Sie sind auf Seite 1von 11

1154

IEEE TRANSACTIONS ON PLASMA SCIENCE, VOL. 39, NO. 4, APRIL 2011

Temporal Evolution of the Ion Fluxes for Various


Elements in HIPIMS Plasma Discharge
Ante Hecimovic and Arutiun P. Ehiasarian

AbstractIn our previously published paper, the life span of


metal ions in high-power impulse magnetron sputtering (HIPIMS)
discharges was measured up to 5 ms from the start of the pulse. To
investigate the influence of the ion mass, ionization energy, and
sputter yield on the time evolution and life span of singly and
doubly charged metal and gas ions in the HIPIMS plasma discharge, the most frequently used materials for thin-film deposition
carbon, aluminium, titanium, chromium, copper, and niobium
have been used. The ion energy distribution function of each
material was measured using energy resolved mass spectrometry
in time-resolved mode. The setup of the mass spectrometer was the
same for all materials. To investigate the influence of working gas
pressure on the time evolution of ion fluxes, measurements have
been performed at two pressures, 0.3 Pa and 3 Pa.
Index TermsPlasma density, plasma properties, plasma transport processes.

I. I NTRODUCTION

N 1999, Kouznetsov et al. [1] suggested a new method


of powering standard circular planar magnetron sources
called high-power impulse magnetron sputtering (HIPIMS)
where a dc power supply would be replaced with pulsed power
supply. Peak power density of 2800 W/cm2 was reported.
Ehiasarian et al. [2], [3] and Machunze et al. [4] used
HIPIMS to deposit hard CrN and TiN coatings with applications in tribological protection of automotive engines and
turning. Konstantinidis et al. [5] report dense zinc-oxide films
deposited using HIPIMS with application as a seed layer in lowemissivity windows.
The main advantage of HIPIMS over dc magnetron sputtering is increased ion-to-neutral metal ratio. Ehiasarian et al. [6]
found that the increase in the ion-to-neutral ratio of Cr and
increase in the ion current density and deposition rate with increase of target current could be linked with the increased electron density in the target vicinity and explained with electron
impact ionization mechanisms. The time and spatial evolution
Manuscript received March 2, 2010; revised August 23, 2010 and
December 20, 2010; accepted December 23, 2010. Date of publication
February 21, 2011; date of current version April 13, 2011. This work was
supported by the Engineering and Physical Sciences Research Council under
Grant EP/D049202/1.
A. Hecimovic is with the Nanotechnology Centre for PVD Research,
Materials and Engineering Research Institute, Sheffield Hallam University,
S1 1WB Sheffield, U.K. and also with the Research Department Plasma with
complex interactions, Ruhr-Universitt Bochum, 44801 Bochum, Germany
(e-mail: Ante.Hecimovic@rub.de).
A. P. Ehiasarian is with the Nanotechnology Centre for PVD Research,
Materials and Engineering Research Institute, Sheffield Hallam University, S1
1WB Sheffield, U.K.
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/TPS.2011.2106516

of the electron density was measured by Bhlmark et al. [7],


measuring electron densities above 1019 m3 .
The ion energy distribution functions (IEDF) of metal and
gas ions in HIPIMS obtained using plasma sampling mass
spectrometry have been reported by various authors [8][12].
During the pulse, a Thompson-like high-energy tail distribution
of metal IEDF with energies up to 100 eV has been reported
[8], [12], whereas after the pulse metal ion IEDF comprised of a
main low energy peak and a high energy tail, for measurements
recorded in the vicinity of the target. Time-resolved measurements comparing the rise of the discharge current to the end of
the pulse have shown the existence of a single low-energy peak
in the post discharge [9], [10]. Several papers used mass spectrometry to describe the influence of pressure on the IEDF in
dc discharges [13] and the influence of peak HIPIMS power on
the IEDF [8], [9]. Increasing the working gas pressure, the highenergy tail of the metal IEDF was reduced and the low-energy
peak of the metal IEDF increased and narrowed as a result of
thermalization through collisions with process gas atoms.
The long life span of Ti neutrals and ions has been investigated, showing that ions and electrons obey ambipolar diffusion
in the postdischarge [14]. The evolution of the electron density
measured at different distances from the target at pressures
from 0.3 to 3 Pa shows that at high pressure, the electron
density decays slowly at all distances compared to a sharp drop
in density after switch off at low pressure [15]. In addition,
the measurements showed delayed arrival of peak electron
density at longer distances and reduction of the propagation
velocity of the electron density peak with increased pressure.
The results obtained with a Langmuir probe on the dependence
of the electron density temporal evolution on the discharge
power [16], show the increase in electron density with discharge
power and delayed arrival of the electron density peak at longer
distances from the target. Furthermore, Langmuir probe was
used to investigate the evolution of the metal ion density as a
function of time and distance and as a function of time and
pressure at 1.1 Pa [17]. The results show two waves of ions, first
one arriving shortly after the end of the pulse and second wave
arriving in the postdischarge, reaching a peak at later times for
longer distances from the target.
The ionization mechanisms during HIPIMS evolution are
governed by electron impact collision processes and depend
strongly on the electron energy distribution function (EEDF).
The Langmuir probe measurements of the EEDF in a similar system [18]a pulsed magnetic multipole hydrogen
dischargeshowed that it changed from a non-Maxwellian distribution with high energy during the discharge to a Maxwellian
distribution in the late postdischarge. Another similar

