You are on page 1of 15

5088

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

3-D Microtransformers for DCDC On-Chip


Power Conversion
Ali Moazenzadeh, Fralett Suarez Sandoval, Nils Spengler, Vlad Badilita, and Ulrike Wallrabe

AbstractWe address the miniaturization of power converters


by introducing novel 3-D microtransformers with magnetic core
for low-megahertz frequency applications. The core is fabricated
by lamination and microstructuring of Metglas 2714A magnetic
alloy. The solenoids of the microtransformers are wound around
the core using a ball-wedge wirebonder. The wirebonding process is fast, allowing the fabrication of solenoids with up to 40
turns in 10 s. The fabricated devices yield the high inductance
per unit volume of 2.95 H/mm3 and energy per unit volume of
133 nJ/mm3 at the frequency of 1 MHz. The power efficiency of
6476% is measured for different turns ratio with coupling factors
as high as 98%. To demonstrate the applicability of our passive
components, two PWM controllers were selected to implement an
isolated and a nonisolated switch-mode power supply. The isolated
converter operates with overall efficiency of 55% and maximum
output power of 136 mW; then, we experimentally demonstrate
how we increased this efficiency to 71% and output power to
408 mW. The nonisolated converter can deliver an overall efficiency of 81% with a maximum output power of 515 mW. Finally,
we benchmarked the results to underline the potential of the technology for power on-chip applications.
Index TermsDCDC power conversion, magnetic layered
films, micromachining, transformer.

I. INTRODUCTION
HE pressure of handheld electronic devices has brought
along an on-going trend of miniaturization applied to all
components. Among various parts, power converters are of a
great importance since they deliver the required power to each
function block of the system. Power converters generally consist
of two main parts: the control part, which is currently available
by means of integrated circuit (IC) technology, and the passive electronics part. Usually, a significant physical volume and

Manuscript received April 30, 2014; revised July 10, 2014 and September 20,
2014; accepted October 31, 2014. Date of publication November 6, 2014; date of
current version April 15, 2015. This work was supported by the DFG Graduate
School Embedded Microsystems under Grant 1103. The work of F. Suarez
Sandoval was supported by the National Council of Science and Technology
(CONACYT, Mexico), and by the General Ministry of International Affairs
of the Secretariat of Public Education (DGRI, SEP, Mexico). The work of V.
Badilita was supported by DFG through Contract BA 4275/2-1. Recommended
for publication by Associate Editor J. A. Cobos.
A. Moazenzadeh, F. Suarez Sandoval, V. Badilita, and U. Wallrabe are
with the Laboratory for Microactuators, Department of Microsystems Engineering (IMTEK), University of Freiburg, 79110 Freiburg, Germany (e-mail:
ali.moazenzadeh@imtek.uni-freiburg.de; fralett.suarez@imtek.uni-freiburg.de;
vlad.badilita@imtek.de; wallrabe@imtek.uni-freiburg.de).
N. Spengler is with the Freiburg Institute of Advanced Studies, University of Freiburg, 79104 Freiburg, Germany and also with the Laboratory for
Microactuators, Department of Microsystems Engineering (IMTEK), University of Freiburg, 79110 Freiburg, Germany (e-mail: nils.spengler@imtek.unifreiburg.de).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TPEL.2014.2368252

weight of the power converter unit is reserved by the magnetic


passive components, e.g., coils and transformers [1]. The main
function of these magnetic components is to periodically store
and release the energy [2]. The more we shrink down the size
of the magnetic components, the more difficult it becomes to
maintain their power handling capability. Therefore, a solution
is needed to overcome this challenge and allow an uncompromised fabrication of high-performance miniaturized converters.
One possibility to maintain the power handling capabilities
while miniaturizing the converters is to integrate a magnetic core
with the passive component [2]. Therefore, the device yields
higher inductance within the same footprint, thus being able
to drive more power. Increasing the switching frequency in the
converter offers another route which leads to further reduction
in the size of inductors [3]. A higher switching frequency means
that the converter needs to process a smaller amount of energy
during each switching cycle, for the same amount of power.
However, both solutions present further disadvantages.
Microstructuring of magnetic materials to obtain miniaturized
cores is still challenging, especially when it comes to highaspect-ratio structures. Electroplating and sputtering are widely
used to obtain thin magnetic layers [4][7]; however, the choice
of magnetic materials is limited for each of these processes,
respectively. Also, these techniques pose inherent limitations
in terms of thickness and the homogeneity of the magnetic
components in the bulk physical volume.
Increasing the converter switching frequency introduces additional high-frequency parasitic losses, via the skin or proximity
effects in the inductors [2]. Magnetic losses are also increased
due to hysteresis and eddy currents losses in the magnetic core
[8]. The hysteresis loss can be controlled by choosing the right
material composition as well as annealing of the core magnetic
material [9]. The eddy current losses can be minimized either
by using high resistivity magnetic materials or by splitting the
bulk magnetic core into thin layers, which are in the same thickness range of the skin depth in the frequency range of interest
[10]. However, both approaches increase the complexity of the
magnetic components fabrication.
The fabrication of 3-D microcoils using an automatic ballwedge wirebonder was initially reported by Kratt et al. [11].
Since then, we used the technology for several unconventional
applications in microsystems [12]. In [13], we report the fabrication of SU-8 core microtransformers for the very high frequency
regime applications. Also, the preliminary results on integrating
a rod-shape multilayer magnetic core to the wirebonded coils
was reported in [14]. Although the rod-shape magnetic core is
the easiest choice for the integration with the wirebonded microcoils, this approach has some disadvantages, mainly due to

0885-8993 2014 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications standards/publications/rights/index.html for more information.

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

5089

TABLE I
COMPARISON BETWEEN STATE-OF-THE-ART SOFT MAGNETIC MATERIALS USED FOR MINIATURIZED CORE FABRICATION
Type

Material

Fabrication

Advantages

Disadvantages

Ferrites

NiZn [23]
MnZn [24]

Screen printing

Resistance 
Simplicity of fabrication

Saturation magnetization
Relative permeability
Coercivity
Incompatible with MEMS fabrication

Metallic alloys

NiFe [4][8], [28], [47]


CoNiFe [48]

Electroplating
Sputtering

Saturation magnetization 
Relative permeability 

Resistance
Limited core thickness

Nanocrystalline
films

CoZrTa [38]
CoZrO [40]
CoFeSiB [39]

Sputtering

Resistance 
Saturation magnetization 
Relative permeability 
Low temperature fabrication

Limited core thickness


Costly fabrication

Amorphous
metal ribbons

Metglas 2714
Vitrovac 6025 [8], [46]

Lamination
Etching

Relative permeability 
Resistance
Simplicity of fabrication
Low temperature fabrication

Saturation magnetization
Sensitive to external processes
Limited ribbons thickness

the high demagnetization field in the open-loop magnetic core


configuration.
In this paper, we report significant progress allowing the integration of a closed-loop magnetic core to our established wirebonded microcoils. Meanwhile, improvements in the wirebonding technology enable us to fabricate the coils with twice the
turns than our previous reports [13], [14] directly on the magnetic core. A new assembly method has been introduced to align
and position the magnetic core structures on the final substrate in
a parallel way to speed up the fabrication process of the microtransformers. The applicability of our passive components has
been verified by implementing them in two different topologies
of switch-mode power supplies, an isolated and a nonisolated
topology. Altogether resulting in a new approach to fabricate
miniaturized on-chip transformers, for low-megahertz regime
(0.330 MHz) applications. In Section II, the state of the art of
soft magnetic materials used in power conversion applications
is presented. Section III reports on the design and fabrication of
the closed-loop magnetic core, as well as the integration with
our established wirebonded microcoil technology. In Section
IV, the frequency-dependent performance of the magnetic core,
as well as the fabricated microtransformers are characterized.
Meanwhile, we benchmarked the results with the state of the art
of published devices to underline the potential of the technology
for power on-chip applications. Finally, Section V reports on the
implementation of the microtransformers and microcoils in an
isolated and nonisolated dcdc converter, respectively.
II. SOFT MAGNETIC MATERIALS USED FOR MINIATURIZED
CORE FABRICATION
The state-of-the-art air core inductors with high quality factors of up to 30 and inductances per unit volume of up to
700 nH/mm3 have been reported [13], [15][18]; thus, magnetic
core inductors with lower inductance per unit volume would not
represent an advantage.
Soft magnetic materials are fundamental for microtransformers magnetic core. They enhance the efficiency and the power
density of transformers, thus enabling further miniaturization.
An appropriate material for the magnetic core of a power

