Experimentally Validated Numerical Model of Thermal and Flow Processes Within The Permanent Magnet Brushless Direct Current Motor

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

International Journal of Thermal Sciences 130 (2018) 406–415

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


journal homepage: www.elsevier.com/locate/ijts

Experimentally validated numerical model of thermal and flow processes T


within the permanent magnet brushless direct current motor
B. Melkaa,∗, J. Smolkaa, J. Hetmanczykb, Z. Bulinskia, D. Makielab, A. Ryfaa
a
Silesian University of Technology, Institute of Thermal Technology, Konarskiego 22, Gliwice, 44-100, Poland
b
Silesian University of Technology, Department of Power Electronics, Electrical Drives and Robotics, B. Krzywoustego 2, Gliwice, 44-100, Poland

A R T I C LE I N FO A B S T R A C T

Keywords: In this paper, a numerical model describing the heat transfer and air flow inside and outside the casing of the
Electric motor permanent magnet brushless direct current (PM BLDC) motor is presented. In the model, a conjugate heat
Permanent magnet brushless DC motor transfer including the heat conduction with heat sources in windings, magnets, bearings, natural and forced
Thermal analysis convection and radiation phenomena within internal and external air domain were analysed. The complex
CFD
geometrical model included windings, a magnet circuit, a rotor with neodymium magnets, a PCB electronic
Power loss
plate, plastic covers, internal air between solid parts as well as a portion of the external air. The validation
LDA
process of the developed model was realised on the basis of the velocity and temperature fields. The experi-
mental data were recorded by 7 constant temperature anemometers at 4 levels above the motor housing, Laser
Doppler Anemometry sensors in two axes inside the rear portion of the motor and 25 thermocouples fixed inside
and outside the motor. The experimental and numerical tests were performed for 4 motor loadings and 3 ro-
tational speeds. The average temperature error for the internal point was of 9 K, while that of the external points
was of 2 K. The velocity field was very well predicted close to the housing wall in terms of horizontal and vertical
components. A slightly worse agreement was found for the vertical component, especially in the vicinity of the
shaft. The results were consistent for all the considered motor loadings and rotational speeds.

1. Introduction with a mechanical commutator requires a special maintenance monitor


and requires systematic brush replacements. Moreover, the efficiency of
Environmental, political and industrial tendencies have directed a brush motors decreases with additional losses of the mechanical com-
reduction of non-renewable sources of energy consumption [1]. One mutation and higher electrical contact resistance [10]. The brushless
way to achieve this goal is to increase the use of electrical power drives solution is free of the above-mentioned disadvantages [11]. A perma-
in vehicles and other machines to replace internal combustion engines nent magnet brushless motor powered by direct current (PM BLDC
of lower efficiency. Hence, a direct current electric motor is expected motor) is characterised by high-efficiency, high-power density and lack
have wider application as a direct source of drive [2]. Today, the pro- of mechanical commutation [12]. The main disadvantages of this ma-
blem with implementing this concept, on the large scale, is the lim- chine are the high market price of the motor with an electronic com-
itation of electric power storage in vehicles and small systems. Never- mutator and the electronic commutation complexity. Nonetheless, they
theless, the prospective for electric power storage technology and its are becoming ever more popular due to the mentioned advantages.
improved maintenance is promising [3–6]. An additional challenge to Therefore, this paper focuses on this type of motor.
solve with increased electric vehicle usage is the integration of charging In electric motor design, thermal analysis plays an important role
stations into the electric grid. Nevertheless, many studies and proposed [13]. It helps estimate the value and location of the maximum tem-
solutions can be found [7,8]. Therefore, in the near future, increased perature in the machine. The specific material and components of the
use of small power electric motors powered by direct currents can be motor should be protected from exceeding the maximum safe working
expected. temperature. Hence, proper estimation of the maximum temperatures
The most popular technologies for motors powered by direct current occurring during exploitation could help reduce the risk of overheating
are brush or brushless solutions. Brush motors are cheaper to produce and failure of the machine [14]. A precise thermal analysis allows for
and their construction is simpler [9]. Nevertheless, the brush solution decreasing the machine size by reducing the safety margin and,


Corresponding author.
E-mail address: Bartlomiej.Melka@polsl.pl (B. Melka).
URL: http://www.itc.polsl.pl/bmelka (B. Melka).

https://doi.org/10.1016/j.ijthermalsci.2018.04.029
Received 11 October 2017; Received in revised form 20 April 2018; Accepted 20 April 2018
Available online 26 May 2018
1290-0729/ © 2018 Elsevier Masson SAS. All rights reserved.
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