0093-3813/$26.00 2011 IEEE

HECIMOVIC AND EHIASARIAN: TEMPORAL EVOLUTION OF ION FLUX IN HIPIMS DISCHARGE

discharge system, multicusp ion source, was investigated experimentally and theoretically by Bretagne et al. [28] showing
a bi-Maxwellian electron energy distribution. We use the term
Maxwellian distribution under the assumption that the plasma
is in local thermal equilibrium. Similar results were reported
in the HIPIMS discharge [19]. The work of Vetushka and
Ehiasarian [20] shows a high-energy group during the ignition
phase of the discharge and a low-energy electron group at the
end. Here, it must be pointed out that all reports on the EEDF
in the HIPIMS discharge are close to a Maxwellian distribution
only at low energies of a few electron volts, while at higher
energies, the distribution is not Maxwellian and these higher
energy electrons are very important in terms of ionization of
gas and metal particles.
The ion saturation current measured at different distances
from the target [21] showed that ions are able to leave the
magnetized region and reach the substrate in greater numbers
depending on the peak current, which is the consequence of
the kinetic pressure exceeding the magnetic pressure. Previous
work by the present authors has suggested [9] that thermalization of the metal ions occurs in the HIPIMS postdischarge.
Simulations of particle transport [22] and the neutral particle
flow in HIPIMS [23] add valuable information on plasma
thermalization and gas rarefaction.
Detailed investigation of the metal ion life- pan for different
target materials has not been reported yet. In this paper, we
investigate the life span of the most frequently used materials
for thin-film deposition carbon (C), aluminium (Al), titanium
(Ti), chromium (Cr), copper (Cu), and niobium (Nb), and the
influence of intrinsic properties on the time evolution of the ion
flux in the HIPIMS plasma discharge.

1155

Fig. 1. UHV machine used for HIPIMS plasma measurements.

II. E XPERIMENTAL D ETAILS


HIPIMS plasma discharge was generated in a CMS
18 ultrahigh vacuum (UHV) system (Kurt J. Lesker), utilizing
a planar unbalanced magnetron with diameter of 75 mm.
The pulse length of the HIPIMS plasma discharge was 70 s
and the repetition frequency was 100 Hz. The peak discharge
current was kept constant for all measurements at 40 A, that is
equivalent to a current density of 1 A/cm2 . The measurements
have been performed at two pressures, 0.3 and 3 Pa. The
pressure of working gas Ar was measured using an MKS SRG2CE spinning rotor gauge. The target of the magnetron for each
measurement of different material was changed. Six elements
were investigated: C, Al, Ti, Cr, Cu, and Nb, and the target of
each element was made with purity of 99.9%.
A schematic cross section of the UHV machine used for
plasma measurements is given in Fig. 1. It shows the position
of the substrate and the plasma sampling mass spectrometer in
relation to the target. The spectrometer was placed at an angle
of 58 with respect to target normal and the distance from the
orifice of the spectrometer to the target was 17 cm.
Results presented in Section III-A and B are measured using
time-resolved mode of the mass spectrometer to show the
temporal evolution of the ion flux for different elements. The
details of the time-resolved measurements can be found in
the previous publication [26]. Time-averaged measurements