transformer should have a high saturation magnetization, high


permeability, high resistivity, low coercivity, and high thermal
conductivity. Among the above-mentioned properties, resistivity, thermal conductivity, and saturation magnetization are intrinsic properties of a material, while the other depend on the
external effects, such as the shape and dimension of the core
[19], fabrication process [20], operating frequency [10], extra
annealing [20], and excitation waveform [21]. All these parameters need to be considered when designing an efficient core.
Table I summarizes the properties of different magnetic materials used for miniaturized core fabrication. Ferrites are popular
candidates as a core material for megahertz frequency regime
applications, mainly due to their high resistivity, which is necessary to damp the induced eddy current in the core at high
frequencies [22][26]. This advantage is paid by the relatively
low permeability and high coercivity, which limit the further
miniaturization of ferrite cores [4], [10]. Besides, the main disadvantage of ferrites is the relatively lower saturation flux density when compared to other types of core magnetic materials.
Furthermore, their typical fabrication is not compatible with
MEMS processing due to the high temperature needed for sintering [23].
Magnetic metallic alloys are other candidates for the core
material. Among them, PermalloyNiFeis commonly used
as a magnetic core material [4][8], [27][35]. A core layer
thickness of up to several micrometers is achievable by electrodeposition that enables manufacturing of 3-D cores [4]. The
fabrication process is well established, mainly due to the wide
usage of NiFe in magnetic recording heads technology [36]. Besides the MEMS compatible fabrication, metal alloys also have
good soft magnetic properties in terms of high permeability
and low hysteresis loss. However, the intrinsic low resistivity
of metals results in high eddy currents losses in the core for
high switching frequency applications [4]. Reducing the metallic film thickness to the nanometer range enables the fabrication
of nanolaminated Permalloy cores, which results in lower core
losses at high frequencies [37]. However, realizing such a core
needs a complex and expensive fabrication method. In order
to improve the resistivity of magnetic alloys and make thick
alloys applicable for high-frequency applications, researchers

5090

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

Fig. 1. (a) Exploded view of an UT magnetic core microtransformer chip. (b) Fabricated UT Metglas core in this paper compared to the smallest commercially
available Metglas core. The UT core has been placed inside a commercial toroidal core.

TABLE II
PHYSICAL PROPERTIES OF METGLAS 2714 [49], [50]
Property
Composition
Saturation magnetization
Max DC permeability (cast)
Saturation magnetostriction
Electrical resistivity
Crystallization temperature
Curie temperature

Value
Co6 6 Fe4 Ni1 Si1 5 B1 4
0.57 (T)
> 80 000
< 0.5
142 ( cm)
550 C
225 C

investigated several combinations of the magnetic materials together with nonmetallic elements, e.g., silicon, tantalum, and
oxygen [1], [38][45]. Compared to NiFe, these new composite
core materials offer higher resistance, the same range of saturation magnetization, and lower coercivity [24], [26].
Another type of composite soft magnetic materials are the
cobalt-based amorphous alloys [8], [46]. These alloys are fabricated based on the technology known as melt spinning and
rapid solidification, meaning that the liquid metal composite
gets cooled rapidly when it is poured onto a disc rolling at a
very high speed [20]. As a result of the rapid cooling, the alloys
have an amorphous atomic structure. This disordered structure
of the metal with no phase boundaries yields ultrasoft magnetic
properties with very low coercivity and two to three times higher
resistivity in comparison to crystalline alloys [20], qualifying
them as an appropriate transformer core material for low-power
and high-frequency applications.
In this paper, cobalt-based amorphous magnetic alloy was
used as a core material. 18-m-thick commercially available
Metglas 2714A ribbons were successively laminated to form
the magnetic core. Metglas 2714A was chosen as the core
material for its high permeability, low energy dissipation, and
high resistivity [49], [50]. Metglas is available as a magnetically anisotropic material, makes it appropriate choice for high-

Fig. 2. Results of the combination of different materials to glue the Metglas


layers and different methods of cutting the laminated stack. (a) Adhesive tape
+ nanosecond UV laser: The induced heat while cutting with the laser oxidizes
the Metglas layers. Also the inherent Gaussian shape of the laser beam resulted
in nonflat cutting walls. (b) Spin-coated epoxy glue (2011, Araldite) + wafer
dicing machine (DAD321, Disco): Pressing the layers resulted in uneven epoxy
layers with different thicknesses. (c) Adhesive tape + dicing machine: The
sticky tape (5601, Nitto) glue got dissolved during the dicing process upon long
term water exposure, resulting in the delamination of the laminated stack. (d)
Adhesive tape + EDM cut. The combination lead to a good surface quality on
the cutting edges without delamination of the stack.

frequency applications. The material is readily available as an


1825-m-thick ribbon, which eliminates the need for an electrodeposition step in the fabrication procedure of the magnetic
core. Metglas toroidal cores are commercially available with the
smallest outer diameter of 8.1 mm, inner diameter of 4.4 mm,
and height of 4.6 mm [49]. Fig. 1(b) compares the UT magnetic core presented in this paper to the smallest commercially
available Metglas core. The properties of Metglas 2714A are
summarized in Table II.

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

Fig. 3.

5091

Wirebonded UT core microtransformers fabrication process flow.

III. DESIGN AND FABRICATION


Fig. 1(a) shows an exploded view of the microtransformer
chip introduced in this paper. The magnetic core and the inductive coils were designed as 3-D structures. This results in
minimizing the footprint of the chip without compromising the
transformer performance. The small footprint of 5.4 mm2 is
enough to contain the magnetic core as well as the primary and
secondary coils. Additionally, when compared to 2-D deposited
cores, the larger cross-sectional area of the 3-D core enhances
the power-handling capability without reaching saturation. Having a 3-D shape for the inductive coils results in more inductance
density, and also more concentrated and uniform magnetic field,
which leads to stronger inductive coupling between the primary
and secondary coils than a 2-D design [13], [51].
The core is composed of two structures, one with a U-shape,
the other one with a T-shape, allowing an easy assembly to form
a closed-loop magnetic core. Winding both the primary and
secondary coils on the central post enhances the self-shielding
properties of the microtransformer. The designed magnetic cores
have the footprint of 3 mm 1 mm and the height of 2.5 mm.
For the fabrication of our multilayered UT magnetic cores,
37 layers of 18-m-thick Metglas 2714A were laminated with a
hand roller (DTS-HR, Asmetec). Prior to the lamination, the alloy surface was cleaned in an ultrasonic bath filled with pure isopropanol. The hand roller also removes dusts, which improves
the adhesion of the layers. Intermediate layers of 10-m-thick
polyester-based double-sided sticky tape (5601, Nitto) provided
both adhesion and electrical insulation between the Metglas layers. The lamination was performed on the surface of a heater
plate at the temperature of 80 C [see Fig. 3(a)]. The whole lamination process was carried out inside a clean environment of a
class II biological cabinet which provides the clean area. The
last step was to place the stack inside a hydraulic press (JAS105,
High force), for 12 h at the force of 4.5 kN and the temperature
of 45 C [see Fig. 3(b)]. This last step was introduced to guarantee the perfect uniformity and high adhesion strength for the
37 layers of Metglas, as well as for the 36 layers of sticky tape.
By employing an electrical discharge machine (EDM) (Robifil

2020SI, GF Agie Charmilles), the magnetic stacks were precisely cut into submillimeter core parts [see Fig. 3(c)]. The
cutting direction was chosen so that the easy magnetization axis
of the cores is aligned with their height. EDM cutting has the
advantage of producing good surface quality on the cut edges
while not changing the magnetic or physical properties of the
Metglas alloy. The use of EDM offers an almost unrestricted
way of structuring the laminated magnetic stacks. However,
this method is not applicable to fabricate cylindrical shape posts
and the EDM cut posts cross section are limited to square or
rectangular shapes. Fig. 2 shows the distinct proven methods to
laminate the Metglas layers as well as several cutting techniques
for the laminated stack.
To enable coil winding, the cores were arranged onto a
4 in borosilicate wafer. Borosilicate wafers were chosen to
avoid eddy currents induced in the substrate, hence minimizing
substrate-related losses. The bond pads were made by means of
standard UV lithography. After evaporating the 15 nm/150 nm
Cr/Au seed layer, a 20-m-thick mold for pads and traces was
patterned using AZ-9260 photoresist. A layer of 12 m of gold
was subsequently electroplated on top of the seed layer. In order
to define the bond positions with respect to the microtransformer
cores, 2 m of ma-N 1420 negative tone photoresist was spun
and patterned to structure the alignment marks on the substrate
[see Fig. 3(d)]. The substrate was finally diced into the 8 mm
8 mm chips using a wafer dicing machine (DAD321, Disco).
To enable the magnetic core assembly on the processed
borosilicate substrate, a 1.5 mm-thick paper-based plastic substrate was used as an alignment aid. Square through-holes were
patterned with a UV laser (TruMark 6330, Trumpf) to perfectly
accommodate the posts of the T-shape cores [see Fig. 3(e)]. The
distance between the through-holes was equal to the distance between the landing marks on the borosilicate wafer. Two vacuum
chucks fixed both the borosilicate (top) and the plastic substrate
(bottom) containing the structures to be positioned. Taking advantage of the borosilicate transparency and using precision
moving tables, the plastic substrate was aligned with respect to
the borosilicate substrate [see Fig. 3(f)]. A Z-direction moving

