com
Email: frank.henning@kit.edu
*Corresponding author

Abstract: Electric and hybrid electric vehicles and lightweight components are
important issues for the automotive industry in the pursuit for the reduction of
carbon emissions. In this study, a fibre-reinforced polymer (FRP) casing for a
high-power electric motor was designed, with focus on the cooling system
effectiveness. Heat transfer simulations were performed to prove the
effectiveness of different proposed cooling systems. A design concept using the
selected cooling system and 30 wt% glass fibre-reinforced polyamide-6 as
the material for the casing was submitted to a thermomechanical analysis.
Results show a maximum temperature of 105°C in the casing and stresses well
within the limits of the material. For this material, the design concept shows a
potential for ~65% weight reduction. The possibility of using other FRPs is
also briefly discussed.

Keywords: computer-assisted engineering; electric motor; fibre-reinforced


polymers; finite element analysis; heat transfer analysis; innovative composite
applications; lightweight design; motor cooling; polyamide-6; sequentially
coupled thermal stress analysis.

Reference to this paper should be made as follows: Müller, T., Blazek, G. and
Henning, F. (2016) ‘Design concept of a lightweight electric motor casing with
support from thermomechanical simulations’, Int. J. Automotive Composites,
Vol. 2, No. 2, pp.167–186.

Biographical notes: Timo Müller studied Mechanical Engineering at KIT.


Since 2009, he is a Research Assistant at the Institute for Vehicle Systems
Technology (FAST) at KIT. His current research interests are simulation and
characterisation of long fibre-reinforced thermoplastics.

Gustavo Blazek is a Research Associate at the Institute for Vehicle Systems


Technology (FAST) of the Karlsruhe Institute of Technology (KIT). His
research interest is focussed on the simulation of the thermomechanical
behaviour of long fibre-reinforced polymers. He has formerly worked on the
thermal and structural characterisation of biodegradable polymeric blends at the
Metallurgical and Material Engineering Department of the Polytechnic School
at University of Sao Paulo (EPUSP), Brazil.

Frank Henning is a Professor for Lightweight Technology at the Institute of


Vehicle Systems Technology (FAST) of the Karlsruhe Institute of Technology

Copyright © 2016 Inderscience Enterprises Ltd.


168 T. Müller et al.

(KIT), Deputy Director of the Fraunhofer Institute for Chemical Technology


(ICT), Professor at the Faculty of Mechanical Engineering and Material
Sciences and Managing Director of the Fraunhofer Project Centre for
Composites Manufacturing at the University of Western Ontario, London,
Canada. His current research interests include the development of methods for
the design of fibre-reinforced lightweight structures through material
modelling, numerical simulation and experimental verification and the
development of production technologies for fibre-reinforced structures and
parts.

1 Introduction

In recent years, several factors have been pushing the automotive industry towards the
use of lightweight components. Initially, fuel price fluctuations and the thrive for high
performance in sports cars stimulated some manufacturers to use fibre-reinforced
polymers (FRPs) as a lighter alternative, but fuel economy alone will not drive weight
reduction of average commercial vehicles (Brady and Brady, 2008b). Due to the growing
concern about the effect of greenhouse gases on the global climate and with cars and vans
representing approximately 12.5% of the CO2 emission in the European Union (EU), new
European regulations have been set to reduce the carbon emissions of the average car
fleet to 130 and to 95 g of CO2 per kilometre by 2015 and 2020, respectively (Transport
and Environment, 2011). These new regulations, combined with increased environmental
awareness, have been driving sports vehicles OEMs not only towards the development of
less environmentally impacting cars, such as electric vehicles and hybrid electric
vehicles, but also towards the search for lightweight solutions (Kastensson, 2014). This
search plays a major role in the increasing demand for FRP components (Osborne, 2013).
Brady and Brady (2008a) report that, in the third annual Automotive Engineering
Plastics Conference (AutoEPCON), many of the obvious substitutions from metal to
polymer have already been made, but newly available additives, plastics and composites
allow us to revisit applications which, due to their requirement for high temperature
resistance and dimensional stability, were usually taken as non-feasible for polymer-
based materials.
Predictive models of fibre composites are essential tools in the development of such
composite components. They reduce costs and time on expensive and time demanding
experimental procedures and prototyping. Those models have become more and more
accurate, helping to overcome the resistance towards change, sometimes present in the
automotive industry (Brady and Brady, 2007, 2008b).
When approaching the development of an electric motor, Kim, Kim and Kim (2013)
and Staton, Boglietti and Cavagnino (2005) point out that, due to the increasing demand
for smaller, more efficient motors with higher torque and output, the thermal design and
cooling technique have to be carefully observed, giving thermal simulations increased
importance in this process. Moreover, when working with polymeric matrix composites,
it is crucial to predict the temperatures reached during service and the materials response
to the applied loads at such temperatures.
Kim, Kim and Kim (2013) performed heat flow analysis to analyse the feasibility of
applying an air cooling system to an in-wheel electric motor and used the obtained results
to propose an improvement method for the cooling structure of the in-wheel motor to
Design concept of a lightweight electric motor casing 169

optimise its cooling performance. Lohse-Busch (2004) developed a lumped capacitance


finite-difference model to simulate the thermal overload conditions of a motor and
inverter unit. The inverter was tested individually and in assembly with the motor, under
regular operational conditions, and the experimental data obtained were used to calibrate
the model. The model was then used to simulate double current conditions. Based on the
obtained simulation results, design changes were proposed to improve the thermal flow
from heat source to heat sink. Chin and Staton (2004) used an analytical lumped circuit
model and a finite element analysis (FEA) model to simulate the thermal behaviour of a
traction motor. When comparing the results with experiments, they found that the lumped
circuit model provides good results with high efficiency, but the FEA model offers more
insight into the temperature variation in the structure of the motor.
In the present work, a lightweight casing concept for a liquid-cooled electric motor is
developed using heat transfer and sequentially coupled thermal stress analysis. Injection-
moulded glass fibre-reinforced polyamide is used for cost efficiency. The FRP casing
makes a necessary different design approach. Because of the complex geometry of
thermally conductive paths, it would be difficult to define a lumped circuit model, and the
FRP is much more sensitive to temperature than metallic casings, coupling the
mechanical performance of the casing with the thermal behaviour. Therefore, a FEA
approach which allows to model thermal and mechanic behaviour is adopted.