Fig. 2. Discharge current and voltage in HIPIMS for C, Al, Ti, Cr, Cu, and
Nb at low pressure of 0.3 Pa.

presented in this paper were recorded at 30 and 70 s in order


to record the IEDF inside the pulse, followed by measurements
at 100, 150, 200, 300, and 400 s to observe the IEDFs in the
postdischarge shortly after the end of the pulse. To observe the
life span of ions in the late postdischarge, IEDF was measured
at 800 s, 1.2, 2.5, 5, and 9.9 ms. Each measurement was
recorded with 20 s gate width and averaged over 50 pulses.
Results presented in Section IV are measured in the timeaveraged mode of the mass spectrometer. In the time-averaged
mode, the detector of the mass spectrometer records the signal
for 300 ms, that is for 30 pulses in our measurements.

1156

IEEE TRANSACTIONS ON PLASMA SCIENCE, VOL. 39, NO. 4, APRIL 2011

Fig. 4. Time evolution of the single-charged ion flux of different elements at


pressure of 0.3 Pa.

Fig. 3. Discharge current and voltage in HIPIMS for C, Al, Ti, Cr, Cu, and
Nb at high pressure of 3 Pa.

The target discharge current for each element are shown on


Fig. 2 for low-pressure measurements and on Fig. 3 for highpressure measurements. The shape of the discharge voltage is
the same for all elements as it is regulated by the power supply
but the amplitude is different, therefore the power on the target
is different as shown Fig. 5. The shape and the ignition time
of the discharge current differ for each element. For most of
the materials, the current starts to rise around 20 s from the
start of the pulse except for Ti, where the discharge started
with delay of 40 s from the start of the pulse. Time-resolved
measurements obtained using optical emission spectroscopy by
Ehiasarian et al. [6], [24] showed that the discharge transforms
from gas plasma phase to a metal plasma phase. Short time
of the HIPIMS pulse with Ti target could lead to lower metal
density due to shorter time between end of the gas plasma phase
and end of the pulse. Consequently, it is possible to expect
lower intensity of Ti ion flux. The discharge current shape can
be described as a triangle with exponential rise of the current
until 70 s, when the voltage pulse is terminated and the current
reaches its peak value of 40 A. After the end of the pulse, the
current linearly reduces to zero within 15 s.
The shape of the discharge current for the high-pressure
measurements differs from the low-pressure measurements.
The rise of the current occurs within 10 s from the start of
the pulse. For Al, Cr, Cu, and Nb target materials, the shape
of the discharge current indicated that saturation has occurred.
After an initial rise up to value of 40 A, the current saturates

Fig. 5. Dependence of the peak intensity of the ion flux on the energy
introduced to the system during the pulse period at pressure of 0.3 Pa.

until the end of the pulse when it is reduced linearly to zero.


For C and Ti, the discharge current has a triangular shape as at
low-pressure measurements.
III. R ESULTS
A. Temporal Evolution of Metal and Gas Ion Fluxes of
Various Elements at Pressure of 0.3 Pa
1) Singly Charged Ions of Various Elements: The time evolution of the singly charged ion fluxes for different materials at
pressure of 0.3 Pa is presented on Fig. 4. The x-axis of Fig. 4
represents the time between pulses of 10 ms and the y-axis is
the ion flux calculated from the integral of the IEDF. For all
elements, the peak in ion flux intensity is detected at 150 s
from the start of the pulse, i.e., 80 s after the end of the pulse.
The highest intensity of the ion flux is measured for Cr and Nb.
The peak intensity can be correlated to the integral of the power
during the pulse, i.e., energy delivered to the system as shown
on Fig. 5.
The maximum ion flux, ion flux at 5 ms, and the energy
introduced into the system as a function of atomic mass are
plotted on Fig. 5. Figure shows an increase of the maximum ion

HECIMOVIC AND EHIASARIAN: TEMPORAL EVOLUTION OF ION FLUX IN HIPIMS DISCHARGE

Fig. 6.