consequently, decreasing the material costs. The optimisation of the presented in Ref. [34]. The authors obtained satisfactory agreement
thermal motor behaviour could also minimise the copper losses of the between the model and experimental results. They suggested the RNG
machine by decreasing the electrical resistivity of the windings de- k− ε model for turbulence modelling in the heat dissipation process via
pending on the temperature. Due to these facts, thermal analysis should natural convection.
always be included in the electric motor design process. The main aim of the presented paper was to determine the hot-spot
One of the simplest and the most popular models to predict the temperature and flow fields in terms of a safe and long-term operation
temperature distribution in electric machines is the lumped parameter of the analysed low-power PM BLDC motor. To achieve this goal, the
model (LPM) [15]. LPM in the thermal analysis of electric machines is temperature and the velocity fields were determined within and around
also known as the thermal network [16]. Complex thermal networks the machine housing under various loadings and rotational speeds. The
contain dozens of the elements - nodes [17]. Usually, LPM allows the fields were computed on the basis of the developed thermal model that
temperature of the entire element to be calculated, e.g., one tempera- included the computational domain of a very complex internal struc-
ture for winding, when it is based on the zero-dimensional model ture of the considered motor. In addition, the block of the surrounding
without spatial resolution. Moreover, it does not contain precise in- ambient air was formulated.
formation about the coolant flow. The essential parameters used in the To validate the model, a test rig was built to measure temperature
thermal networks, e.g., convective heat transfer coefficients, are as- and velocity fields at specified locations inside and outside the motor
sumed or calculated usually from the empirical equations, which are housing. A set of calibrated thermocouples was used to record the va-
presented in the literature for the general simplified shapes and geo- lues at the stator magnetic core, housing and in the vicinity of the
metries. These equations were derived from the heat transfer field windings. Moreover, the temperatures were also captured at a number
analysis, e.g., in Ref. [18]. The main advantage of thermal networks is of locations above the machine housing. The velocity field was mea-
their sufficient accuracy with short computational times. Even simple sured by means of the Laser Doppler Anemometry (LDA) technique in
thermal networks produce reasonable temperature response for short- the rear part of the motor. The velocity validation inside the motor
time thermal transient [19]. It is possible, especially for the machines housing and around the motor is a challenging task. Therefore, this
with high thermal inertia, while the thermal capacitance of applied problem is rarely presented in the literature. In the presented study, the
materials was correctly assumed. The temperature prediction in the LDA technique was used showing its strengths and weaknesses in this
electric motors with permanent magnets can also be realised by means application. In addition, the anemometers were used to find the velocity
of numerical simulation and on-line measurements together [20]. fields above the motor at four levels. All the mentioned parameters
A more complex and accurate method for estimation of the tem- were measured for the motor working at four loading conditions at
perature distribution is Computational Fluid Dynamics (CFD). In this various rotational speeds reaching the maximum value of 3500 rpm.
method, many characteristic parameters, such as the heat transfer Namely, the considered PM BLDC motor was experimentally tested at
coefficient, can be calculated directly and locally within a selected 11 operating conditions.
computational domain. This is possible by extending the model geo- A temperature comparison of the solid motor parts shows very good
metry to space outside the machine and, consequently, by considering agreement between the numerical and experimental results. Similarly,
the heat generated inside the motor and then dissipated directly to the the velocity profile inside and the field outside the housing was pre-
ambient air [21,22]. As a result, a more detailed analysis leads to more dicted accurately for most measurement points. The study also shows
accurate heat dissipation predictions. In the literature, there are many some problems in the velocity predictions at some near-wall locations.
studies that compare CFD and LPM with respect to experimental data
[23]. One of the most important works about thermal modelling of the 2. Test rig
electric motor operation describes the CFD model formulation and the
usage of the CFD field results in the LPM analysis [24]. In the literature, 2.1. Motor description and test rig components
there are also studies based on the Finite Element Method, which is
usually applied when researchers focus only on the thermal conduction The studied machine was a low-power electric motor with perma-
in solids. One of the studies with thermal conduction analysis in the nent magnets on the rotor. The motor was supplied by the electronic
motor housing is the study of [25]. In the cited work, the authors as- commutator transforming direct current. The hall effect sensor was
sumed heat flux from the internal surface of the housing frame and used to control the rotor position in the control process. The rated
analytically calculated the heat transfer coefficient. Consequently, the parameters of the rotor were as follows: the output power of 431 W, the
estimated temperature field of the finned housing was compared with input voltage of 24 V and the current of 21 A.
the experimental study conducted. Zhang et al. [26] presented a cou- The test rig consisted of two PM BLDC motors connected to each
pled analysis between the electromagnetic and thermal solvers. How- other by a universal coupling. The first motor worked in the motor
ever, the external heat transfer coefficient was calculated analytically, mode, while the second one worked in the generator mode. The ma-
while radiation between internal elements was assumed to be negli- chines were mounted with aluminium brackets and the natural rubber
gible. An interesting work was presented by Jungreuthmayer et al. distances to the aluminium base. The distances were used to reduce
[27], where thermal analysis was conducted for a water-cooled electric heat conduction towards the base. The test rig was covered by an ac-
motor. Oil spray cooling of internal parts of electric motor was in- rylic glass cover. The cover was applied to reduce velocity fluctuations
vestigated experimentally in Ref. [28] and also numerically in Ref. caused by the laboratory space. The cover was a square-shaped channel
[29]. Heat storage mechanism allowing cooling intensification of the open from the bottom and closed at the top. In addition, an 8 mm hole
transient mode of electric motor using phase change material was in- was drilled through the top wall for natural air flow around both mo-
vestigated in Ref. [30]. The multi-physics simulation connecting tors. The above-mentioned elements were the main parts of the test rig
thermal analysis with mechanical analysis was presented for case of and all of them were taken into consideration for the formulated CFD
aircraft actuator in Ref. [31]. model. The picture of the test rig is presented in Fig. 1 with schema-
Nevertheless, studies showing the flow field inside and outside the tically marked boundary conditions set in the model. The motor con-
motor casing are not prevalent in the literature. The velocity mea- struction with measuring probes positions were schematically pre-
surements were made in a water-filled axial machine in the work of sented in Fig. 2.
Aubert et al. [32]. Preliminary results of the air velocity validation were The remaining elements used in the test rig but not included in the
also described in our previous work [33]. Moreover, the velocity vali- model were the main power supply unit, an electronic commutator
dation caused by natural convection from a small heat source using consisting of mosfet transistors, a control system for the PM BLDC
Particle Image Velocimetry technique in the closed cavity was motor, and system of resistors used to dissipate electric power produced

407
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

and numbers coloured red. Thermocouples from #T12 to #T17 were


mounted on the external part of the motor housing. Thermocouples
#T18 - #T21 were mounted above the analysed motors at heights of 15,
25, 38 and 48 cm from the housing wall, respectively.
The vertical component of the velocity vector outside the motor was
measured by seven constant temperature anemometers at four height
levels. The positions of these sensors are presented in Fig. 2 as green
square markers with prefixes A and numbers in green colour. The
horizontal distance between anemometers was approximately 0.04 m,
while vertical distance was at the level of 0.047 m.
The velocities within the motor housing in its rear part were mea-
sured using a LDA system from Dantec Dynamics. The LDA system was
used to measure velocity components in two directions: horizontal and
vertical. The velocity measurements within the housing were performed
in two sections.
The first section was on the height of shaft axis. This section con-
tained the distance from the housing wall to the shaft wall. Its height is
marked by a blue triangle with the label of LDA1 in Fig. 2. The second
section, where the velocity measurements were also conducted, in-
cluded the distance from the wall to the centre of the motor. The height
of this section is marked by a blue triangle with the label of LDA2 in
Fig. 2. The distance between each recorded point was 1 mm.
The electric measurements were carried out using a Sanwa PC5000a
digital multimeter, Chauvin Arnoux E3N current probes and a
Tektronix MSO 3014 oscilloscope.
Fig. 1. Test rig with schematically marked boundary conditions.
2.3. Measurement procedure and motor losses calculation
by the generator. These devices were connected by wires and con-
nectors. The temperature, velocity and electrical measurements were exe-
cuted for three rotational speeds at four different loads. The load was
modified by changing a connection of four resistor sets located on the
2.2. Measuring devices generator side. This allowed the examination of 11 different operational
states presented in Table 1. In the table, the calculated losses are also
Temperature was measured using 22 calibrated T-type thermo- presented. The copper losses were calculated according to the multi-
couples. The thermocouple calibration was performed using a Fluke plication of square current in A and the winding resistance in Ω. The
9100S unit with an accuracy of ± 0.25 °C. The first six thermocouples resistance of the motor windings was estimated using the DC mea-
were fixed in the region of the motor windings. Thermocouple #T7 was surement technique. The windings resistance was determined each time
mounted between the stator magnetic core and the aluminium motor when the steady state conditions were obtained in the experimental
housing. Thermocouples from #T8 to #T11 were fixed within the rear test. The resulting volumetric heat sources were obtained based on the
portion of the motor. The positions of thermocouples mounted inside Joule power losses equation (see Eq. (12)). In this way, the tempera-
the motor are schematically presented in Fig. 2 as dots with prefixes T ture-dependent resistance and the resulting power losses were