5092

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

same trajectory, and as previously demonstrated [13], the coils


are supposed to be very similar due to the high precision of the
wirebonding process.
In order to show the effects of adding the rod-shape Metglas
micromachined core, both coils have been characterized in the
range of 1 MHz to 1 GHz using an impedance analyzer (E4991A,
Agilent). The coil with the Metglas core resonates at 560 MHz,
while the other resonates at around 1 GHz [see Fig. 5(a)]. This
fact shows that the Metglas coil has more internal inductance
and capacitance since the resonance frequency is a function of
these two parameters:
f=

Fig. 4. Full view of a UT core microtransformer as well as some closeup


views of the laminated stack, the remaining gap on top of the core between the
T-shape and U-shape parts, the contact area between the T-shape and U-shape
core on bottom of the core as well as the wedge and ball bonds (in clockwise
direction starting from left).

table was used to raise the plastic substrate and bring the T-shape
structures in contact with the borosilicate wafer. Fast adhesive
(4204, Loctite) served as the bonding material. After the adhesive got cured, the plastic substrate was brought down leaving
behind the T-shape structures positioned [see Fig. 3(g)]. After a
short O2 -plasma cleaning step for 2 min, at 40 C, 0.3 mbar and
the power of 1000 W at 2.54 GHz, a modified, automated ballwedge wirebonder (3100plus , ESEC) was employed to wind the
primary and secondary solenoids directly around the post of the
T-structures. The starting height of the coil was set to 600 m
above the surface of the substrate allowing for a precise 100 m
clearance from the horizontal part of the T-structure core. Within
10 s, each solenoid was wound with up to 40 turns of insulated
25-m-thick copper or gold wire [see Fig. 3(h)]. After the winding process, the U-shape structure was positioned following the
same methodology as used for the T-structures, in this way the
microtransformer fabrication being completed [see Fig. 3(i) and
(j)]. Fig. 4 shows a full view of a UT core microtransformer chip
as well as some closeup SEM pictures of the important regions.
IV. CHARACTERIZATION RESULTS
A. Metglas Core Performance
Metglas is a soft magnetic material with ultrahigh dc permeability. However, its permeability is a function of different physical parameters. In order to characterize the high-frequency permeability of the Metglas, a rod-shape (magnetically open-loop)
cubic multilayered core was fabricated with the same process
as mentioned in Section III. The cubic core had the dimensions
of 0.8 mm 0.8 mm 1.2 mm (L W H). A wirebonded
microcoil with seven turns was wound around the post. Another
core with nominal exact shape and dimension as the Metglas
cubic core was fabricated by means of a thick SU-8 lithography
process, which later was wirebonded to realize the same coil
with seven turns. Both coils, the one on the Metglas rod-shape
core and the one on the SU-8 post, were wirebonded using the

2 LC

(1)

Fig. 5(b) shows the frequency-dependent inductance for each


coil. The SU-8 core coil shows a constant inductance over almost the full measurement range with an inductance of 70 nH;
the Metglas core coil, however, exhibits 4.42 times higher inductance at the frequency of 1 MHz. Until 10 MHz, it shows a
slight decrease of 10 nH in the inductance value; however, the
inductance falls rapidly with further increase in frequency.
The total magnetic flux density that is generated by the Metglas core coil is
B = 0 (Hc + M ) = 0 (Hc Nd M + M )

(2)

where Hc is the magnetic field produced by the coil, Nd is


the demagnetization factor of the magnetic core, and M is the
magnetization of the Metglas. The magnetic flux density of the
SU-8 core coil is
B = 0 Hc .

(3)

A demagnetization factor Ndz of a rectangular rod, magnetized along its long axis (z-direction), is found to be as expressed
below [52]:
1
(4)
Ndz =
(2n + 1)
where n is the dimensional ratio of the rectangular rod. Based
on the dimension of the rod-shape Metglas core, the demagnetization factor of the cubic Metglas core is 0.25 [52].
An electrical current I flowing through a coil creates a magnetic flux proportional to the current. The proportional constant is defined as the inductance L of the coil [20]. The inductance of the coil is only a function of its geometry and the
permeability of the surrounding medium. Since in our case,
both the rod-shape Metglas core and the SU-8 core coils were
geometrically identical, we can consider the inductance ratio
of the coils as the effective permeability e of the rod-shape
Metglas core. In order to calculate the relative permeability r
of the Metglas core from its effective permeability, we need to
consider the demagnetization factor of the core [53]:


 e (1 Ndz ) 
.
(5)
r = 
1 Ndz e 
By substituting the values of the demagnetization factor and
the effective permeability in equation (5), the relative permeability of the rod-shape Metglas core is measured to be 31.57 at
the frequency of 1 MHz. For an ideal closed-loop UT Metglas

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

5093

Fig. 5. Characterization result of two similar wirebonded coils, one with the rod-shape Metglas core and the other with a SU-8 (magnetically like air) core. (a)
Resonance frequency of the coil with magnetic core was decreased as a result of an increase in the internal inductance and capacitance. (b) Coil inductance has
been increased 4.42 times at the frequency of 1 MHz.

(106-683, Cascade Microtech). The scattering parameters of the


microtransformers were recorded simultaneously as a function
of a logarithmic frequency sweep from 300 kHz up to 300 MHz
as a S2 P file. The power efficiency of the microtransformers,
defined as the output load power versus the input power, was
computed directly from the scattering parameters Sxy , using
(6), where x and y denote the port numbers [54]:
Z 0 =

Fig. 6. Two-port characterization setup for the frequency-dependent


impedance measurements of the microtransformers.

core, the effective permeability of the core is expected to be the


same as its relative permeability; then, both should be equal to
31.57 at the frequency of 1 MHz. However, any airgap in the
closed-loop core structure could impact its effective permeability. Also, it should be considered that two-third of the magnetic
core volume in the both closed-loop (UT core) and open-loop
(rod-shape core) consist of Metglas and the rest was filled with
the double-sided tape, which is a nonmagnetic material.
B. UT Core Microtransformer Electrical Performance
The high-frequency characterization of the UT core microtransformers was performed using the two-port on-wafer
measurement method. The scattering parameters of the microtransformers were measured using a vector network analyzer
(E5071A, Agilent) connected to a probe-station (9000, Cascade
Microtech), equipped with two microprobes (SG/GS-500, Cascade Microtech) as shown in Fig. 6.
Prior to the measurements, open, short, load, and through
calibrations were done using an impedance standard substrate

|S21 |2
1 |S11 |2

(6)

where Z0 = 50 is the nominal impedance of the network analyzer which served as the load impedance of the measurements.
By converting the S-parameters to Z-parameters and by extracting the real and imaginary parts of the impedance for each port at
each frequency, the inductance, electrical resistance and quality
factor, as well as the coupling factor of the microtransformers
were calculated using
Lxy =

 [Zxy ]
2f

Rxy =  [Zxy ]
 [Zxy ]
 [Zxy ]

 [Z12 ]  [Z21 ]
.
k=
 [Z11 ]  [Z22 ]

Qxy =

(7)

On that basis, a UT core microtransformersample 1with


copper solenoids of 40 turns and turns ratio of 1:1 on a borosilicate substrate was characterized. The self-resonance frequency
of the microtransformer was found to be 28 MHz. The selfinductance of the primary and secondary solenoids were measured to be L11 = 43.7 H and L22 = 41.7 H, at 300 kHz,
as can be observed in Fig. 7. The self-inductance, however, decreased to the values of 39.91 and 37.93 H at the frequency of
1 MHz as a result of the internal loss of the Metglas core.
Thus, we achieved an inductance per unit volume of up to

5094

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

Fig. 7.

Two-port measured frequency-dependent inductance of sample 1.

Fig. 8.

Two-port measured frequency-dependent resistance of sample 1.

Fig. 9. Two-port measured frequency-dependent quality factor of the primary


() and the secondary () coils and impedance of the primary () and the
secondary () coils of sample 1.

2.95 H/mm3 and an energy per unit volume of 133 nJ/mm3


at a frequency of 1 MHz, both being higher than most of the
other published on-chip power microinductors and microtransformers [38], [55].
The two-port frequency-dependent resistance of the sample 1
is shown in Fig. 8. For both solenoids, the resistance increased
as the frequency increased. This is caused, on one hand, by the
skin and proximity effects in the inductor and, on the other hand,
by internal losses of the magnetic core, resulting from the hysteresis and eddy current losses. The quality factor of the primary
solenoid reached a maximum value of 7.1, whereas it was 6.8
in the secondary solenoid (see Fig. 9), which is comparable to
state-of-the-art power microinductors [38]. The coupling factor
had a constant value of 97.3% over the whole frequency range
below the self-resonance frequency indicates that the microtransformers had a minimal leakage inductance (see Fig. 10).
The high coupling factor is attributed to the 3-D geometry, the

Fig. 10. Two-port measured frequency-dependent efficiency () and coupling


factor (o) of sample 1.