Average energy of element ions at pressure of 0.3 Pa.

flux with delivered energy. This effect seems to be independent


of the element.
The energy delivered into the system is calculated as the
integral of power with respect to time from the start of one
pulse to the beginning of the consequent pulse. Higher ion flux
for plasma discharge of element with high energy delivered into
the system stems from two processes. First process is correlated
with increased amount of sputtered material with increased
energy delivered to the system and consequently higher density
of secondary electrons. The energy delivered in the system
is primarily distributed to electrons, ions and neutrals in the
plasma, therefore higher delivered energy produces electrons
with higher kinetic energy which lead to increased ionization,
e.g., higher ion fluxes.
Fig. 4 shows that from 150 s up to 1 ms, the reduction of
the ion flux intensity for all elements is exponential in time
and very sharp, whereas after 1 ms, the decrease of intensity is
exponential at slower rate. After the initial peak, the ion fluxes
decrease up to a time of 5 ms when the last signal is detected for
all elements. A notable exception is Nb for which a small signal
of metal ions is detected at 9.9 ms. The ion flux at 5 ms plotted
on Fig. 5, compared with the maximum ion flux, shows that the
life span of the elements depend on the mass of the element
since the ion flux of the light elements is reduced at 5 ms i.e.,
the life span of light elements is shorter than the life span of the
heavy elements.
Fig. 6 represents the average energy of ions as a function of
time. Average energy is calculated as a ratio of sum of product
of the energy and signal intensity (counts/s) over sum of the
signal intensity (counts/s)

E Intensity
.
Eav = 
Intensity
It is calculated for each IEDF separately measured at different times as described in Section II. Average energies of
up to 8 eV have been detected at the end and shortly after
the pulse at 70 and 100 s. Such high average energies of
ions originate from particles sputtered from the target with
Thompson distribution [25]. After the end of the pulse, the
flux of sputtered material is stopped and metal ions are losing
energy through collisions with gas atoms and ions. Gas particles

1157

Fig. 7. Time evolution of the doubly charged ion flux of different elements at
pressure of 0.3 Pa.

have thermal energy equivalent to room temperature. After


400 s, all metal ions are thermalized, as shown in our previous
publication [26], and have an average energy below 1 eV that
is equivalent to the plasma potential of 1 V as measured at a
pressure of 0.3 Pa in the postdischarge at 400 s from the start
of the pulse [9].
2) Doubly Charged Ions of Various Elements: The temporal
evolution of doubly charged ions of various elements is shown
on Fig. 7. The trends are similar to those of single-charged ions
presented on Fig. 4. The ion flux of all elements has a peak at
150 s followed by an exponential decrease in time. The life
span of doubly charged ions for C, Al, and Cu is around 1 ms.
In HIPIMS of Cr and Ti discharges, the life span is 2.5 ms and
for Nb discharge, the doubly charged ions are detected up to
5 ms. Long life span of Nb, Cr, and Ti ions could be due to
high intensity of ion fluxes detected shortly after the end of the
pulse. The highest intensity of the ion flux is detected for Nb,
Cr, and Ti elements. The high amount of doubly charged ions
detected after the end of the pulse for these elements is due
to heavy mass of ions and due to low ionization potential of
singly charged ions, which is 13.58 eV for Ti, 14.32 eV for Nb,
and 16.5 eV for Cr. Furthermore, Cr and Nb have the highest
quantity of singly charged ions in the same period as shown on
Fig. 4. The lack of Cu2+ ions could be explained by their high
second ionization potential which is 20.3 eV.
3) Singly Charged Ar Ions: In addition to singly and doubly
charged ions of various elements, the singly charged ions of
Ar were measured during the discharge of each element and
the results are shown on Fig. 8. The temporal evolution of Ar
ions is similar to the temporal evolution of metal ions on Fig. 4.
Peak intensity of the Ar ions flux is reached shortly after the
end of the pulse at 200 s. After the peak, the ion flux of Ar
ions reduces exponentially in time until 5 ms when the last
Ar ions are detected. At the last measuring point at 9.9 ms,
only a small signal was detected. Highest intensity of Ar ion
flux is detected for Nb, C, and Cr discharges. The metal flux in
HIPIMS discharges of Nb and Cr is higher compared to other
materials, as shown on Fig. 5, due to higher energy introduced
into the system, in turn producing more secondary electrons that
ionize gas atoms, explaining the high intensity of gas ions. The

1158

Fig. 8. Temporal evolution of the Ar ion flux at HIPIMS discharge of different


elements at pressure of 0.3 Pa.