2 State-of-the-art electric motor

A state-of-the-art (SoA) liquid-cooled permanent magnet electric motor was taken as the
starting point for the development. The motor consists of an internal rotor with permanent
magnets and will not be presented in this paper. The stator is composed of copper
windings and a steel-laminated core enclosed in an aluminium casing with incorporated
stainless-steel cooling pipes and a connection box, as represented in Figure 1. The motor
produces a continuous power of 79 kW and a torque of 260 Nm. The isolation of the
stator windings and plastic slot liners limits the operating temperature of the stator to
160°C.

Figure 1 Simplified representation of the SoA motor parts without the rotor (see online version
for colours)
170 T. Müller et al.

2.1 Heat transfer analyses of unaltered SoA motor design


A heat transfer analyses of the SoA motor were made to check the validity of the model
and to serve as a basis for comparison. A heat flux of 4 kW represents the power losses in
the windings. The ambient air is completely still and has a temperature of 40°C.
A constant stream of water with a temperature of 40°C is used for cooling. The
manufacturer specifies an inlet temperature of the cooling liquid of 35°C, and a
maximum temperature at the outlet of 45°C, for a flow rate between 5 and 10 l/min. For
the simulation, the water temperature is set to be a constant 40°C through the cooling
system. This is in very good agreement with the values. Kral, Haumer and Bauml (2008)
provided for a smaller but similar fluid-cooled motor. The air inside the motor has a
temperature of 80°C.
Gmelin et al. (1999) compiled an extensive overview of published values for thermal
contact conductance between different material combinations, including stainless steel,
copper and aluminium.
The thermal contact conductance for the material combination of stainless steel and
aluminium was set to 2900 W/m2K at 10 MPa contact pressure and room temperature
according to Cengel (2008). He does not cite a source for the values he presents, but the
values are in good agreement with the values presented by Mills (1999), who gives a
range from 3000 to 4500 W/m2K for moderate pressures and usual finishes. Staton,
Boglietti and Cavagnino (2005) used these values for the simulation of electric motors.
Lewis and Perkins (1968) measured values ranging from approximately 200 to
900 BTU/hft2°F for a pressure range from 0 to 500 psi. In SI units, this amounts to
1136–5110 W/m2K for a pressure range between 0 and 3 MPa. More recently, Fukuoka
and Nomura (2013) measured values for the combinations of carbon steel, stainless steel
and aluminium. For the combination of stainless steel and aluminium at a pressure of 10
MPa, a temperature of approximately 50°C and a surface roughness of 6.4 µm, a value
just below 3000 W/m2K was presented. They also proposed an empirical equation for the
thermal contact coefficient between dissimilar metals. Directional effects on the contact
conductance were measured by Lewis and Perkins and Fukuoka and Nomura but were
not included in this analysis. Fieberg and Kneer (2007) measured similar values for
contact pressures below 10 MPa, with the contact conductance showing a strong increase
with rising contact pressure, similar to the measurements of Parmenter and Marschall
(1995). The value for the combination of stainless steel and aluminium was also used for
the contact conductance between stator and housing. Staton, Boglietti and Cavagnino
(2005) and Malumbres et al. (2015) suggest that the effective value could be lower,
although Malumbres et al. used a perfect contact in their calculations. Because the stator
is pressed into the aluminium casing in this particular case, we assumed a good contact.
Recently, Camilleri, Howey and McCulloch (2014) measured the contact conductance
between stator laminations and aluminium casings with different press fits and found
values between 500 and 5000 W/m2K. Due to the fact that the actual contact pressure was
not known, using 2900 W/m2K should be a good approximation.
Design concept of a lightweight electric motor casing 171

The thermal contact conductance for the material combination of aluminium and
aluminium was set to 11400 W/m2K for 150°C and a pressure between 1.2 and 2.5 MPa
according to Cengel (2008). Fukuoka and Xu (1999) measured the contact conductance
for pairs of carbon steel, stainless steel and aluminium, different surface finishes and
contact pressures. He also proposed an empirical formula for the contact conductance.
Using the values he provided for aluminium AW2024, a contact pressure of 1.6 MPa and
a total surface roughness of 6.4 µm, the formula yields a contact conductance of
11380 W/m2K, correlating well with the values proposed by Cengel. Parmenter and
Marschall (1995) measured a similar value for contact pressures between 1 and 2 MPa for
aluminium, after a blasting with fine beads. Fukuoka and Xu (1999) and Parmenter and
Marschall (1995) both found a strong increase of the contact conductance with rising
contact pressure. This was not considered in this analysis but would only improve the
heat conduction.
The heat transfer coefficients in the cooling pipes and on the outer surface of the
housing were calculated according to the formulas compiled by Staton and Cavagnino
(2008). The result for the convection coefficient in the cooling pipes was 8500 W/m2K
similar to the value calculated by Alexandrova, Semken and Pyrhönen (2014). The
coefficient for forced convection on the end windings was calculated with Schubert’s
formula, taking the mean velocity of the rotor at rated speed as the average air speed,
resulting in the value of 56 W/m2K. This correlates well with typical values [see, for
example, Staton, Boglietti and Cavagnino (2005)]. The properties used for the thermal
simulation are shown in Table 1. The difference in axial and radial conductivity of the
lamination stack was not considered. According to Staton and Šusnjic (2009), the
conductivity in an axial direction can be 20–40 times smaller due to the insulating layers
between the individual sheets.
Table 1 Thermal properties of the components of the SoA model