Fig. 11.

Two-port measured frequency-dependent inductance of sample 2.

Fig. 12.

Two-port measured frequency-dependent resistance of sample 2.

magnetic core, the precision of the wirebonder, and the compact


design of the transformers. The efficiency of the microtransformer reached a maximum value of 71% at a frequency of
1.11 MHz for the 50 load, provided by the network analyzer
(see Fig. 10).
Another UT core microtransformer chipsample 2with
12 primary turns and a turns ratio of 1:2.5 made of gold wire
was fabricated on a printed circuit board (PCB) substrate. This
microtransformer was used later for the implementation of an
isolated dcdc converter, presented in the next section. The
measurement results are depicted in Figs. 1113. The PCB chip
was produced on a standard FR4 substrate with the thickness
of 500 m. 200-nm-thick gold layer was deposited on top of
the usual 35 m of copper to enable wirebonding on the chip.
Table III summarizes the main measurement results of the both
microtransformer chips.
To conclude this part, we benchmarked our results to previously published microinductors (pure RF microinductors were
excluded). Fig. 14 illustrates the maximum quality factor as a

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

5095

Fig. 13. Two-port measured frequency-dependent efficiency () and coupling


factor (o) of sample 2.
TABLE III
MEASUREMENT RESULTS OF THE MICROTRANSFORMER CHIPS, SAMPLE 1
AND SAMPLE 2

L1 1 (H)
L2 2 (H)
Q11m ax
Q22m ax
(%)
m a x (%)
Np rim a ry
N1 : N2
Material

@ f (MHz)

Sample 1

@ f (MHz)

Sample 2

1
1
0.320
0.320
< 28
1.11

39.9
37.9
7.1
6.8
97.3
71
40
1:1
Cu

1
1
0.548
0.355
< 55
2.57

3.7
22.1
4.3
6.2
97.1
66
12
1 : 2.5
Au

Fig. 15. Benchmark of the state-of-the-art microtransformers in terms of inductance per dc resistance (L/R D C ) as a function of the frequency at which
the maximum efficiency appears. The size of the bubbles indicates the power
efficiency of each transformer, whereas the color of a bubble distinguishes between air and magnetic core transformers. The label of each bubble represents
the reference, transformer type and coupling factor. The microtransformers presented in this paper are labeled with TW1 and TW2. All the devices except
in [32] and [45] were characterized using a 50 load.

(pure RF microtransformers were excluded). Fig. 15 shows the


inductance per dc resistance (L/RDC ), which represents the
low-frequency performance of the microtransformers as a function of the frequency of the maximum efficiency. The size of
the bubbles indicates the power efficiency of each transformer,
whereas the color of a bubble distinguishes between air and
magnetic core transformers. The label of each bubble represents the reference, transformer type, and coupling factor. The
transformers presented in this paper and represented by the
biggest bubbles convince through their high power efficiency,
high inductance per dc resistance, and high coupling factors.
For higher frequencies applications, perhaps lower inductance
is needed; then, our former microtransformer prototypes are
applicable [13], [14].
V. APPLICATION OF THE MICROTRANSFORMERS
AND MICROCOILS: DCDC CONVERTERS

Fig. 14. Benchmark of the state-of-the-art published microinductors in terms


of maximum quality factor as a function of the inductance per unit volume
(nH/mm3 ). The shape of each point indicates the type of magnetic core of each
inductor, whereas the color of each point distinguishes the frequency range
where each prototype is applicable. The inductor presented in this paper are
represented by the TW label.

function of the inductance per unit volume (nH/mm3 ) for several miniaturized inductors. The shape of each point indicates
the magnetic material type of each inductor, whereas the color of
each point distinguishes the frequency range where each prototype is applicable. The inductor presented in this paper is labeled
with TW and convinces through its high inductance per unit
volume and reasonably high quality factor. We further benchmarked the microtransformers to previously published devices

In order to assess the behavior of our microtransformers and


microinductors, we implemented them in two different switchmode power supplies, an isolated and a nonisolated topology,
respectively.
A. Isolated Converter
The isolated power supply is an asymmetrical half-bridge
Flyback converter, as depicted in the schematic of Fig. 16.
This topology uses a primary-side controller and the output
gets regulated based on the transformer turns ratio. The primary
side of the converter is a modified synchronous Buck converter
that comprises two switches, the high-side switch SHS and the
low-side switch SLS , as well as input and output capacitors Cin
and Cpri , respectively. If the filtering inductor of a conventional
Buck converter is replaced by the transformer T1 , having turns

5096

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

Fig. 16.

General schematic of an asymmetrical half-bridge Flyback converter.

Fig. 17.

Microtransformer chip connected to the Fly-buck converter PCB.

ratio of N1 :N2 and, furthermore, if the voltage of the secondary


winding is rectified by the diode D1 , one obtains the so-called
Fly-buck converter, which is a combination of the Buck and
Flyback topologies.
The pulse-width modulation (PWM) controller used in our
isolated converter is the TPS55010 (Texas Instruments). We designed the converter to supply an output voltage of 5 V, from an
input voltage range of 4.55.5 V at an adjustable frequency of
up to 2 MHz. After the component values needed for a proper
operation of the controller were calculated, a PCB layout was
designed and fabricated as illustrated in Fig. 17. The total footprint of the demonstration board was 18 mm 63 mm including
connectors used during characterization.
To connect our microtransformer to the converters board, we
designed a second PCB with an area of 5 mm 5 mm, shown
in the zoomed-in view of Fig. 17. Bond pads for the winding
process can be found on the top layer of the chip, pads for
soldering a pin header are located on the bottom face, while
the connection between both chip layers is realized by four vias.
The converter board is equipped with a pin receptacle that allows
electrical connection to the microtransformer.
The efficiency curve as well as the percentage of load and
line regulation [56] of the Fly-buck converter were extracted using four precision digital multimeters (34401A, Agilent). These
three figures of merit for power supplies were obtained by simultaneously recording input and output voltages, as well as
currents of the converter. The converter was operating with a

Fig. 18. Efficiency of a 5 V output Fly-buck converter operating with sample


2 at a switching frequency of 1.82 MHz.

Fig. 19. Characterization of a 2.7 V output synchronous Buck converter operating with an UT microcoil with 12 turns made out of a gold wire with diameter
of 25 m, at a frequency of 0.75 MHz. (a) Efficiency curve for different input
voltages. (b) Percentage of load regulation for different input voltages.

microtransformer with 12 turns in the primary winding and


turns ratio of 1:2.5sample 2at a switching frequency of
1.82 MHz. Fig. 18 shows the efficiency of the converter as a
function of the output current.
The maximum converter efficiency that we measured was
55% as shown by Fig. 18, while the percentage of load regulation
was 37%, measured from the no-load to full-load conditions.

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

5097

TABLE IV
STATE-OF-THE-ART BUCK CONVERTERS SPECIFICATIONS
Parameter

Item nr.
[57]

Type

f [M Hz]

V i n [V ]

V o [V ]

Po m a x
[W ]

[% ]
@P o m a x

m a x
[% ]

I o [m A] @
m a x

PLOR1 [% ]

PLNR2 [% ]

Power
density
[W/mm3 ]

Commercial
module

0.4

3.3

16

85

95

1200

2.77 @
I o = 3.9 A

0.022

1.5
0.6
1.5

3.3
5
12

1.5
1.5
3.3

12
15
1.65

77
82
82

89
91
82

2000
3000
500

0.6 @
I o = 0.5 A

0.27 @
I o = 3.9 A,
V o = 2.25 V

0.03 @
T = 25 C,
V o = 9.5 V

[58]
[59]
[60]

0.031
0.019
851.6
106

0.6

5.5

3.3

9.9

93.5

93.5

3000

1.6

3.3

0.99

85

85

300

0.002

1.5
5

5
12

1.2
1.2

48
12

85
86

89
87

20,000
60,000

0.9 @
I o = 0.3A

0.041
0.048

5.5

4.2

1.8

1.08

85

88

200

0.51 @
I o = 0.6 A

0.157

0.8
< 5 PFM

5
3.6

2
1.8

0.6
0.6

70
87

80
90

60
200

0.69 @
I o = 0.2 A,
Vo = 3 V

1.8

3.6

92

93.4

140

0.081

100
200
30
200
50

1.2
3.3
3
3.6
3.3

0.95
2.3
1.5
2.2
2

0.18
0.161
0.45
0.730
0.546

83.7
62
62.5
77
68.7

87.5
62
71.7
77
68.7

90
70
120
380
300

9.9 @
I o = 0.2 A

3.57 @
I o = 0.05 A,
V o = 0.5 V

0.009
0.4
0.110

0.075

170

1.2

0.9

0.315

75

77.9

190

0.230

[75]