IEEE TRANSACTIONS ON PLASMA SCIENCE, VOL. 39, NO. 4, APRIL 2011

Fig. 10. Time evolution of the single-charged ion flux of different elements at
pressure of 3 Pa.
TABLE I
E NERGY I NTRODUCED I NTO THE S YSTEM C ALCULATED AS I NTEGRAL OF
P OWER I NTEGRATED F ROM 0 TO 70 s, I . E ., D URING THE
P ULSE AT P RESSURE OF 3 Pa

Fig. 9. Average energies of the Ar ions for HIPIMS plasma discharges of


various elements at pressure of 0.3 Pa.

high amounts of gas ions in the C discharge could be explained


by the high ionisation potential of C atoms of 11.26 eV. As C
atoms are sputtered from the target into the plasma, the plasma
electrons are cooled less compared to metals and the ionization
of Ar is more efficient, producing higher number of Ar ions.
Fig. 9 shows the average energy of Ar ions for different
target materials. The figure shows an increase in average energy
during the pulse and peak at 100 s. Increased average energy
above values of plasma potential are due to collisions with hot
metal particles ejected from the target. After termination of the
pulse, the Ar ions lose their kinetic energy in charge exchange
collisions with thermalized gas particles within next 100 s.
After 200 s, Ar ions for all discharges are below 1 eV which
is equal to measured plasma potential [9]. Average energy of
6.8 eV and 4.7 eV are calculated in plasma discharges of Nb and
Cu. The kinetic energy transferred to Ar ions in elastic collision
with heavy Nb and Cu ions and atoms is significant and
therefore the average energy of Ar ions in mentioned discharges
is high. Mass spectrometer is placed at the distance of 17 cm
and mean free path of ions is around 7 cm, therefore Ar ion will,
in average, collide between two and three times before reaching
the entrance of the mass spectrometer, resulting in increased
average energy. The lowest average energy is detected in C

discharge and following the same argument as in case of heavy


elements, the kinetic energy transferred to Ar ions in collision
with light C particles is insignificant. Maximum energy transferred from a C particle to an Ar particle in an elastic collision,
assuming a billiard balls model, is only 35% of the C particle
kinetic energy, compared to above 60% for all other elements.
B. Temporal Evolution of Metal and Gas Ion Fluxes of
Various Elements at Pressure of 3 Pa
1) Singly Charged Ions of Various Elements: The measurements presented in previous publication [26] show the life span
of metal ions up to the start of consequent pulse at distances
above 10 cm at pressure of 3 Pa. The measurements presented
on Fig. 10 are measured at distance of 17 cm from the target and
at pressure of 3 Pa, therefore long life spans are expected. The
x-axis represents the timescale from the start of the pulse until
the start of the following pulse at 10 ms and y-axis is the ion flux
calculated by integrating the IEDF of each element. The ion flux
scale on Figs. 4 and 10 is the same to ease comparison. Contrary
to the low-pressure case, the high-pressure measurements show
existence of two peaks, first one occurring at the end of the
pulse and second one at around 1.2 ms from the start of the
pulse.
At the end of the pulse, the highest intensity is detected for
elements Cu and Nb due to heavy atomic mass and high amount
of energy introduced into the system during the pulse. The
element-to-Ar mass ratio is 2.3 and 1.6 for Nb and Cu ions, respectively, and means that due to heavy mass, the change of the
ion trajectory angle is small in elastic collisions with gas atoms

HECIMOVIC AND EHIASARIAN: TEMPORAL EVOLUTION OF ION FLUX IN HIPIMS DISCHARGE

1159

TABLE II
I NTENSITY OF THE F IRST AND S ECOND P EAK FOR E ACH E LEMENT AND R ELATIVE C HANGE IN
I NTENSITY OF THE S ECOND P EAK C OMPARED TO F IRST P EAK AT P RESSURE OF 3 Pa