Fig. 2. Simplified scheme of the studied motor with thermocouple and anemometer positions.

408
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

Table 1 presented in Fig. 3 on a cross-section built on the surface rotated at an


Operational points tested on the test rig. angle of 45° in the domain centre. It is presented in the figure that most
Var. No Res. of the Rot. Av. Copper Iron Mech. of the motor components were discretised using structured HEX ele-
generator speed, Current losses, W losses, losses, ments. The same type of the grid was employed in the gap between the
circuit, Ω RPM per phase, W W stator and the rotor. Due to complex and irregular shape of the front
A and rear air region, the TET element mesh was used in those zones.
I 0.25 1415.1 3.41 10.22 3.98 0.88
Outside the motor housing, in the machine vicinity the TET mesh was
II 0.25 3489.0 14.37 43.10 13.22 2.94 also used, while in the further region the HEX elements dominated
III 0.33 1415.1 2.31 6.92 3.98 0.88 (region not presented in Fig. 3). In the meshing strategy, the size of
IV 0.33 2500.5 6.25 18.74 8.33 1.85 elements was gradually increased. The worst element characterised by
V 0.33 3571.5 10.53 31.59 13.67 3.04
the smallest orthogonal quality was on the level of 0.07 and was located
VI 0.5 1415.1 1.00 2.99 3.98 0.88
VII 0.5 2500.5 3.45 10.35 8.33 1.85 in some cell in the air outside the machine in the TET zone. Orthogonal
VIII 0.5 3571.5 6.32 18.95 13.67 3.04 quality of the rest of elements was higher than 0.2 and in the most
IX 1 1388.9 0.46 1.39 3.89 0.86 regions did not fell below 0.5. The minimum value of the Y+ parameter
X 1 2500.5 1.19 3.58 8.33 1.85
in the numerical domain was at the level of 0.09 located within motor
XI 1 3489.0 2.26 6.78 13.22 2.94
in its rear part and the maximum value of Y+ reached 14.6 located in
the air outside the machine.
estimated. In Fig. 3, the global coordinate system is presented. The gravita-
The core losses, also known as the iron losses, were calculated from tional body force was set in the opposite direction of the x-axis of this
the characteristics collected during the idle state measurements. They system.
were dependent on the rotational speed. The mechanical losses were
first included in the characteristics. However, in the next step, they 3.1. Governing equations
were separated. The mechanical losses were calculated using metho-
dology presented by the bearing manufacturer. All the measurements The CFD model was built on the basis of the governing equations
were performed until a thermal steady state had been reached. describing the mass, momentum and energy conservation principles
that were expressed by Eq. (1), through Eq. (5). A motion of the motor
rotor was modelled in accordance with the Multiple Reference Frame
3. Numerical model approach [35].
∂ρ
To simulate the phenomena occurring during the experiment, which + ∇⋅(ρ→
v)=0
∂t (1)
was executed on the test rig, a numerical model was proposed. The
numerical mesh generated within the computational domain included where t is the time, ρ is the density defined as a function of temperature
more than 8 million elements. The visualisation of the mesh is in the fluid zone and therefore air is treated as ideal gas and → v is the

Fig. 3. Mesh displayed in the centred vertical cross-section.

409
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

velocity vector. coefficient, n is the refractive index, σs is the scattering coefficient, σ is


∂ → the Stefan-Boltzmann constant of 5.669⋅10−8 W⋅m−2⋅K−4 , Φ is the phase
(ρ v ) + ∇⋅(ρ→→ →
v v ) = −∇p + ∇ (τ ) + ρg function and Ω′ is the solid angle.
∂t (2)
where p is the pressure, τ is the stress tensor that can be calculated from 3.2. Source terms

Eq. (3) and g is the gravitational acceleration vector.
2 The power losses in the motor were implemented as a homogenous
τ = μ ⎡ (∇→
v + ∇→
v ) − ∇⋅→
T
v ⎤ volumetric heat source in the case of copper and iron losses. These heat
⎣ 3 ⎦ (3)
sources were set in the volume of windings, laminated core and bear-
where μ is the dynamic viscosity and  is the unit tensor. ings. The heat source values were set individually for the load and the
∂ specific speed. The heat sources were calculated on the basis of the
(ρh) + ∇⋅(→
v ρh) = ∇⋅(kij ∇T ) + Sh
∂t (4) electrical measurements. The heat sources in the winding were esti-
mated as the copper losses from Eq. (12).
where h is the specific enthalpy, kij is the thermal conductivity in the
specific direction due to the anisotropic conductivity of the windings I 2⋅R (T )
Sh =
and core and Sh is the volumetric heat source. The anisotropic thermal VCu (12)
conductivity and the heat sources are described in the following sec-
where I is the electric current, R (T ) is temperature-dependent re-
tions.
sistance and VCu is the windings volume.
∂ In the model, the mechanical losses were set on the external bearing
(ρE ) + ∇⋅(→
v (ρE + p)) = ∇⋅(keff ∇T + (τeff ⋅→
v ))
∂t (5) surfaces where balls were in contact with bearing walls.
where the total energy E can be calculated using Eq. (6). Moreover, in The estimated copper, iron and mechanical loss values are presented
the former equation, keff is a sum of k and kt , where k is the laminar for every studied load in Table 1.
thermal conductivity and kt is the turbulent thermal conductivity.
Analogically, the stress tensor τeff is calculated. 3.3. Material properties