Thermal conductivity
Material (W/mK) Specific heat (J/gK)
Stainless steel 24.3 460
32.9 at 20°C
Lamination M530-50A 500
36.7 at 150°C
Aluminium 167 896

The software was set to simulate heat generation in the stator for a motor running at rated
power until thermal equilibrium was reached. The temperature distribution for the SoA
aluminium casing is shown in Figure 2. A maximum temperature of around 84°C is
observed in the casing as well as a smooth temperature distribution. A maximum
temperature of 98°C can be observed in the stator core. These values are in agreement
with those provided by the manufacturer and also with the simulations by Kral, Haumer
and Bauml (2008).
172 T. Müller et al.

Figure 2 SoA aluminium casing: maximum casing temperature is around 84°C (see online
version for colours)

Note: Temperature profile stabilises within 40 min


To illustrate the need for a new design concept for a FRP casing, the aluminium in the
SoA motor design was replaced with FRP. The values used for the contact conductance
coefficient in between FRP and metallic parts are approximated by the values presented
by Marotta and Fletcher (1996) for the contact pair PA6 and aluminium (400 W/m2K).
The simulation did not reach a thermal equilibrium at rated power, the temperature of the
stator rises until stator and casing are no longer operational. The whole stator and parts of
the polymeric casing reach temperatures above 160°C. These temperatures would cause
not only the melting of the slot liners in the stator but also the degradation of the
polymeric casing.
Such high and poorly distributed temperatures are a consequence of the lower thermal
conductivity of the PA6GF30 (0.35 W/mK) as compared to aluminium (167 W/mK).
Even if thermal conductive formulations of PA6 or other FRPs were used to replace the
aluminium in this design, the limitations generated by a thermal conductivity of the two
orders of magnitude lower than the original would still have to be overcome. Even when
using carbon fibres or fillers to improve the thermal conductivity, the difference would
stay in the same order of magnitude [see, for example, Li et al. (2013)].
A particular feature of the SoA casing design has a great influence on the motor
temperature. In this design, the casing envelops the cooling system, preventing the direct
contact between the steel pipes and the stator. For a polymer-based casing, this creates an
insulating layer between the heat source and the heat sink, limiting the heat dissipation
capacity of the system.
Through this analysis, it becomes clear that a simple material substitution in the
unaltered SoA design does not offer a solution. A new design concept focussing on the
cooling capacity has to be developed.
Design concept of a lightweight electric motor casing 173

3 Design of a new cooling system concept

For fully enclosed liquid-cooled motors the use of a metallic water jacket along the outer
diameter of the stator is a proven and effective solution and motors of this type have been
described by Crescimbini et al. (2005), Kral, Haumer and Bauml (2008) and
Jungreuthmayer et al. (2011). The goal for the new design was to have a fully enclosed
cooling system, made of stainless steel, with the same cooling capacity as the SoA
casing. The stainless steel allows the safe operation with different cooling fluids without
corrosion problems, while the closed system prevents water leaking into stator or other
components. At the same time, the new design should be lighter and more economical
than a metallic water jacket. To dissipate the heat generated in the stator, a thermal
conductive path between the copper windings and the cooling system must be established
without insulating layers of FRP. Round stainless-steel pipe was chosen for the cooling
system because it is can be formed into complex shapes by bending, leaking is
improbable, and the stock is readily available and can be overmoulded even at high
injection pressures.
Different strategies can be applied to improve the thermal conductivity between stator
and cooling system:
• Bring the pipes directly into contact with the stator either through a simple line
contact or through deformation of the pipes by 1 or 2 mm to obtain a flat contact
area.
• Place a thin aluminium jacket around the stator to act as a heat conductor.
• Add aluminium clamping profiles to improve the thermal contact surface between
the stator and undeformed pipes. Profiles covering only part of the circumference of
the stator and profiles with a continuous aluminium jacket were considered.
• Increase the contact length of the cooling pipe along the stator.
These possibilities are combined with a FRP casing resulting in the configurations
described in Table 2. The basic wall thickness of 3 mm for the FRP casing was chosen
for production reasons.
Table 2 Proposed configurations

Number of Casing
pipes Profiles/jacket Type of pipe thickness (mm)
Base 16 None Deformed - 2 mm 3
configuration
16 Jacket with Non-deformed 3
profiles
Clamps/jacket
16 Small profiles Non-deformed 3
variations
16 Jacket - no Deformed - 2 mm 3
profiles
Varying number 12 None Deformed - 2 mm 3
of pipes
14 None Deformed - 2 mm 3
18 None Deformed - 2 mm 3
20 None Deformed - 2 mm 3
174 T. Müller et al.

Table 2 Proposed configurations (continued)

Number of Casing
pipes Profiles/jacket Type of pipe thickness (mm)
Type of pipes 16 None Non-deformed 3
variations 16 None Deformed - 1 mm 3
16 None Deformed - 2 mm 2
Casing thickness
16 None Deformed - 2 mm 4
variations
16 None Deformed - 2 mm 5

Simulations were used to select the most appropriate configuration.