2.25

2.6

1.2

0.8

48

58

183

0.4

[76]
[77]
[78]

3
660
180

4
2.2
3.3

3
0.8
2.5

1
0.044
0.675

73.7
31
77.9

75
31
80.5

266
55
170

1.98 @
I o = 0.6 A

0.005
0.006
0.2

0.75

3.3

2.7

0.515

62

81.2

85.7

4.63 @
I o = 0.15 A

0.81 @
I o = 0.06 A,
Vo = 2 V

0.724

63

65.3

270

0.6

52

52

225

0.35 @
I o = 0.15 A
1.78 @
I o = 0.15 A

[61]
[62]

Research
module

[63]
[64]
[65]

PwrSiP
product

[66]
[67]
[68]

PwrSiP
research

[69]
[70]
[71]
[72]
[73]

[74]

This work

PwrSoC
research

Research
module

0.108

0.045
0.104

PLOR stands for percentage of load regulation [56] and is defined as PLOR = 100 (V o I o = 0 V o F L )/V o F L ,where V o I o = 0 .is the output voltage of the converter at
zero output current and V o F L is the output voltage at full load (maximum output current).
2
PLNR stands for percent of line regulation [56] and is defined as PLNR = 100 V o /(V o n o m V i n ),where V o is the total variation of the output voltage when the
converter is subjected to the total variation of the input voltage V i n , and V o n o m is the converters nominal output voltage.
1

The percentage of line regulation was 3.2% for a variation in


the input voltage from 4.5 to 5.5 V at the fixed value of output
current of 20 mA. The power density of the microtransformer
was 1.11 W/cm3 .
B. Nonisolated Converter Implementation
and Characterization
The UT-shape magnetic core microcoil was also tested as
part of a synchronous Buck converter implemented with the
PWM controller TPS43000 (Texas Instruments). Synchronous
rectification was chosen to enhance the overall circuit efficiency
at the desired output voltage level. The converter was designed
to supply an output voltage of 2.7 V, for an input voltage range
of 3.35 V at an adjustable switching frequency of up to 2 MHz.

The way of connecting the microcoil was similar as explained


before. The footprint of this converters PCB was 18 mm
54 mm including connectors used during characterization. The
efficiency curves as a function of output current for three different values of input voltages were extracted and are depicted
in Fig. 19(a). Fig. 19(b) shows the load regulation for an output
current change of 250 mA.
The total converter efficiency diminishes with increasing input voltage. This is expected because the converter losses such
as conduction and switching losses in both MOSFETs and in
the inductor, as well as the quiescent loss of the IC increase
with increasing input voltage. An input voltage of 3.3 V offers
a range of 70 mA working at the high efficiency of 81%. The
power density of the microcoil was 3.7 W/cm3 . The efficiency
at an input voltage of 5 V is higher at light loads (030 mA) than

5098

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

TABLE V
STATE-OF-THE-ART ISOLATED CONVERTERS SPECIFICATIONS
Parameter

Item nr
[18]

Type

f [M Hz]

Topology

Transformer
volume
[m m 3 ]

Vin
[V ]

Vo
[V ]

Po m a x
[W ]

[% ]
@P o m a x

m a x
[% ]

I o [m A]
@ m a x

PLOR [% ]

PLNR [% ]

Research
module

Half-bridge
With ZVS
Forward
Forward
Resonant

31.24

12

0.5

34

34

100

10
0.819
23

10
6
12

3
1.9
5.5

0.3
0.08
0.66

33
3
49.5

33
3
49.5

100
42
120

0.136

47

55

22.5

37 @
I o = 0.04 A

0.296

62.5

67.2

42

14 @
I o = 0.04 A

0.408

64.2

71

65

11 @
I o = 0.04 A

3.2 @
I o = 0.02 A,
Vo = 1 V
1.2 @
I o = 0.02 A,
Vo = 1 V
0.8 @
I o = 0.02 A,
Vo = 1 V

[79]
[80]
[81]
This work

15
6.5
6
Research
module

1.82

Fly-buck Wire:
Au 25 m
Fly-buck Wire:
Cu 34 m

7.5

Fly-buck Wire:
Cu 100 m

the other two cases of input voltage, because the controller was
set to enter an automatic pulsed frequency modulation (PFM)
mode. In this mode, the IC operates in either PWM or in PFM
mode. When the device is initially powered, it operates in fixed
PWM mode until completion of the soft-start. It remains in this
mode until it senses that the converter is on the verge of breaking
into discontinuous operation. At this point, the controller goes
to sleep mode until the output voltage has decreased by 2%.
The controller then starts again at a fixed PWM frequency for a
short duration and increases the output voltage. If the controller
again senses discontinuous operation, the cycle repeats. Since
the duty cycle in PFM mode is low, all losses are reduced which
results in efficiency improvement at light loads. The two modulation modes are clearly identified in Fig. 19(b) for the three
input voltages.
The percentage of load regulation of the synchronous Buck
converter for an output current change of 150 mA is 4.63%,
0.35% and 1.78% with an input voltage of 3.3, 4, and 5 V, respectively, as summarized in Table IV. As depicted in Fig. 19(b),
with an input voltage of 3.3 V and an output current greater than
140 mA, the efficiency reduces dramatically due to the decrease
in output voltages, leading to the conclusion that a compromise should be made between the overall circuit efficiency and
possible output voltage for load currents greater than 140 mA.
The highest percentage of line regulation that was measured is
0.81%, at a load current of 60 mA and a change in input voltage
from 3 to 5 V.
C. Effects of the Wire Diameter on the Microtransformer:
Impact on the Circuit Efficiency
At the efficiency maxima of the isolated converter, we have
calculated that the ohmic loss in the microtransformer accounts
for 57.5% of the total losses. In order to improve the efficiency
of the converter, as well as to increase the power density of
the microtransformer, we replaced the gold wire of 25 m in
diameter with a hand-wound copper wire with a diameter of
34 m in the first case, and of 100 m in the second case.
The new efficiency curves were measured and are presented in

Fig. 20. Efficiency curve of a 5-V output Fly-buck converter operating with
an UT microtransformer with 12 turns on the primary side and turns ratio of
1:2.5. The transformers were made of wirebonded gold with diameter of 25 m
(), hand-wound copper with diameter of 34 m () and hand-wound copper
with diameter of 100 m ().

Fig. 20. We have maintained the efficiency curve from Fig. 18


to ease the comparative process.
Increasing the wire diameter as well as using copper, which
has higher conductivity than gold, reduces the ohmic loss of the
microtransformer. Fig. 20 shows that the efficiency maximum
increased from 55% to 67% and 71% for the copper wires with
34 and 100 m diameter, respectively. With increasing wire
diameter, the maximum possible output current has increased
as well.
Furthermore, high converter efficiencies are maintained for
a wider range of loads at increased wire diameters. The lowefficiency maxima of sample 2 is caused by the poor load regulation of the converter when working with thin wire diameters.
The percentage of load regulation has decreased from 37% to
14% and to 11% for the output current range of 040 mA,
whereas the percentage of line regulation decreased from 3.1%

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

TABLE VI
POWER LOSS DISTRIBUTION OF THE FLY-BUCK CONVERTER OPERATING WITH
DIFFERENT WIRE DIAMETERS FOR THE UT MICROTRANSFORMERS
Measured input power [mW]
Measured output power [mW]
Power losses

172.26
94.77

UT transformer [mW]
MOSFETs [mW]
Diode [mW]
Input, primary and output capacitors [mW]
Total [mW]

43.03
22.64
11.25
0.57
77.49

399.61
283.8
[%]
55.53
29.21
14.52
0.74
100

[%]
42.54
37.23
32.5
3.58
115.81

36.7
32.15
28.06
3.09
100

to 1.1% and to 0.8%, for a variation in the input voltage from


4.5 to 5.5 V at a fixed value of output current of 20 mA. Table V summarizes the most important characteristics of the Flybuck converter operating with the distinct microtransformers
in comparison with other reported results. In contrast with the
large number of references listed in Table IV, Table V shows
only a few isolated converters using miniaturized transformers.
Moreover, the reported efficiencies and output power capabilities of these are considerably lower than those of nonisolated
converters.
The power losses in the microtransformer made from 100m-diameter copper wire account for only 36% of the total loss
in converter, comparable to the total power losses of the semiconductor devices. Table VI summarizes the different sources of
loss at the efficiency maxima for the smallest and the largest wire
diameter used. The power losses in the semiconductor devices
account for both switching and conduction losses. The losses
in the primary and secondary windings of the microtransformer
were summed and are accounted as one single value.
The high converter efficiency that we measured with the
hand-wound microtransformers clearly indicates that a moderate increase in wire diameter may lead to a drastic increase
in efficiency while maintaining batch fabrication capability, as
wirebonder machines can be configured to work with wire diameters of 100 m.
VI. CONCLUSION
Three-dimensional microtransformers were manufactured by
combining a newly developed closed-loop micromachined magnetic core with wirebonded microcoils meant for on-chip power
conversion applications in the low-megahertz region. No cleanroom processes were needed, neither for the fabrication of the
magnetic core nor for the fabrication of the inductive coils.
Thus, we present a process which is relatively cost effective and
straightforward to realize in a usual laboratory environment.
This fabrication technique lends itself to direct on-chip integration of the microtransformers into any converter circuits.
Former reports on wirebonded microcoils treated individual
solenoids showing low mutual inductances [82], whereas in
this study, transformers with strongly coupled microsolenoids
are presented. The effective permeability of the new cores has
increased seven times at the frequency of 1 MHz, as the result
of the closed-loop design. Moreover, improvements in the wirebonding technology enabled the fabrication of coils with twice
more turns directly on the magnetic core. As a result, the new