p →
v
2
In the steady state, which was assumed in the model, only thermal
E=h− +
ρ 2 (6) conductivity influences the results obtained. In the electric motor, most
of the generated heat is released in the materials which have aniso-
Moreover, the standard k-ε model for turbulence was employed in
tropic thermal conductivity. It directly influences the temperature field.
the solver in a form expressed by Eqs. (7) and (8).
Therefore, anisotropic thermal conductivity was set for the windings
∂ ∂ ∂ ⎡⎛ μ ∂k and laminated core. The following thermal conductivities were as-
(ρktur ) + (ρktur →
vi ) = ⎜ μ + t ⎞ tur ⎤ + Gk + Gb − ρε − YM

sumed in the model:
∂t ∂x i ∂x j ⎢
⎣⎝ σ ⎥
k ⎠ ∂x j ⎦

+ Sk (7)
• windings in the axial direction (along the wires) 167.3 W⋅m ⋅K −1 −1

[36];
• windings in the radial direction 1.4 W⋅m ⋅K [36];
∂ ∂ ∂ ⎡⎛ μ ∂ε ⎤ ε
(ρε ) + (ρε→
vi ) = ⎜ μ + t⎞ ⎟ + C1ε (Gk + C3ε Gb) −1 −1
∂x j ⎢ ⎥
∂t ∂x i ⎣ ⎝ σε ⎠ ∂x j ⎦ k tur
• the laminated core in the radial direction 22.2 W⋅m ⋅K [36]; −1 −1

− C2ε ρ
ε2
+ Sε • the laminated core in the axial direction 4.9 W⋅m ⋅K [36];−1 −1

ktur (8) • aluminium housing (the most popular aluminium alloy 6061) 167
−1 −1
W⋅m ⋅K ;
where ktur is the turbulence kinetic energy, ε is the turbulent dissipation
rate, Gk is the generation of turbulence kinetic energy due to the mean • electronic plate in a form of the Printed Circuit Board (PCB) 0.55
W⋅m−1⋅K−1 [37].
velocity gradients, Gb is the generation of turbulence energy due to
buoyancy, Ym is the contribution of the fluctuating dilatation in com-
The air properties were defined as the temperature-dependent
pressible turbulence to the overall dissipation rate, C1ε , C2ε , Cμ , σk , σε are
properties including the density that was defined according to the in-
the model constants of 1.44, 1.92, 0.09, 1.0, 1.3, respectively. The
compressible ideal gas equation.
turbulent viscosity μt was calculated using Eq. (9). The constant C3ε was
Thermal contact resistance was added in the crucial parts where it
calculated by Eq. (10).
could play an important role. The thermal contact resistance was de-
k2 fined at the interface between windings and plastic layer which covered
μt = ρCμ
ε (9) the iron core. In this case, the thickness of the equivalent material was
→ of 8.5e-5 m. The thermal conductivity of the equivalent contact re-
v sistance material was assumed on the level of 0.0242 W⋅m−1⋅K−1 and its
C3ε = tanh →
v⊥ (10) value was close to the air properties.
→ →
where v and v⊥ are components of the velocity vector parallel and
3.4. Boundary conditions
perpendicular to the gravity direction, respectively.
The heat transfer by radiation was modelled using Discrete
To solve the governing equations presented in Section 3.1, the
Ordinates (DO) [35]. In the presented model, the angle division was set
thermal and flow boundary conditions were defined on the domain
to 4 for each octant of space, while the pixilation of the control angle
boundaries. The constant-pressure inlet located at the bottom surfaces
was defined as 3. The DO model can be expressed by Eq. (11).
under the plastic cover, where the air could freely be exchanged with
σT 4 the laboratory room, was set equal to 0 Pa. This boundary condition
∇⋅(IR (→
r,→
s )→
s ) + (a + σs ) IR (→
r,→
s ) = an2
π was schematically marked by blue arrows in Fig. 1. The constant-
σ 4π
+ ∫ IR (→
r,→
s ′)Φ(→
s,→
s ′) dΩ′ pressure outlet was set on the top wall where an 8 mm hole was drilled.
4π 0 (11)
Pressure outlet was presented in Fig. 1 by orange arrow. On the vertical

where IR is the radiation intensity depending on the position r and and top walls, the ambient air temperature and local heat transfer
direction →
s vectors. In the former equation, a is the absorption coefficient (HTC) were calculated based on the empirical equations for

410
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

vertical and horizontal plates in natural convection [18]. One of the


walls, where this boundary condition was set, was marked by green
lines in Fig. 1. The average HTC value for the horizontal external walls
of the plastic cover was calculated and reached approximately 3.4
W⋅m−2⋅K−1 and 2.62 W⋅m−1⋅K−1 for vertical walls in Variant V. The
emissivity of the plastic cover was set equal to 0.95. The rotational
speed was assigned to walls which were moving according to the data
listed in Table 1.
The internal parts of the motor working as the generator (the ma-
chine on the right-hand side in Fig. 2) were not included in the model to
reduce the number of elements in the computational domain. The
measured temperature collected from the generator housing were used
to build the temperature profile and implemented in the model as a
boundary condition.