3 Simplified model - cooling system selection

To analyse the heat dissipation potential of the configurations, a cross-section of the


motor including part of the stator, casing and cooling pipes was modelled. Symmetry was
used whenever applicable. For this model, the copper windings and the steel-laminated
core of the stator were represented separately. Figure 3 depicts the configuration using a
jacket with clamping profiles. Two of the proposed designs are exhibited in Figure 4.

Figure 3 Simplified model representing a quarter of the motor cross-section (variant-aluminium


jacket with clamping profiles and non-deformed pipes)

Note: The stator consists of the laminated core and the copper windings

Figure 4 Two of the proposed configurations

Note: White areas represent the polymeric casing and black the aluminium
clamps. Left - 2 mm deformed pressed pipes; right - aluminium clamps
promote the thermal contact between stator and non-deformed pipes
Design concept of a lightweight electric motor casing 175

A heat transfer analysis of each configuration was performed until thermal equilibrium
was reached and a stable temperature profile was obtained. The maximum temperatures
of the configurations are depicted in Figure 5.

Figure 5 Top to bottom, left to right - maximum temperature variation according to the type of
aluminium clamps/jacket, pipe geometry, casing thickness and number of pipes

The results were analysed to find a design which would combine casing temperatures in
the order of 100°C, which is well below the continuous service temperature for PA6GF30
stated by the manufacturer and a minimal use of metallic components for weight
reduction purposes.
The configuration chosen was the one with a 3-mm thick casing and 16 undeformed
cooling pipes, connected to the stator through small aluminium clamping profiles. The
heat transfer simulation of this configuration indicates that an efficient thermal
connection can be established through the small profiles, obtaining a satisfactory service
temperature range. The use of a continuous aluminium jacket with clamps, or of a higher
number of steel pipes, would unnecessarily increase the overall mass and cost of the
system without improving the cooling capacity.

4 Concept motor casing design

Once the variant had been chosen, the whole set of cooling system, stator and casing was
fully designed and modelled so a complete thermomechanical simulation could be
performed to prove the capacity of the structure to support the defined thermal and
mechanical loads.
The cooling system consists of 16 aluminium clamping profiles placed around the
stator in an equally spaced circular pattern. In this particular machine, the winding heads
have a larger diameter than the rest of the stator due to the automation of the winding
process of the copper wires. This prohibits extending the clamping profiles and cooling
pipes over the winding heads. Due to the fact that the winding heads are typically the
most critical parts for cooling, two aluminium rings were integrated to promote thermal
connection between the cooling pipes and the stators winding heads. The rings are
machined to fit over the winding heads and the stator. In most other configurations,
176 T. Müller et al.

where the diameter of the winding heads is smaller than the stator, the shape of the rings
could be simplified to a basic shape. The rings also serve a second purpose, providing a
defined surface to position the stator and cooling system assembly in the mould during
the overmoulding process. The uneven surface of the winding heads, even when
encapsulated with resin, is not suitable for clamping and sealing of the mould cavity. The
cooling system is depicted in Figure 6.

Figure 6 Individual drawing of an aluminium clamping profile and full design of the cooling
system

The 3-mm thick PA6GF30 casing is then injection moulded over the stator and cooling
system. In the developed concept, the connection box is to be injected with the casing as
a one piece component, while the rear end plate, also made of PA6GF30, is produced
separately. The front end plate can be made of FRP or aluminium depending on the
mechanical demands. The material used for the front end plate was PA6GF30. The casing
concept design is illustrated in Figure 7.

Figure 7 Casing concept design with connection box incorporated. Cooling system and end
plates are also depicted

4.1 Mould filling simulation - concept motor casing


A mould filling simulation was performed to calculate the fibre orientation in the FRP
casing. A Cross-WLF viscosity model and a two domain Tait equation of state were used
to simulate the moulding material with parameters provided by the material supplier. The
Folgar-Tucker model was used to predict the fibre orientation. The software
MOLDFLOW was used for the simulation.
Four gate locations were defined along the circumference of the casing, coinciding
with the bosses used to mount the front end plate. The mould was filled in 2.6 s with a
maximum injection pressure at the gate of 67 MPa. Due to the long flow length, a high-
fibre orientation along the casing is achieved as illustrated in Figure 8.
Design concept of a lightweight electric motor casing 177

Figure 8 Fibre orientation in the FRP casing (see online version for colours)

4.2 Heat transfer simulation - concept motor casing


A heat transfer simulation of the new electric motor design was performed until thermal
equilibrium was reached. The assumptions made for this simulation are the same as the
one made for the SoA model, except that the stator is now divided into copper windings
and the steel-laminated core. The properties of pure copper were assumed for the
winding, a more detailed approach as described by Boglietti, Cavagnino and Staton
(2008) was not used, because of the missing information about the components and
structure of the winding. An experimental setup, like the one used by Hoang et al. (2010),
was not available. Instead, an additional thermal contact conductance between the
winding and metallic parts was used, for the pair aluminium-copper a value of 12000
W/m2K according to Cengel (2008) and a value of 10000 W/m2K for the pair
steel-copper approximated from Gmelin et al. (1999). Both the values are relatively
high, but the windings in this particular motor where encapsulated with a high-
thermal conductivity epoxy resin, which should lead to a relatively low resistance
(Crescimbini et al., 2008).
The thermal conductivity of the FRP was assumed to be isotropic with a value of
0.359 W/mK. In reality, the different thermal conductivities of matrix and fibre lead to
anisotropic behaviour depending on the fibre orientation. For glass fibres, this effect is
less pronounced [see, for example, Choy et al. (1992)]. Due to the generally low values
compared to the metallic parts, the anisotropy of the FRPs thermal conductivity was
neglected. The results of the heat transfer analysis are exhibited in Figures 9 and 10.
The results show a maximum temperature of 113°C in the copper windings, well beneath
the previously established limit of 160°C. The highest temperatures are reached in the
winding heads. In the FRP parts, a maximum temperature of around 105°C is observed in
the contact surface to the winding head cooling rings. This underlines the importance of
the presence of the aluminium cooling rings, which promote the thermal connection
between this critical area and the cooling pipes. If the cooling pipes could be extended
over the winding heads, the temperature could be reduced to the 90°C range seen in the
middle of the stator.
178 T. Müller et al.