5099

microtransformers generate at 1 MHz 7.4 times more inductance density than the rod-shape core microtransformers previously reported in [14] and 26.7 times more inductance density
than our air core microtransformers reported in [13]. We benchmarked the devices to previously published microinductors and
microtransformers, both from dc and ac performances points
of view, to underline the potential of the technology for power
on-chip applications.
We further investigated how the winding wire diameter influences the system efficiency of the power converters. The power
losses of the microtransformer made from 100-m-thick handwound copper wire account for the 36% of the total loss in the
circuit. This is 19% less than for the microtransformer wound
with the 25-m-thick wirebonded gold wire. These results prove
that if the wirebonder machine is calibrated to wind coils made
of increased wire diameters (>25 m), the system efficiency of
the converters could be even higher than what we have achieved
so far with the hand-wound devices, because the winding precision of this machine also reduces the leakage inductance of our
microtransformers, thus reducing switching losses.
Therefore, adaptation of our wirebonder-based coil winding technique to thicker copper wire appoints to more efficient
magnetic devices, consequently better system efficiency. Moreover, our laboratory style manual pick-and-place assembly of
the magnetic cores can be easily automated by the use of pickand-place machines to push our process toward industrial manufacturing standards.
ACKNOWLEDGMENT
The authors would like to thank S. M. Torres Delgado
(IMTEK, Laboratory of Simulations) for access to electrical
characterization equipment, J. Hempel (IMTEK, Laboratory of
Electrical Instrumentation) for access to probe station, and A.
Gehringer (IMTEK, Laboratory for Process Technology) for
EDM cut of the magnetic cores. They would also like to thank M.
Pauls (IMTEK, Laboratory for Microactuators) and the Gisela
and Erwin Sick Chair for Micro-optics for access to the SEM.
They further acknowledge Prof. D. P. Arnold (University of
Florida) for the valuable technical discussion on the demagnetization effect.
REFERENCES
[1] D. Yao, C. G. Levey, and C. R. Sullivan, Microfabricated V-groove power
inductors using multilayer Co-Zr-O thin films for very-high-frequency
DC-DC converters, in Proc. IEEE Energy Convers. Congr. Expo., 2011,
pp. 18451852.
[2] C. R. Sullivan, Integrating magnetics for on-chip power: Challenges and
opportunities, in Proc. IEEE Custom Intergr. Circuits Conf., Sep. 2009,
pp. 291298.
[3] T. ODonnell, N. Wang, R. Meere, F. Rhen, S. Roy, D. OSullivan,
and S. C. OMathuna, Microfabricated inductors for 20 MHz dcdc converters, in Proc. Appl. Power Electron. Conf. Expo., 2008,
pp. 689693.
[4] J. Y. Park and M. G. Allen, Ultralow-profile micromachined power inductors with highly laminated Ni/Fe cores: Application to low-megahertz
DC-DC converters, IEEE Trans. Magn., vol. 39, no. 5, pp. 31843186,
Sep. 2003.
[5] C. R. Sullivan and S. R. Sanders, Design of microfabricated transformers
and inductors for high-frequency power conversion, IEEE Trans. Power
Electron., vol. 11, no. 2, pp. 228238, Mar. 1996.
[6] R. J. Rassel, C. F. Hiatt, J. DeCramer, and S. A. Campbell, Fabrication
and characterization of a solenoid-type microtransformer, IEEE Trans.
Magn., vol. 39, no. 1, pp. 553558, Jan. 2003.

5100

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

[7] M. Brunet, T. ODonnell, J. OBrien, P. McCloskey, and S. C. OMathuna,


Design study and fabrication techniques for high power density microtransformers, in Proc. Appl. Power Electron. Conf. Expo., 2001,
pp. 11891195.
[8] D. Flynn, A. Toon, L. Allen, R. Dhariwal, and M. P. Y. Desmulliez,
Characterization of core materials for microscale magnetic components
operating in the megahertz frequency range, IEEE Trans. Magn., vol. 43,
no. 7, pp. 31713180, Jul. 2007.
[9] R. M. Bozorth, Ferromagnetism. New York, NY, USA: Wiley, 1993,
p. 968.
[10] D. P. Arnold, I. Zana, and M. G. Allen, Analysis and optimization
of vertically oriented, through-wafer, laminated magnetic cores in silicon, J. Micromech. Microeng., vol. 15, no. 5, pp. 971977, May
2005.
[11] K. Kratt, V. Badilita, T. Burger, J. G. Korvink, and U. Wallrabe, A fully
MEMS-compatible process for 3D high aspect ratio micro coils obtained
with an automatic wire bonder, J. Micromech. Microeng., vol. 20, no. 1,
pp. 015021-1015021-11, Jan. 2010.
[12] A. C. Fischer, J. G. Korvink, N. Roxhed, G. Stemme, U. Wallrabe, and
F. Niklaus, Unconventional applications of wire bonding create opportunities for microsystem integration, J. Micromech. Microeng., vol. 23,
no. 8, pp. 083001-1083001-18, Aug. 2013.
[13] A. Moazenzadeh, N. Spengler, R. Lausecker, A. Rezvani, M. Mayer,
J. G. Korvink, and U. Wallrabe, Wire bonded 3D coils render air core
microtransformers competitive, J. Micromech. Microeng., vol. 23, no. 11,
pp. 114020-1114020-11, Nov. 2013.
[14] A. Moazenzadeh, N. Spengler, and U. Wallrabe, High-performance, 3Dmicrotransformers on multilayered magnetic cores, in Proc. IEEE Micro
Electro Mech. Syst., 2013, pp. 287290.
[15] J. Kim, F. Herrault, X. Yu, M. Kim, R. H. Shafer, and M. G. Allen,
Microfabrication of air core power inductors with metal-encapsulated
polymer vias, J. Micromech. Microeng., vol. 23, no. 3, pp. 035006-1
035006-7, Mar. 2013.
[16] C. D. Meyer, S. S. Bedair, B. C. Morgan, and D. P. Arnold, Micromachined wiring board with integrated microinductor for chip-scale power
conversion, IEEE Trans. Power Electron., vol. 29, no. 11, pp. 60526063,
Jul. 2014.
[17] C. D. Meyer, S. S. Bedair, B. C. Morgan, and D. P. Arnold, Highinductance-density, air-core, power inductors, and transformers designed
for operation at 100500 MHz, IEEE Trans. Magn., vol. 46, no. 6,
pp. 22362239, Jun. 2010.
[18] S. Tang, S. Hui, and H. Chung, A low-profile low-power converter
with coreless PCB isolation transformer, IEEE Trans. Power Electron.,
vol. 16, no. 3, pp. 311315, May 2001.
[19] B. Jamieson, T. ODonnell, S. Kulkarni, and S. Roy, Shape-independent
permeability model for uniaxially-anisotropic ferromagnetic thin films,
Appl. Phys. Lett., vol. 96, no. 20, pp. 202509-1202509-4, 2010.
[20] B. D. Cullity and C. D. Graham, Introduction to Magnetic Materials, 2nd
ed. New York, NY, USA: Wiley, 2009, p. 832.
[21] W. Wieserman, G. Schwarze, and J. Niedra, Magnetic and electrical
characteristics of cobalt-based amorphous materials and comparison to a
permalloy type polycrystalline material, in Proc. 3rd Int. Energy Convers.
Eng. Conf., 2005, pp. 5720-15720-17.
[22] P. Galle, X. Wu, L. Milner, S.-H. Kim, P. Johnson, P. Smeys, P. Hopper, K. Hwang, and M. G. Allen, Ultra-compact power conversion
based on a CMOS-compatible microfabricated power inductor with minimized core losses, in Proc. Electron. Compon. Technol. Conf., 2007,
pp. 18891894.
[23] J. Y. Park and M. G. Allen, Development of magnetic materials and
processing techniques applicable to integrated micromagnetic devices, J.
Micromech. Microeng., vol. 8, pp. 307316, 1998.
[24] I. Kowase, T. Sato, K. Yamasawa, and Y. Miura, A planar inductor
using Mn-Zn ferrite/polyimide composite thick film for low-Voltage and
large-current DC-DC converter, IEEE Trans. Magn., vol. 41, no. 10,
pp. 39913993, Oct. 2005.
[25] M. Raimann, A. Peter, D. Mager, U. Wallrabe, and J. G. Korvink,
Microtransformer-based isolated signal and power transmission, IEEE
Trans. Power Electron., vol. 27, no. 9, pp. 39964004, Sep. 2012.
[26] Y. Fukuda, T. Mizoguchi, S. Yatabe, and T. Tachi, Planar inductor with
ferrite layers for DCDC converter, IEEE Trans. Magn., vol. 39, no. 4,
pp. 20572061, Jul. 2003.
[27] T. O. Donnell, N. Wang, M. Brunet, S. Roy, A. Connell, J. Power,
C. O. Mathuna, and P. Mccloskey, Thin film micro-transformers for
future power conversion, in Proc. Appl. Power Electron. Conf. Expo.,
2004, pp. 939944.