4. CFD results and validation

The following results are presented for the temperature and velocity
fields. First, the temperature field in a vertical cross-section in the
centre of the domain is illustrated. Next, a comparison of measurements Fig. 5. The temperature comparison between the CFD and experimental results
and CFD results at the specific points inside and outside the motor is at the load of 3 resistors and the rotational speed of 3500 rpm - Variant V.
discussed. Then, the velocity measurements and CFD results above the
studied machines are presented. Finally, the velocity measurements temperature in this view was higher directly above the end caps of both
collected inside the motor are compared with the CFD results calculated machines. The highest air temperature outside the motor occurred
for some turbulence models. The following results are presented for above the back-end cap of the generator. The remaining temperature
Variants IV and V, but the calculations were executed for all the var- field located 0.1 m above the machines was practically uniform and was
iants described in Table 1. The two variants, presented in the results present at half of the domain height.
section, vary between each other by the rotational velocity and the A comparison of the temperature fields obtained using CFD and
losses level. experimental measurements is presented in Fig. 5. The results collected
at the specific locations are presented in four groups, which are sepa-
4.1. Temperature field inside and outside the motor housing rated by the dotted green lines. Those locations denoted as red coloured
points are schematically shown in Fig. 2:
The temperature field of the load described as Variant V is presented
in Fig. 4 in the middle vertical cross-section of the computational do- • inside at the front part of the motor #T1 -#T7;
main. In the figure, the motor is on the left-hand side, and the generator • inside motor and in the rear part of the motor #T8 -#T11;
is on the right-hand side. • on the external part of the motor housing #T12-#T19;
The highest temperature, approximately 78 °C, occurred in the • air above the motor #T20 -#T24.
windings region. The presented cross-section does not cut the model
directly through the windings. The external shape of the machines are In that figure, the thermocouples fixed on the windings recorded the
visible in the presented temperature field because the surrounding air highest temperature values between 80 and 85 °C. The last thermo-
temperature was notably lower. Moreover, it is observed that the air couple in the first group shows a value of 71.4 °C. The CFD results
present more uniform temperature distribution in the first group of
thermocouples and reach approximately 84 °C. Thermocouples #T8 and
#T9 fixed in the back part of the motor show an internal air tem-
perature of 68 and 70 °C, respectively. One of largest differences be-
tween measurements and CFD results occurs in those points. In parti-
cular, the CFD results are underestimated by approximately 5 K. Point
#T10 fixed inside the motor on the housing reached 65 °C. The last
measured point #T11 located on the PCB recorded a temperature of
84 °C. This temperature was similar to the temperature occurring on the
windings. It was caused by the thermocouple directly fixed where one
of the phase electric paths was located. The current flows through that
path caused a local heat generation. The next seven points, where the
temperature was measured, were attached to the external surface of the
housing. The temperature for these points was approximately 65 °C.
Slightly higher temperatures were observed in the middle of the motor
length. The maximum difference between these temperatures was 1 K.
Hence, temperature distribution was uniform on the housing walls
mainly due to the high thermal conductivity. The last four temperatures
were measured for different heights of the air above the considered
motor. The temperature of these points was at a level of 35 °C. The
differences between the measurements and CFD results at these points
Fig. 4. Temperature field in °C displayed in the domain cross-section through are in the range of 1 K.
the motor axis at the load of 3 resistors and the rotational speed 3500 rpm The average heat transfer coefficient obtained on the basis of the
(Variant V). average convective heat flux, the outer casing wall temperature and the

411
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

value of the vertical velocity component from all the measured points
was 0.089 m⋅s−1, while the average of these values from the thermal
model was 0.131 m⋅s−1. In the figure, most of the CFD results are in the
error bar ranges estimated from the measurements. Considering that
calculations were conducted in a steady state mode, the agreement
between model and measurement can be treated as satisfactory.

4.3. Velocity field inside the motor housing

A reduced model was created to validate the velocity calculation.


This model was limited to the domain of air in the back part of the
motor, where the LDA apparatus was used to record velocity compo-
nents as described in Section 2.2. For this model, three numerical me-
shes were tested during the calculation process that contained between
0.7 and 3 million elements. The temperature profiles were implemented
for all investigated loads from the full model described in the previous
sections.
First, a comparison was made between the horizontal component of
the velocity vector for the CFD results and the measurements in section
Fig. 6. Vertical velocity component comparison between CFD and experimental
LDA1 as shown in Fig. 7. The first line marked as continuous in red
results at the load of 3 resistors and a rotational speed of 3500 rpm (Variant V).
colour represents a position where the LDA measurements were per-
formed. In this figure, the second line represents the velocity in the
average air temperature was in the range of 4–6 W⋅m−2⋅K−1. These section moved 5 mm to the front of the motor. A comparison of the two
values are consistent with the heat transfer coefficient obtained from positions was used to visualise the velocity field in the analysed region.
the Nusselt number correlation for the natural convection flow around a The results of this configuration are presented for Variants IV and V. For
cylindrical shape [18]. both presented loads, the velocity obtained by the model and the
measurements is negative in the region near the wall. The first value
4.2. Velocity field outside the motor housing was measured at 0.001 m from the wall. That means the horizontal
velocity component is directed to the front of the motor in this region.
A comparison of the velocity results from CFD and measurements, According to the measurements, the direction of the horizontal velocity
collected outside the motor, is presented in Fig. 6. The chart describes component changes at approximately 1 cm from the wall. The same
the vertical component of the velocity vector in 28 points measured by tendency can be observed in the model results. Moreover, a rapid in-
7 anemometers at 4 height levels above both the motor and generator. crease of the horizontal velocity in the direction of the motor front at a
The values were grouped according to these levels with the group de- distance of approximately 1 cm from the wall can be observed. The
scription above the chart. The sensor numbering at the bottom of the experimental results show a sudden decrease in the direct shaft vicinity.
chart is consistent with the numbers with prefixes A in Fig. 2. The error At this location, the CFD results reveal the changes between the two
bars were added to the measured values based on the standard devia- presented lines. The second line moved 5 mm to the front of the motor
tion analysis. As shown in this figure, the error bars are wide for many shows better agreement with the measured values than the first line
points. This can be explained by the natural convection flow that recorded in the middle of the back part of the motor.
caused high instability of the measurement. In the case of the external In Fig. 8, the vertical component of velocity on the height of the
air flow, that was the main mechanism of the heat dissipation from the shaft axis (LDA1) is compared between results from the experiment and
motor walls. the model for Variants IV and V. In this section, the vertical velocity
According to Fig. 6, in the first group on the lowest presented level, increases slightly from wall vicinity to shaft direction approaching the
the CFD model results were in the range of the estimated velocity errors end of the recorded range. In the distance behind recorded range from
for 5 out of 7 measured points. In the next group, there were only 3 measurements, i.e., more than 0.022 m, the values derived from the
values within the error ranges. In the third group, 4 of the measured CFD model increased rapidly to achieve the linear velocity of the shaft.
points were consistent with the modelled values. Finally, the highest The values in the measured range are similar between experiment and
level contained 6 points in which the CFD results were in the range of CFD results. The measured values of velocity were limited by the re-
the standard deviations calculated from the measurements. The average solution (of 1 mm) in which each measured point could be tested. The

Fig. 7. Horizontal component of the velocity vector within motor on a height of the shaft axis (LDA1) for Variant No. V (left) and Variant IV (right).