Figure 9 Temperature profile of the concept design - maximum temperature of 113°C in the
copper winding heads (see online version for colours)

Figure 10 Temperature profile of the concept casing - maximum temperature 105°C (see online
version for colours)

The casing temperature distribution observed at Figure 10 shows temperatures ranging


mostly from 80 to 90°C in the internal area, especially at the regions of direct contact
between stator and FRP casing. The external surface temperature of the casing nears
90°C on the area above the cooling rings, while the other regions reach maximum
temperatures of circa 81°C, but are, mostly in lower ranges of 50–65°C.

4.3 Thermomechanical simulation - concept motor casing


For the thermomechanical analysis, the temperature profile obtained in the previous
section was used as input. Mechanical loads and thermal expansion coefficients were
Design concept of a lightweight electric motor casing 179

added to the model as well as the temperature-dependent mechanical properties of


PA6GF30.
The mechanical behaviour of PA6GF30 is complex and depends on many factors
especially the orientation, length and volume fraction of the reinforcing fibres and the
influence of temperature, strain rate and moisture content on the matrix material.
To model the influence of the fibres, local fibre orientations were mapped from the
mould filling simulation. Isotropic elastic behaviour was assumed for PA6 matrix and
glass fibres. A homogenisation scheme as proposed by Mori and Tanaka (1973) in the
form described by Benveniste (1987) was used to generate the elastic properties of the
composite with unidirectional aligned fibres. A fibre aspect ratio of 16.7 was assumed.
Eshelby's tensor for isotropic matrix and inclusion was utilised (Mura, 1982). Orientation
averaging, as proposed by Advani and Tucker (1987), using the orthotropic fitted closure
by Cintra and Tucker (1995) to compute the fourth-order orientation tensor was then used
to compute the elastic properties for all local fibre orientations. This approach allows
calculating stress states in fibre and matrix. Mises equivalent stress in the matrix was
used as yield criterion. Pressure-dependent yield criteria would be more accurate for
polyamide [see, for example, Donato and Bianchi (2012) or Caddell, Raghava and Atkins
(1974)] but the necessary material test data were not available at this stage. The same was
true for the yield of the composite. The constitutive behaviour was implemented as a
user-defined subroutine for Abaqus.
More advanced models, taking plastic behaviour of the matrix into account, for
example Nguyen et al. (2009) or Kammoun et al. (2011), were considered but deemed to
complex at this stage of development. They should be considered in later stages of
development.
The influences of temperature and moisture on the elastic properties were considered
by a temperature-dependent modulus of the matrix based on manufacturers data,
measured on samples which were conditioned until they reached their saturation moisture
content (see Table 3). The properties of the glass fibres were considered constant in this
temperature range.
Table 3 Temperature-dependent modulus and yield strength at saturation moisture content

Composite
Matrix elastic Matrix yield elastic modulus Composite yield
Temperature (°C) modulus (MPa) stress (MPa) (MPa) stress (MPa)
0 3570 64.29 7842 145.28
23 1500 30.72 5179 107.11
40 825 17.30 4554 95.80
60 565 14.5 4138 83.69
100 450 13.11 3156 65.27
150 385 9.5 2826 (140°C) 51.50 (140°C)

An important step for improving the accuracy of the simulation would be to simulate the
creep behaviour of the FRP, which would be used to calculate the reduction of the
residual stresses after moulding, and also to predict if critical deformations of the material
would occur during the life span of the motor. This should be considered in the next stage
of development.
180 T. Müller et al.

A structure simulation was performed to obtain the stress distribution in the system,
taking into consideration also the strain and stress generated due to the thermal expansion
of the components. The inclusion of thermal expansion effects is especially important
when dealing materials with different thermal expansion coefficients. In this case, the
coefficient used for PA6GF30 is two to three times larger than the coefficient for
aluminium and seven to eight times larger than the one for steel. The orthotropic
coefficients were calculated depending on the fibre orientation.
The mechanical loads considered are the following:
• the weight of the system automatically calculated by Abaqus through the density
input
• the rotor’s weight added as a load at the ball bearings
• a torque of 260 Nm applied to the stator.
The motor is fixed in space by the front end plate and beam elements are used to simulate
the bolts connecting both end plates to the casing. The resulting stress distribution is
shown in Figures 11–13.
Figure 11 shows the stress in the composite in the direction of the main fibre
orientation. Stresses range from 24 to −19 MPa well within the limit of the material and
the measured strength of the composite is 95 MPa at this temperature. The highest
stresses are concentrated in the transition from the bosses to the casing. The distribution
of tensile and compressive stresses in this area suggests torsion or bending loads probably
due to the torque of the engine. The geometry in this area should be improved to lower
the stresses, for example, through the use of numerical topology optimisation. Figure 12
shows the equivalent stress in the matrix of the composite reaching 12 MPa close to the
bolts and approximately 5 MPa where the composite shows the highest loads.