[28] N. Wang, T. ODonnell, S. Roy, P. McCloskey, and C. OMathuna, Microinductors integrated on silicon for power supply on chip, J. Magn. Magn.
Mater., vol. 316, no. 2, pp. e233e237, Sep. 2007.
[29] N. Wang, T. ODonnell, S. Roy, S. Kulkarni, P. Mccloskey, and S. C.
OMathuna, Thin film microtransformer integrated on silicon for signal
isolation, IEEE Trans. Magn., vol. 43, no. 6, pp. 27192721, Jun. 2007.
[30] C. H. Ahn, Y. J. Kim, and M. G. Allen, A fully integrated planar toroidal
inductor with a micromachined nickel-iron magnetic bar, IEEE Trans.
Compon. Packag. Manuf. Technol., vol. 17, no. 3, pp. 463469, Sep. 1994.
[31] M. Xu and T. Liakopoulos, A microfabricated transformer for highfrequency power or signal conversion, IEEE Trans. Magn., vol. 34,
no. 4, pp. 13691371, Jul. 1998.
[32] M. Brunet, T. C. ODonnell, L. Baud, J. OBrien, P. McCloskey, and S.
C. OMathuna, Electrical performance of microtransformers for DC-DC
converter applications, IEEE Trans. Magn., vol. 38, no. 5, pp. 31743176,
Sep. 2002.
[33] T. M. Liakopoulos and C. H. Ahn, 3-D microfabricated toroidal planar
inductors with different magnetic core schemes for MEMS and power
electronic applications, IEEE Trans. Magn., vol. 35, no. 5, pp. 3679
3681, Sep. 1999.
[34] A. Zolfaghari, A. Chan, and B. Razavi, Stacked inductors and transformers in CMOS technology, IEEE J. Solid-State Circuits, vol. 36, no. 4,
pp. 620628, Apr. 2001.
[35] T. M. Andersen, C. M. Zingerli, F. Krismer, J. W. Kolar, N. Wang, and
C. OMathuna, Modeling and Pareto optimization of microfabricated
inductors for power supply on chip, IEEE Trans. Power Electron., vol.
28, no. 9, pp. 44224430, Sep. 2013.
[36] T. Osaka, Recent development of magnetic recording head core materials
by plating method, Electrochim. Acta, vol. 44, pp. 38853890, Jun. 1999.
[37] J. Kim, M. Kim, P. Galle, F. Herrault, R. Shafer, J. Y. Park, and
M. G. Allen, Nanolaminated permalloy core for high-flux, highfrequency ultracompact power conversion, IEEE Trans. Power Electron.,
vol. 28, no. 9, pp. 43764383, Sep. 2013.
[38] D. S. Gardner, G. Schrom, F. Paillet, B. Jamieson, T. Karnik, and S. Borkar,
Review of on-chip inductor structures with magnetic films, IEEE Trans.
Magn., vol. 45, no. 10, pp. 47604766, Oct. 2009.
[39] H. Kurata, K. Shirakawa, O. Nakazima, and K. Murakami, Solenoid-type
thin-film micro-transformer, IEEE Transl. J. Magn. Jpn., vol. 9, no. 3,
pp. 9094, May 1994.
[40] S. Prabhakaran, Y. Sun, P. Dhagat, W. D. Li, and C. R. Sullivan, Microfabricated V-groove power inductors for high-current low-voltage fasttransient DC-DC converters, in Proc. IEEE 36th Conf. Power Electron.
Spec., 2005, pp. 15131519.
[41] D. W. Lee, K. Hwang, and X. W. Wang, Fabrication and analysis of
high-performance integrated solenoid inductor with magnetic core, IEEE
Trans. Magn., vol. 44, no. 11, pp. 40894095, Nov. 2008.
[42] D. Yao, C. Levey, R. Tian, and C. Sullivan, Microfabricated V-groove
power inductors using multilayer CoZrO thin films for very-highfrequency DCDC converters, IEEE Trans. Power Electron., vol. 28,
no. 9, pp. 43844394, Sep. 2013.
[43] M. Yamaguchi, K. Suezawa, K. I. Arai, Y. Takahashi, S. Kikuchi, Y.
Shimada, W. D. Li, S. Tanabe, and K. Ito, Microfabrication and characteristics of magnetic thin-film inductors in the ultrahigh frequency region,
J. Appl. Phys., vol. 85, no. 11, pp. 79197922, 1999.
[44] M. Mino, T. Yachi, A. Tago, K. Yanagisawa, and K. Sakakibara, Planar microtransformer with monolithically-integrated rectifier diodes for
micro-switching converters, IEEE Trans. Magn., vol. 32, no. 2, pp. 291
296, Mar. 1996.
[45] K. Yamaguchi, S. Ohnuma, T. Imagawa, J. Toriu, H. Matsuki, and
K. Murakami, Characteristics of a thin film microtransformer with circular spiral coils, IEEE Trans. Magn., vol. 29, no. 5, pp. 22322237, Sep.
1993.
[46] O. Dezuari, S. E. Gilbert, E. Belloy, and M. A. M. Gijs, High inductance
planar transformers, Sens. Actuators A Phys., vol. 81, pp. 355358, Apr.
2000.
[47] B. Orlando, R. Hida, R. Cuchet, M. Audoin, B. Viala, X. Gagnard, P. Ancey, X. U. M. R. Cnrs, and L. Cedex, Low-resistance integrated toroidal
inductor for power management, IEEE Trans. Magn., vol. 42, no. 10,
pp. 33743376, Oct. 2006.
[48] R. W. Filas, T. M. Liakopoulos, and A. Lotfi, Micromagnetic device
having alloy of cobalt, phosphorus and iron, U.S. Patent 6 624 498
B22003, 2003.
[49] (2014, Jan. 31). [Online]. Available: http://www.hitachimetals.com/
product/amorphous/magampsquareloopcores/documents/magamp_opt.pdf