412
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

Fig. 8. Vertical component of the velocity vector within the motor on a height of the shaft axis (LDA1) for Variant V (left) and Variant IV (right).

highest expected vertical velocity should be reached in the wall vicinity increasing tendency in the same section, but the values are approxi-
as it was observed in the CFD results. Therefore, it seems that the values mately two times greater than the measured values. At distances farther
recorded during the experiment in those locations are probably wrong than 0.02 m from the wall, this velocity increased again achieving ap-
(the last non-zero measured value). The LDA device could reflect the proximately 0.25 m⋅s−1 . In this region, the velocity from the experi-
laser beam or did not cross laser beam in the region where the seed ment and the model are similar. In the vicinity close to the shaft, the
(smoke) occurred. Therefore, the maximum velocity value could occur vertical velocity component significantly decreases in the CFD results.
in the distance between the measured points. In Fig. 8, the two lines For Variant IV, the behaviour of this velocity component is similar to
obtained from the CFD model were also presented. The differences that described with the higher load. Nevertheless, the values from the
between the lines are negligible for the vertical velocity component CFD results are approximately three times greater than the values from
along the height of the shaft axis. the experiments up to a distance of 0.02 m from the wall. This differ-
In Fig. 9, the horizontal component of the velocity vector is pre- ence decreased by less than twice. In Fig. 10, the second line obtained
sented on the line directly above the shaft (LDA2). Similar to the pre- from the CFD results shows slightly smaller values than those from the
vious figures, the CFD results were presented for the two mapped lines. first line and values more similar to the experimental measurements.
In the figure, the CFD results are consistent with the measured values
for the entire range of investigated section. Firstly, near the wall, the 5. Conclusions
horizontal velocity was directed in the front of the motor with negative
values. The maximum value in this direction reached approximately 0.1 In this paper, a validated numerical model of heat and flow phe-
m⋅s−1. Similar to the previously described section, the horizontal velo- nomena in a PM BLDC motor was presented. Experiments were per-
city changed its direction at a distance of approximately 0.01 m towards formed for 11 loadings at a thermal steady state. The experiments in-
the back of the motor (positive values). From this point, the measured cluded the following: temperature measurements with thermocouples
values began to gradually increase. The CFD results for load Variant V inside and outside the motor, velocity measurements above the motor
were more in agreement with the measured range than in Variant IV, and generator with constant temperature anemometers, velocity mea-
especially in the part with negative values of the horizontal velocity. surements inside the rear part of the motor using Laser Doppler
Nevertheless, the CFD results of the horizontal velocity component Anemometry, and electrical measurements by current probes, a digital
reasonably reflected the experiment in section LDA2 to a distance of multimeter and oscilloscope.
0.02 m from the wall. The first line of the CFD results still increased but The numerical model presented in this paper was based on the
as quickly as the measured values. In this case, the second line behind complex geometry, which realistically duplicated the real machine. The
the distance of 0.02 m from the wall began decreasing and reached compared points where the temperatures were measured showed rea-
approximately 0 m⋅s−1 in the shaft vicinity. sonable agreement with the experimental data. A very high accuracy of
In Fig. 10, the vertical component of velocity on the line above the the temperature prediction was obtained for the ambient air and the
shaft (LDA2) is presented. For Variant V, the measured velocity values motor housing. In those regions, the maximum discrepancy did not
increased from 0 m⋅s−1 in the vicinity of the wall to approximately 0.2 exceed 3 K. Inside the housing, the numerical and experimental tem-
m⋅s−1 at a distance of 0.02 m from the wall. The CFD results showed an peratures were also accurately simulated. One of the main reasons was

Fig. 9. Horizontal component of the velocity vector within the motor on a height above the shaft axis (LDA2) for load Variant V (left) and Variant IV (right).

413
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

Fig. 10. Vertical component of the velocity vector within the motor on a height above the shaft axis (LDA2) for Variant V (left) and Variant IV (right).