Figure 11 Stress in direction of the maximum fibre orientation in the casing at 40°C (see online
version for colours)

Note: Circles indicate stress concentrations


Design concept of a lightweight electric motor casing 181

Figure 12 Equivalent matrix stress in the casing at 40°C (see online version for colours)

Note: Circles indicate stress concentrations


Figure 13 shows the stress in the composite in the direction of the main fibre orientation
at operating temperature. The distribution is similar, although the values increase,
especially the compressive values. The simulation does not predict any damage although
the safety margins are smaller compared with the measured strength of 65 MPa.

Figure 13 Stress in direction of the maximum fibre orientation in the casing at operating
temperature (see online version for colours)

Note: Circles indicate stress concentrations


182 T. Müller et al.

In both load cases, the stress in all parts of the casing apart from the bosses is below
10 MPa.

5 Discussion

The simulation results indicate that the use of fibre-reinforced polyamide-6 as material
for a motor casing is feasible when the design is developed keeping in mind the material
characteristics and aiming for optimal heat flux in between heat source and heat sink. For
promoting such thermal contact, a cooling system with small strategically placed
aluminium parts, as the one proposed in this project, proved to be efficient. Profiles of
this kind can be efficiently produced from extruded aluminium.
When using PA6GF30, the concept developed shows a high weight reduction
potential in comparison with the SoA aluminium casing.
For manufacturing the frontal end plate, which is used for mounting the motor,
PA6GF30 could also be used according to the stress distributions obtained. If a higher
safety margin is required, it could also be manufactured from aluminium. In any of these
cases, it also offers a potential for weight reduction when compared to the SoA frontal
end plate due to dimensional changes made for the concept design.
The whole SoA casing, made in aluminium, has a total weight of circa 11 kg, while
the concept PA6GF30 casing with an aluminium front end plate, a stainless-steel cooling
pipe and aluminium clamping profiles and cooling rings weights around 4 kg. These
represent a reduction of approximately 65% of the original weight as detailed in the table
below.
Table 4 Weight reduction potential

Mass - concept casing Mass - SoA casing


Part (kg) (kg)
Casing (PA6GF30) 1.17 7.44
End plate A (aluminium) 0.683 1.614
End plate B (PA6GF30) 0.439 1.840
Cooling pipes 0.723 Not provided
Aluminium clamping profiles 0.447 –
Cooling rings and copper winding 0.303 –
heads support
Connection box lid (PA6GF30) 0.064 –

Threaded inserts 0.024 –


Sum 3.853 (65% reduction) 10.894

By using other materials than PA6GF30, or even different formulations of it, the weight
reduction potential might vary as well as the temperature and stress distributions. Bulk
moulding compound formulations, for instance, might offer better mechanical properties
and dimensional stability at higher temperatures but at the same time pose new
challenges for the manufacturing process.
Fibre-reinforced polymers formulated with special fillers to enhance thermal
conductivity usually have higher density values but will offer a more homogeneous
Design concept of a lightweight electric motor casing 183

temperature distribution and lower temperatures during the motors service. High
temperature plastics, like PA66, PEEK and PPS, will lead to an increase in costs but will
also improve the mechanical performance of the part. Increasing the fibre volume
fraction and the fibre length, by using a long fibre-reinforced material with 40 wt% glass
fibres, for example, would be a very efficient alternative to improve material performance
especially against creep.
In all cases, simulative methods allow the prediction of the behaviour of the part,
helping developers to find the balance between, for instance, lightweight potential and
mechanical/thermal performance as well as between other desired part features. They also
facilitate the feasibility analysis of bringing materials into new applications. In
summation, simulative tools play a major role at the support of the development process,
reducing experimental and prototyping costs.

6 Conclusion

Heat transfer analyses performed for a SoA electric motor design show that a simple
substitution of PA6GF30 for aluminium as casing material is not feasible, and that design
changes promoting the heat flux between heat source and heat sink are needed to develop
a lightweight FRP casing.
Through heat transfer simulations of simplified models, it was found that a system of
small aluminium clamping profiles provides an efficient thermal contact between stator
and cooling pipes, and that the use of a full aluminium jacket is not needed.
A design of a new concept for the electric motor casing and cooling system was
proposed, including two aluminium rings around the winding heads for improved
cooling. The thermomechanical analysis of the new design concept confirmed the
predicted effectiveness of the proposed cooling system with the results showing that the
proposed design and material would be able to fulfil the mechanical and thermal
demands.
The new PA6GF30 casing concept shows a 65% weight reduction potential when
compared to the aluminium SoA casing. The design could be further improved by using
numerical optimisation techniques, especially in the area where casing and end plates are
hold together by screws. Even if thermomechanical fatigue or creep in some applications
would require a more durable casing, the weight reduction should still be substantial.
The results of this work not only indicate that PA6GF30 is a feasible alternative
material for this specific application but also encourage the review of the role of polymer-
based composites in higher temperature ranges. The use of an injection moulding to
produce the complete casing should also offer economic benefits.

Acknowledgements

The authors would like to thank the German Federal Ministry for Economic Affairs and
Energy for supporting this work. The authors would also like to express our gratitude to
Jonas Friedrich for his design of the casing, and Marina Mrkonjic for revising the model
of the casing.
184 T. Müller et al.