MOAZENZADEH et al.: 3-D MICROTRANSFORMERS FOR DCDC ON-CHIP POWER CONVERSION

[50] S. Brugger and O. Paul, Field-concentrator-based resonant magnetic


sensor with integrated planar coils, J. Microelectromech. Syst., vol. 18,
no. 6, pp. 14321443, Dec. 2009.
[51] K. Ehrmann, N. Saillen, F. Vincent, M. Stettler, M. Jordan, F. M. Wurm,
P.-A. Besse, and R. Popovic, Microfabricated solenoids and Helmholtz
coils for NMR spectroscopy of mammalian cells, Lab Chip, vol. 7,
no. 3, pp. 37380, Mar. 2007.
[52] M. Sato and Y. Ishii, Simple and approximate expressions of demagnetizing factors of uniformly magnetized rectangular rod and cylinder, J.
Appl. Phys., vol. 66, no. 2, pp. 983985, 1989.
[53] S. Tumanski, Induction coil sensorsA review, Meas. Sci. Technol.,
vol. 18, no. 7, pp. 3146, Mar. 2007.
[54] D. M. Pozar, Microwave Engineering, 4th ed. Amherst, MA, USA: Wiley,
2011, pp. 559570.
[55] S. C. OMathuna, N. Wang, S. Kulkarni, and S. Roy, Review of integrated magnetics for power supply on chip (PwrSoC), IEEE Trans.
Power Electron., vol. 27, no. 11, pp. 47994816, Nov. 2012.
[56] M. K. Kazimierczuk, Pulse-Width Modulated DC-DC Power Converters.
Chichester, U.K.: Wiley, 2008, pp. 69.
[57] Vishay product, FX5545G108.
[58] Linear Tec. product, LTM4608.
[59] ISL8201M.
[60] ROHM product, BP5275-50.
[61] PicoTLynx, APTH003A0X-SRZ.
[62] M. Mino, K. Tsukamoto, K. Yanagisawa, A. Tago, and T. Yachi, A
compact buck-converter using a thin-film inductor, in Proc. Appl. Power
Electron. Conf., 1996, vol. 1, pp. 422426.
[63] Q. Li, Y. Dong, F. Lee, and D. Gilham, High-density low-profile coupled inductor design for integrated point-of-load converters, IEEE Trans.
Power Electron., vol. 28, no. 1, pp. 547554, Jan. 2013.
[64] F. C. Lee and Q. Li, High-frequency integrated point-of-load converters:
Overview, IEEE Trans. Power Electron., vol. 28, no. 9, pp. 41274136,
Sep. 2013.
[65] Texas instruments product, TPS 82671, 2011.
[66] F. Sato, T. Ono, N. Wako, S. Arai, T. Ichinose, Y. Oba, S. Kanno, E. Sugawara, M. Yamaguchi, and H. Matsuki, All-in-one package ultracompact
micropower module using thin-film inductor, IEEE Trans. Magn., vol.
40, no. 4, pp. 20292031, Jul. 2004.
[67] S. Sugahara, K. Yamada, M. Edo, T. Sato, and K. Yamasawa, 90% high
Efciency and 100-W/cm3 high power density integrated DCDC converter
for cellular phones, IEEE Trans. Power Electron., vol. 28, no. 4, pp. 1994
2004, Apr. 2013.
[68] Z. Hayashi, Y. Katayama, M. Edo, and H. Nishio, High-efficiency DCDC converter chip size module with integrated soft ferrite, IEEE Trans.
Magn., vol. 39, no. 5, pp. 30683072, Sep. 2003.
[69] H. J. Bergveld, K. Nowak, R. Karadi, S. Iochem, J. Ferreira, S. Ledain,
E. Pieraerts, and M. Pommier, A 65-nm-CMOS 100-MHz 87%-efficient
DC-DC down converter based on dual-die system-in-package integration,
in Proc. IEEE Energy Convers. Congr. Expo., 2009, pp. 36983705.
[70] K. Onizuka, H. Kawaguchi, M. Takamiya, and T. Sakurai, Stacked-chip
implementation of on-chip buck converter for power-aware distributed
power supply systems, in Proc. IEEE Asian Solid-State Circuits Conf.,
2006, pp. 127130.
[71] N. Wang, J. Hannon, R. Foley, K. McCarthy, T. ODonnell, K. Rodgers, F.
Waldron, and C. O Mathuna, Integrated magnetics on silicon for power
supply in package (PSiP) and power supply on chip (PwrSoC), in Proc.
3rd Electron. Syst. Integr. Technol. Conf., 2010, pp. 16.
[72] M. Bathily, B. Allard, and F. Hasbani, A 200-MHz integrated buck
converter with resonant gate drivers for an RF power amplifier, IEEE
Trans. Power Electron., vol. 27, no. 2, pp. 610613, Feb. 2012.
[73] Y. Ahn, H. Nam, and J. Roh, A 50-MHz fully integrated low-swing buck
converter using packaging inductors, IEEE Trans. Power Electron., vol.
27, no. 10, pp. 43474356, Oct. 2012.
[74] J. Wibben and R. Harjani, A high efficiency DC-DC converter using 2nH
on-chip inductors, in Proc. IEEE Symp. VLSI Circuits, 2007, pp. 2223.
[75] M. Wens and M. S. J. Steyaert, A fully integrated CMOS 800-mW fourphase semiconstant on/off-time step-down converter, IEEE Trans. Power
Electron., vol. 26, no. 2, pp. 326333, Feb. 2011.
[76] Y. M. Nguyen, M. Brunet, J.-P. Laur, D. Bourrier, S. Charlot, Z. ValdezNava, V. Bley, and C. Combettes, Low-profile small-size ferrite cores for
powerSiP integrated inductors, in Proc. 15th Eur. Conf. Power Electron.
Appl., 2013, pp. 17.
[77] M. Alimadadi, S. Sheikhaei, G. Lemieux, S. Mirabbasi, W. G. Dunford,
and P. R. Palmer, A fully integrated 660 MHz low-swing energy-recycling

[78]
[79]

[80]

[81]
[82]

5101

DCDC converter, IEEE Trans. Power Electron., vol. 24, no. 6, pp. 1475
1485, Jun. 2009.
J. Ni, Z. Hong, and B. Y. Liu, Improved on-chip components for integrated
DC-DC converters in 0.13 m CMOS, in Proc. ESSCIRC, 2009, pp. 448
451.
M. Mino, T. Yachi, K. Yanagisawa, A. Tago, and K. Tsukamoto, Switching converter using thin film microtransformer with monolithicallyintegrated rectifier diodes, in Proc. Power Electron. Spec. Conf., 1995,
vol. 2, pp. 665670.
K. Yamaguchi, Y. Naitou, O. Nakajima, H. Matsuki, and K. Murakami,
Characteristics of a DC-DC converter using a thin film microtransformer
and a microinductor, IEEE Transl. J. Magn. Jpn., vol. 9, no. 6, pp. 8489,
Nov. 1994.
K. Yamasawa, A DC-DC converter using a microtransformer, IEEE
Transl. J. Magn. Jpn., vol. 9, no. 4, pp. 120126, Jul. 1994.
J. Lu, H. Jia, X. Wang, K. Padmanabhan, W. G. Hurely, and Z. J. Shen,
Modeling, design, and characterization of multiturn bondwire inductors
with ferrite epoxy glob cores for power supply system-on-chip or systemin-package applications, IEEE Trans. Power Electron., vol. 25, no. 8,
pp. 20102017, Aug. 2010.

Ali Moazenzadeh was born in Shiraz, Iran. He


started his academic studies from 2003 in the field of
physics. He received the M.Sc. degree in photonics
from the Laser and Plasma Research Institute, Shahid
Beheshti University, Tehran, Iran, in 2010. In November 2010, he joined the Laboratory for Microactuators
as a Ph.D. student of the Graduate School Embedded
Microsystems, Department of Microsystems Engineering (IMTEK), University of Freiburg, Freiburg,
Germany.
His research interests include the development of
micromagnetic devices rapid manufacturing methods.

Fralett Suarez Sandoval was born in Morelia, Michoacan Mexico, in 1988. She received the B.Eng. degree in electronic engineering from the Morelia Institute of Technology, Morelia, Mexico, in 2011, and the
M.Sc. degree in microsystems engineering from the
University of Freiburg, Freiburg, Germany, in 2013.
Since March 2014, she has been working toward the
Ph.D. degree at the Laboratory for Microactuators,
Department of Microsystems Engineering (IMTEK),
University of Freiburg.
Her research interests include power electronics
and wireless power transmission with magnetic microdevices.

Nils Spengler was born in Berlin, Germany. He received the M.S. degree in microsystems engineering
from the University of Freiburg, Freiburg, Germany,
in 2010. After one-year research at the Palo Alto Research Center (PARC), Palo Alto, CA, USA, he has
been working toward the Ph.D. degree at the Laboratory for Microactuators, Department of Microsystems
Engineering (IMTEK), University of Freiburg, since
2011.

5102

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 30, NO. 9, SEPTEMBER 2015

Vlad Badilita received the B.Sc. and M.Sc. degrees


from the University of Bucharest, Bucharest, Romania, in 1997 and 1999, respectively, and the Ph.D. degree in microoptoelectronics with a thesis focused on
the physics of coupled-cavity surface emitting lasers

from the Ecole


Polytechnique Federale de Lausanne,
Lausanne, Switzerland, in 2004.
In 2007, after two years as a Postdoctoral Research Associate at the University of Maryland, College Park, MD, USA, he joined the University of
Freiburg, Freiburg, Germany, as a Group Leader for
magnetic microsystems. His research interests include the broader area of miniaturized electromagnetic devices with a focus on electromagnetic actuators and
magnetic resonance detectors for sample- and volume-limited samples.

Ulrike Wallrabe received the Ph.D. degree in mechanical engineering of microturbines and micromotors from Karlsruhe University, Karlsruhe, Germany,
in 1992.
From 1989 to 2003, she was with the Institute
for Microstructure Technology, Forschungszentrum
Karlsruhe (today KIT), working on microactuators
and optical microelectromechanical systems. She has
been a Professor of microactuators at the Department
of Microsystems Engineering, IMTEK, University of
Freiburg, Freiburg, Germany, since 2003. In 2010, she
received an internal fellowship at the Freiburg Institute of Advanced Studies,
FRIAS. She has published more than 110 papers in the field of microsystems
technology. Her work focus lies in magnetic microstructures including processes
for magnetic materials and microcoils, in adaptive optics, using piezoactuators
to tune elastic lenses and mirrors, and in microenergy harvesting.

You might also like