the temperature-dependence of the winding power losses that were mobility, J Power Sources 256 (2014) 110–124, http://dx.doi.org/10.1016/j.
based on the experimental values for the electric resistance. The largest jpowsour.2014.01.085.
[6] T.-F. Yi, S.-Y. Yang, Y. Xie, Recent advances of Li 4 Ti 5 O 12 as a promising next
difference between the computed results and measured results as re- generation anode material for high power lithium-ion batteries, J Mater Chem A 3
gistered in the region of the stator core was approximately 9 K. The (11) (2015) 5750–5777, http://dx.doi.org/10.1039/C4TA06882C http://pubs.rsc.
reason for the mentioned discrepancies could be caused by the assumed org/en/content/articlehtml/2015/ta/c4ta06882c.
[7] H. Lund, W. Kempton, Integration of renewable energy into the transport and
values of the anisotropic thermal conductivities taken from the litera- electricity sectors through V2G, Energy Pol 36 (9) (2008) 3578–3587, http://dx.
ture. The second reason of the result difference could be caused by the doi.org/10.1016/j.enpol.2008.06.007.
temperature-dependent, but still homogeneous power losses defined for [8] Lopes JAP, Soares FJ, Almeida PMR. Integration of electric vehicles in the electric
power system. Proc IEEE 99(1). https://doi.org/10.1109/JPROC.2010.2066250.
the windings and the core. In the further research, the two-way coupled [9] X.D. Xue, K.W.E. Cheng, N.C. Cheung, Selection of electric motor drives for electric
model of thermal and electromagnetic processes is planned. vehicles, Power engineering conference, 2008. AUPEC '08. Australasian
The validation of the velocity results inside the motor was one of the Universities, 2008, pp. 1–6.
[10] K. Hameyer, R.J.M. Belmans, Permanent magnet excited brushed DC motors, IEEE
novelties presented in the paper. The agreement of the velocity field
Trans Ind Electron 43 (2) (1996) 247–255, http://dx.doi.org/10.1109/41.491348.
above the machines was also reasonable because the computed values [11] K.T. Chau, C.C. Chan, C. Liu, Overview of permanent-magnet brushless drives for
were in the error bar ranges for the majority of the compared points. electric and hybrid electric vehicles, IEEE Trans Ind Electron 55 (6) (2008)
The velocity results obtained with the CFD model inside the motor 2246–2257, http://dx.doi.org/10.1109/TIE.2008.918403.
[12] E. a. Lomonova, E. Kazmin, Y. Tang, J.J. Paulides, In-wheel PM motor: compromise
showed a similar tendency as in the experiment. Those observations are between high power density and extended speed capability, COMPEL Int J Comput
especially justified for the horizontal component of the velocity vector. Math Electr Electron Eng 30 (1) (2011) 98–116, http://dx.doi.org/10.1108/
In that case, the predicted values were close to those found in the ex- 03321641111091467.
[13] D. Staton, A. Boglietti, A. Cavagnino, Solving the more difficult aspects of electric
perimental tests along the whole path from the wall to the shaft. In the motor thermal analysis in small and medium size industrial induction motors, IEEE
case of the vertical component of the velocity vector, the numerical Trans Energy Convers 20 (2005) 620–628, http://dx.doi.org/10.1109/TEC.2005.
results were very close to the measured values in the central region of 847979.
[14] N. Putra, B. Ariantara, Electric motor thermal management system using L-shaped
the investigated region. In the wall and the shaft region, higher dis- flat heat pipes, Appl Therm Eng 126 (2017) 1156–1163, http://dx.doi.org/10.
crepancies were noted. 1016/j.applthermaleng.2017.01.090 https://ac.els-cdn.com/
Nevertheless, the CFD results from the second additional line prove S1359431117305458/1-s2.0-S1359431117305458-main.pdf?{\_}tid=ff89478c-
a52d-11e7-8763-00000aacb35f{\&}acdnat=1506700444{\_}
that the velocity field was more distorted in the tested machine than
dfefeb721193f0207dc927459f32d63f.
that obtained from the CFD model. [15] R. Khlissa, S. Vivier, G. Friedrich, K. El Kadri Benkara, B. Assaad, Thermal modeling
of an asymmetrical totally enclosed permanent magnet integrated starter generator,
Math Comput Simulat 130 (2016) 32–47, http://dx.doi.org/10.1016/j.matcom.
Acknowledgements
2015.06.015 https://ac.els-cdn.com/S0378475416000094/1-s2.0-
S0378475416000094-main.pdf?{\_}tid=cc757612-a9e3-11e7-a144-
Financial assistance was provided by grant no. DEC-2011/03/D/ 00000aab0f27{\&}acdnat=1507218332{\_}
ST8/04171 funded by the National Science Centre, Poland. The work of b8d5be2a4402cae3fe49bd2ab47c545cΩhttp://linkinghub.elsevier.com/retrieve/
pii/S0378475416000094.
BM was also partially supported by statutory research funds of the [16] P. Mellor, D. Roberts, D. Turner, Lumped parameter thermal model for electrical
Faculty of Energy and Environmental Engineering of the Silesian machines of TEFC design, IEEE Proc B Elec Power Appl 138 (5) (1991) 205, http://
University of Technology within grant no. BKM-559/RIE6/2016. dx.doi.org/10.1049/ip-b.1991.0025.
[17] A. Boglietti, A. Cavagnino, D. Staton, M. Shanel, M. Mueller, C. Mejuto, Evolution
and modern approaches for thermal analysis of electrical machines, IEEE Trans Ind
References Electron 56 (2009) 871–882, http://dx.doi.org/10.1109/TIE.2008.2011622.
[18] Y.A. Cengel, Heat transfer: a practical approach, 2nd ed., McGraw-Hill, London,
2007.
[1] C. Thiel, W. Nijs, S. Simoes, J. Schmidt, A. van Zyl, E. Schmid, The impact of the EU
[19] A. Boglietti, E. Carpaneto, M. Cossale, S. Vaschetto, Stator-winding thermal models
car CO2 regulation on the energy system and the role of electro-mobility to achieve
for short-time thermal transients: definition and validation, IEEE Trans Ind Electron
transport decarbonisation, Energy Pol 96 (2016) 153–166, http://dx.doi.org/10.
63 (5) (2016) 2713–2721, http://dx.doi.org/10.1109/TIE.2015.2511170.
1016/j.enpol.2016.05.043.
[20] Y.-Z. Li, S.-J. Zhu, Y. Li, Q. Lu, Temperature prediction and thermal boundary si-
[2] J.H. Langbroek, J.P. Franklin, Y.O. Susilo, The effect of policy incentives on electric
mulation using hardware-in-loop method for permanent magnet synchronous mo-
vehicle adoption, Energy Pol 94 (2016) 94–103, http://dx.doi.org/10.1016/j.enpol.
tors, IEEE/ASME Trans Mechatron 21 (1) (2015), http://dx.doi.org/10.1109/
2016.03.050 http://www.sciencedirect.com/science/article/pii/
TMECH.2015.2443800 1–1 http://ieeexplore.ieee.org/document/7121001/.
S0301421516301550.
[21] B. Melka, J. Smolka, Z. Bulinski, J. Hetmanczyk, An experimental and numerical
[3] K.G. Gallagher, S. Goebel, T. Greszler, M. Mathias, W. Oelerich, D. Eroglu, et al.,
analysis of temperature and velocity field in PM BLDC motor, Przeglad
Quantifying the promise of lithium - air batteries for electric vehicles, Energy
Elektrotechniczny 1 (3) (2016) 94–97, http://dx.doi.org/10.15199/48.2016.03.22
Environ Sci 7 (2014) 1555, http://dx.doi.org/10.1039/c3ee43870h.
http://sigma-not.pl/publikacja-97014-2016-3.html.
[4] L. Chen, L.L. Shaw, Recent advances in lithium-sulfur batteries, J Power Sources
[22] O. Meksi, A.O. Vargas, Numerical and experimental determination of external heat
267 (2014) 770–783, http://dx.doi.org/10.1016/j.jpowsour.2014.05.111 https://
transfer coefficient in small TENV electric machines, 2015 IEEE Energy Conversion
doi.org/10.1016/j.jpowsour.2014.05.111.
Congress and Exposition, ECCE 2015(3) (2015), pp. 2742–2749, http://dx.doi.org/
[5] S.M. Rezvanizaniani, Z. Liu, Y. Chen, J. Lee, Review and recent advances in battery
10.1109/ECCE.2015.7310044.
health monitoring and prognostics technologies for electric vehicle (EV) safety and
[23] M. Fasil, D. Plesner, J.H. Walther, N. Mijatovic, J. Holbøll, B.B. Jensen, Numerical