References
Advani, S.G. and Tucker, C.L. (1987) ‘The use of tensors to describe and predict fiber orientation
in short fiber composites’, Journal of Rheology, Vol. 31, pp.751–784. Available at
http://dx.doi.org/10.1122/1.549945.
Alexandrova, Y., Semken, R.S. and Pyrhönen, J. (2014, April) ‘Permanent magnet synchronous
generator design solution for large direct-drive wind turbines: thermal behavior of the LC DD-
PMSG’, Applied Thermal Engineering, Vol. 65, Nos. 1–2, pp.554–563. ISSN 1359-4311.
Available at http://dx.doi.org/10.1016/j.applthermaleng.2014.01.054.
Benveniste, Y. (1987) ‘A new approach to the application of Mori-Tanaka's theory in composite
materials’, Mechanics of Materials, Vol. 6, No. 2, pp.147–157. ISSN 0167-6636. Available at
http://dx.doi.org/10.1016/0167-6636(87)90005-6.
Boglietti, A., Cavagnino, A. and Staton, D. (2008, July–August) ‘Determination of critical
parameters in electrical machine thermal models’, IEEE Transactions on Industry
Applications, Vol. 44, No. 4, pp.1150–1159. Available at http://dx.doi.org/10.1109/
TIA.2008.926233.
Boglietti, A., Cavagnino, A., Staton, D., Shanel, M., Mueller, M. and Mejuto, C. (2009, March)
‘Evolution and modern approaches for thermal analysis of electrical machines’, IEEE
Transactions on Industrial Electronics, Vol. 56, No. 3, pp.871–882. Available at
http://dx.doi.org/10.1109/TIE.2008.2011622.
Brady, M. and Brady, P. (2007) ‘Automotive composites—the search for efficiency, value and
performance’, Reinforced Plastics, Vol. 51, No. 1, pp.26–29.
Brady, M. and Brady, P. (2008a) ‘Development in automotive plastics’, Reinforced Plastics,
Vol. 52, No. 10, pp.37–40.
Brady, M. and Brady, P. (2008b) ‘The road to lightweight performance’, Reinforced Plastics,
Vol. 52, No. 10, pp.32–36.
Caddell, R.M., Raghava, R.S. and Atkins, A.G. (1974) ‘Pressure dependent yield criteria for
polymers’, Materials Science and Engineering, Vol. 13, No. 2, pp.113–120. ISSN 0025-5416,
Available at http://dx.doi.org/10.1016/0025-5416(74)90179-7.
Camilleri, R., Howey, D.A. and McCulloch, M.D. (2014) ‘Experimental investigation of the
thermal contact resistance in shrink fit assemblies with relevance to electrical machines’, IET
Power Electronics Machines and Drives Conference (PEMD), Available at http://dx.doi.org/
10.1049/cp.2014.0472.
Cengel, Y.A. (2008) Introduction to Thermodynamics and Heat Transfer, 2nd ed. The McGraw-
Hill Companies, New York.
Chin, Y.K. and Staton, D.A. (2004) ‘Transient thermal analysis using both lumped-circuit approach
and finite element method of a permanent magnet traction motor’, AFRICON, 2004. 7th
AFRICON Conference in Africa, Vol. 2, pp.1027–1035. 15–17 September, 2004. Available at
http://dx.doi.org/10.1109/AFRICON.2004.1406847.
Choy, C.L., Leung, W.P., Kowk, K.W. and Lau, F.P. (1992) ‘Elastic moduli and thermal
conductivity of injection-molded short-fiber–reinforced thermoplastics’, Polymer Composites,
Vol. 13, pp.69–80. Available at http://dx.doi.org/10.1002/pc.750130202.
Cintra, J.S. and Tucker, C.L. (1995) ‘Orthotropic closure approximations for flow‐induced fiber
orientation’, Journal of Rheology, Vol. 39, pp.1095–1122. Available at http://dx.doi.org/
10.1122/1.550630.
Crescimbini, F., Di Napoli, A., Solero, L., and Caricchi, F. (2005, September–October) ‘Compact
permanent-magnet generator for hybrid vehicle applications,’ IEEE Transactions on Industry
Applications, Vol. 41, No. 5, pp.1168–1177. Available at http://dx.doi.org/10.1109/
TIA.2005.855048
Donato, G.H. and Bianchi, M. (2012, April–June) ‘Pressure dependent yield criteria applied for
improving design practices and integrity assessments against yielding of engineering
polymers’, Journal of Materials Research and Technology, Vol. 1, No. 1, pp.2–7. Available at
http://dx.doi.org/10.1016/S2238-7854(12)70002-9.
Design concept of a lightweight electric motor casing 185

Fieberg, C. and Kneer, R. (2008, March) ‘Determination of thermal contact resistance


from transient temperature measurements’, International Journal of Heat and Mass Transfer,
Vol. 51, Nos. 5–6, pp.1017–1023. Available at http://dx.doi.org/10.1016/j.ijheatmasstransfer.
2007.05.004.
Fukuoka, T. and Nomura, M. (2013) ‘Evaluation of thermal contact resistance at the interface of
dissimilar materials’, ASME, Journal of Pressure Vessel Technology, Vol. 135, No. 2,
pp.021403–021403-7. Available at http://dx.doi.org/10.1115/1.4007958.
Fukuoka, T. and Xu, Q. (1999) ‘Evaluations of thermal contact resistance in an atmospheric
environment’, Transactions of the Japan Society of Mechanical Engineers Series A, Vol. 65,
No. 630, pp.248–253. Available at http://dx.doi.org/10.1299/kikaia.65.248.
Gmelin, E., Asen-Palmer, M., Reuther, M. and Villard, R. (1999) ‘Thermal boundary resistance of
mechanical contacts between solids at sub-ambient temperatures’, Journal of Physics D:
Applied Physics, Vol. 32, No. 6, pp.R19–R43.
Hoang, E., Lecrivain, M., Hlioui, S., Ben Ahmed, H. and Multon, B. (2010) ‘Element of slot
thermal modelling’, 2010 International Symposium on Power Electronics Electrical Drives
Automation and Motion (SPEEDAM), pp.303–305, 14–16 June, Available at http://dx.doi.org/
10.1109/SPEEDAM.2010.5544907.
Jungreuthmayer, C., Bauml, T., Winter, O., Ganchev, M., Kapeller, H., Haumer, A. and Kral, C.
(2011) ‘Heat and fluid flow analysis of an internal permanent magnet synchronous machine by
means of computational fluid dynamics’, 2011 IEEE International Electric Machines &
Drives Conference (IEMDC), pp.515–520, 15–18 May, Available at http://dx.doi.org/
10.1109/IEMDC.2011.5994651.
Kammoun, S., Doghri, I., Adam, L., Robert, G. and Delannay, L. (2011, December) ‘First pseudo-
grain failure model for inelastic composites with misaligned short fibers’, Composites Part A:
Applied Science and Manufacturing, Vol. 42, No. 12, pp.1892–1902. Available at
http://dx.doi.org/10.1016/j.compositesa.2011.08.013.
Kastensson, A. (2014) ‘Developing lightweight concepts in the automotive industry: taking on the
environmental challenge with the SåNätt project’, Journal of Cleaner Production, Vol. 66,
pp.337–346.
Kim, S.C., Kim, W. and Kim, M.S. (2013) ‘Cooling performance of 25 KW in-wheel motor for
electric vehicles’, International Journal of Automotive Technology, Vol. 14, No. 4,
pp.559–567.
Kral, C., Haumer, A. and Bauml, T. (2008, October) ‘Thermal model and behavior of a totally-
enclosed-water-cooled squirrel-cage induction machine for traction applications’, IEEE
Transactions on Industrial Electronics, Vol. 55, No. 10, pp.3555–3565. Available at
http://dx.doi.org/10.1109/TIE.2008.927242pp
Lewis, D.V. and Perkins, H.C. (1968, September) ‘Heat transfer at the interface of stainless steel
and aluminum—the influence of surface conditions on the directional effect’, International
Journal of Heat and Mass Transfer, Vol. 11, No. 9, pp.1371–1383. Available at
http://dx.doi.org/10.1016/0017-9310(68)90182-8.
Li, M., Wan, Y., Gao, Z., Xiong, G., Wang, X., Wan, C. and Luo, H. (2013, October) ‘Preparation
and properties of polyamide 6 thermal conductive composites reinforced with fibers’,
Materials & Design, Vol. 51, pp.257–261, ISSN 0261-3069. Available at
http://dx.doi.org/10.1016/j.matdes.2013.03.076.
Lohse-Busch, H. (2004) Thermal Overload Capabilities of an Electric Motor and Inverter Unit
Through Modeling Validated by Testing’, Unpublished MSc. Thesis, Virginia Polytechnic
Institute and State University, Blacksburg, USA.
Malumbres, J.A., Satrustegui, M., Elosegui, I., Ramos, J.C. and Martínez-Iturralde, M. (2015, 22
January) ‘Analysis of relevant aspects of thermal and hydraulic modeling of electric machines.
Application in an open self ventilated machine’, Applied Thermal Engineering, Vol. 75,
pp.277–288. ISSN 1359-4311. Available at http://dx.doi.org/10.1016/j.applthermaleng.
2014.10.012.
186 T. Müller et al.

Marotta, E.E. and Fletcher, L.S. (1996) ‘Thermal contact conductance of selected polymeric
materials’, Journal of Thermophysics and Heat Transfer, Vol. 10, No. 2, pp.334–342.
Mills, A.F. (1999) Heat Transfer, 2nd ed., Prentice Hall, Upper Saddle River, NJ.
Mori, T. and Tanaka, K. (1973) ‘Average stress in matrix and average elastic energy of materials
with misfitting inclusions’, Acta Metallurgica, Vol. 21, p.571.
Mura, T. (1982) Micromechanics of Defects in Solids, Martinus Nijhoff, The Hague.
Nguyen, B.N., Bapanapalli, S.K., Kunc, V., Phelps, J.H. and Tucker, C.L. (2009, February)
‘Prediction of the elastic–plastic stress/strain response for injection-molded long-fiber
thermoplastics’, Journal of Composite Materials, Vol. 43, No. 3, 217–246. Availabe at
http://dx.doi.org/10.1177/0021998308099219.
Osborne, J. (2013) ‘Automotive composites—in touch with lighter and more flexible solutions’,
Reinforced Plastics, Vol. 57, No. 2, pp.20–24.
Parmenter, K.E. and Marschall, E. (1995) ‘Influence of surface preparation on thermal contact
conductance of stainless steel and aluminum’, Experimental Heat Transfer, Vol. 8, No. 3,
pp.195–208. Available at http://dx.doi.org/10.1080/08916159508946501.
Staton, D.A., Boglietti, A. and Cavagnino, A. (2005) ‘Solving the more difficult aspects of electric
motor thermal analysis in small and medium size industrial induction motors’, IEEE
Transactions on Energy Conversion, Vol. 20, No. 3, pp.620–628.
Staton, D.A. and Cavagnino, A. (2008, October) ‘Convection heat transfer and flow calculations
suitable for electric machines thermal models’, IEEE Transactions on Industrial Electronics,
Vol. 55, No. 10, pp.3509–3516. Available at http://dx.doi.org/10.1109/TIE.2008.922604.
Staton, D.A. and Šusnijc, L. (2009, December) ‘Induction motors thermal analysis’, Journal for
Theory and Application in Mechanical Engineering, Vol. 51 No. 6, pp.623–631.
Stewart, R. (2011) ‘Rebounding automotive industry welcome news for FRP’, Reinforced Plastics,
Vol. 55, No. 1, pp.38–44.
Transport and Environment. (2011) ‘Consultation response: reducing CO2 emissions from road
vehicles’, Available at http://www.transportenvironment.org/whatwe-do/cars-and-co2
(accessed 16 February 2015).

You might also like