414
B. Melka et al. International Journal of Thermal Sciences 130 (2018) 406–415

and experimental investigation of heat flow in permanent magnet brushless DC hub for a permanent-magnet synchronous motor with phase-change-material packaging,
motor, SAE Int J Alternative Powertrains 4 (1) (2014), http://dx.doi.org/10.4271/ Appl Therm Eng (2016), http://dx.doi.org/10.1016/j.applthermaleng.2016.08.036
2014-01-2900 2014–01–2900 http://papers.sae.org/2014-01-2900/. https://ac.els-cdn.com/S1359431116313825/1-s2.0-S135431116313825-main.
[24] D. Staton, E. So, Determination of optimal thermal parameters for brushless per- pdf?{\_}tid=a7e65828-a530-11e7-8511-00000aacb361{\&}acdnat=
manent\nmagnet motor design, Conference record of 1998 IEEE industry applica- 1506701585{\_}14b77c6e1ead2795eb38eed0d755bf1c.
tions conference. Thirty-third IAS annual meeting, vol. 1, 1998, pp. 41–49. [31] Y. Li, Z. Ji, L. Yang, P. Zhang, X. Bing, J. Zhang, Thermal-fluid-structure coupling
[25] M. Grabowski, K. Urbaniec, J. Wernik, K.J. Wołosz, Numerical simulation and ex- analysis for valve plate friction pair of axial piston pump in electrohydrostatic ac-
perimental verification of heat transfer from a finned housing of an electric motor, tuator (EHA) of aircraft, Appl Math Model 47 (2017) 836–858 https://ac.els-cdn.
Energy Convers Manag 125 (2016) 91–96, http://dx.doi.org/10.1016/j.enconman. com/S0307904X16304395/1-s2.0-S0307904X16304395-main.pdf?{\_}tid=
2016.05.038. 7334fb28-f84d-11e7-a73d-00000aacb362{\&}acdnat=1515839899{\_}
[26] Y. Zhang, J. Ruan, T. Huang, X. Yang, H. Zhu, G. Yang, Calculation of temperature bbdeeba2a35f55686f6b08ab53c02c5e.
rise in air-cooled induction motors through 3-d coupled electromagnetic fluid-dy- [32] A. Aubert, S. Poncet, P. Le Gal, S. Viazzo, M. Le Bars, Velocity and temperature
namical and thermal finite-element analysis, IEEE Trans Magn 48 (2) (2012) measurements in a turbulent water-filled Taylor-Couette-Poiseuille system, Int J
1047–1050, http://dx.doi.org/10.1109/TMAG.2011.2174433. Therm Sci 90 (2015) 238–247, http://dx.doi.org/10.1016/j.ijthermalsci.2014.12.
[27] C. Jungreuthmayer, T. Bauml, O. Winter, M. Ganchev, H. Kapeller, a. Haumer, 018 https://doi.org/10.1016/j.ijthermalsci.2014.12.018.
et al., A detailed heat and fluid flow analysis of an internal permanent magnet [33] B. Melka, J. Smolka, Z. Bulinski, J. Hetmanczyk, D. Makiela, A validated numerical
synchronous machine by means of computational fluid dynamics, IEEE Trans Ind model of heat and mass transfer in a PM BLDC electric motor, (2016), pp.
Electron 59 (12) (2012) 4568–4578, http://dx.doi.org/10.1109/TIE.2011. 1409–1413, http://dx.doi.org/10.1109/SPEEDAM.2016.7525885 http://
2176696. ieeexplore.ieee.org/stamp/stamp.jsp?arnumber=7525885.
[28] T. Davin, J. Pell, E.,S. Harmand, R. Yu, Experimental study of oil cooling systems for [34] X. Zhang, J. Yu, G. Su, Z. Yao, P. Hao, F. He, PIV measurement and simulation of
electric motors, Appl Therm Eng (2015), http://dx.doi.org/10.1016/j. turbulent thermal free convection over a small heat source in a large enclosed
applthermaleng.2014.10.060 https://ac.els-cdn.com/S1359431114009314/1-s2.0- cavity, Build Environ 90 (2015) 105–113, http://dx.doi.org/10.1016/j.buildenv.
S1359431114009314-main.pdf?{\_}tid=ee8b7ad4-a52f-11e7-89dd- 2015.03.015 https://doi.org/10.1016/j.buildenv.2015.03.015.
00000aab0f27{\&}acdnat=1506701274{\_} [35] ANSYS® Academic Research, Release 17.0, help system: ANSYS, Inc.
248c775d636751144956e0823059cc8b. [36] R. Wrobel, P.H. Mellor, A general cuboidal element for three-dimensional thermal
[29] D. Hyun Lim, S. Chul Kim, Thermal performance of oil spray cooling system for in- modelling, IEEE Trans Magn 46 (8) (2010) 3197–3200, http://dx.doi.org/10.1109/
wheel motor in electric vehicles, Appl Therm Eng (2014), http://dx.doi.org/10. TMAG.2010.2043928.
1016/j.applthermaleng.2013.11.057 https://ac.els-cdn.com/ [37] K. Azar, J. Graebner, Experimental determination of thermal conductivity of printed
S1359431113008612/1-s2.0-S1359431113008612-main.pdf?{\_}tid=f7089b5c- wiring boards, Twelfth annual IEEE semiconductor thermal measurement and
a52e-11e7-ab8b-00000aacb35d{\&}acdnat=1506700859{\_} management symposium. Proceedings, IEEE, 1994, pp. 169–182, , http://dx.doi.
7e66e286db5bc5ee982e3cb4843b11cb. org/10.1109/STHERM.1996.545107 http://ieeexplore.ieee.org/lpdocs/epic03/
[30] S. Wang, Y. Li, Y.-Z. Li, J. Wang, X. Xiao, W. Guo, Transient cooling effect analyses wrapper.htm?arnumber=545107.

415

You might also like