and ions, and more ions are reaching the mass spectrometer
orifice at given time. Furthermore, the energy introduced into
the system, presented in Table I, has a high value for Nb and
Cu discharges. During the HIPIMS discharge with Al target,
the energy introduced into the system is also high but the peak
value of ion flux is lower compared to Nb and Cu due to the
low mass of Al ions. Al-to-Ar mass ratio is 0.68 and the change
of the Al ion trajectory angle, in other wordsscattering,
is significant in the elastic collisions with gas particles.
A second peak is detected for all elements and it is highly
reproducible. For all elements, it appears at 1.2 ms except Ti
discharge where the peak is detected at 2.5 ms. The intensity
of second peak for light elements is higher than first peak. In
Table II, the intensity of first and second peak for each element
and the ratio of first peak intensity to second peak intensity
is given. It is clear that for light elements, the second peak
intensity is increased whereas it is reduced for heavier elements.
The first peak comprises of hot and fast ions created during
the pulse that fly directly to the mass spectrometer and due
to heavy mass, high amount of Cu and Nb ions reaches the
mass spectrometer. Masses of Cu and Nb ions are 1.6 and
2.3 times the mass of Ar and since the scattering angle is
linearly dependant on the mass ratio scattering of Cu and Nb
ions will be much smaller compared to lighter elements.
Second peak comprises of thermalized ions that are diffusing
through the chamber, obeying ambipolar diffusion [14] and
arriving 1.2 ms from the start of the pulse to the mass spectrometer. Light elements C and Al lose more energy in collisions
with Ar atoms and ions during the pulse and therefore the ion
flux reaching mass spectrometer is reduced in the peak after the
end of the pulse. The thermalized ions diffuse slowly reaching
the mass spectrometer 1.2 ms from the start of the pulse in
greater numbers compared to first peak. On the other hand, the
number of thermalized heavy element ions during the pulse is
smaller since their scattering was not restricted by Ar particles,
resulting in the ion flux at 1.2 ms being reduced compared to
the first peak.
Slow diffusion of metal ions at pressure of 3 Pa leads to a
long life span until the start of consequent pulse for all elements
except C. The ion flux of elements at 9.9 ms is around 10% of
maximum value for each element, indicating a reasonably high
amount of ions impinging on the substrate at high pressure.
Fig. 11 shows the average energy of IEDF for each element
as a function of time. Only first millisecond is shown since
no change in average energy occurs after that time. Somewhat
surprisingly, the highest average energy is detected for C ions,
followed by heavy elements Cu and Nb. Due to their mass,

Fig. 11. Average energy of singly charged ion flux of different elements at
pressure of 3 Pa.

Fig. 12. Peak average energy plotted as a function of element mass to Ar


atomic mass ratio at pressure of 3 Pa.

heavy elements lose less kinetic energy in collisions with thermalized gas particles and this could explain somewhat higher
average energy. Lowest average energy is measured for Ti, Al,
and Cr, and it could be that elements with mass similar to Ar
lose more kinetic energy in collisions with Ar. The dependence
of the peak average energy on element to Ar mass ratio is
plotted on Fig. 12. It shows that elements that have similar mass
to the mass of cold Ar gas can only maintain a lower average
energy while heavy elements Cu and Nb retain higher average
energy.
The high average energy of C ions could be explained by
low-energy transfer factor to Ar particles as mentioned in

1160

Fig. 13. Temporal evolution of the doubly charged ion flux of different
elements at pressure of 3 Pa.

Fig. 14. Average energy of doubly charged ions at pressure of 3 Pa.

Section III-A3. The collision cross section of two hard spheres


is proportional to the square of the sum of the colliding particle
radii and radius of C ion is 0.16 compared to more than
50 for other elements ions. Therefore, the C ions maintain
its high initial energy of sputtering, much better the rest of the
atoms considered due to its small size and smaller collision
cross section.
2) Doubly Charged Ions of Various Elements: The flux intensity of singly charged ions at high pressure is two orders
of magnitude lower compared to low-pressure measurements.
The intensity of doubly charged ion flux is also two orders of
magnitude lower. Therefore, at the high pressure, only a weak
signal of Nb, Cr, and Ti ions has been detected, shown on
Fig. 13 Peak of the ion flux for all three elements is detected
at 150 s from the start of the pulse. After the peak Nb2+ ,
ion flux has a shoulder similar to Nb1+ ion flux but at lower
intensity and with life span until 2 ms. Second peak of Cr2+
ion flux peaks at 0.8 ms and then falls to zero after 1.2 ms.
The peak of Ti2+ ions is detected with life span of only 200 s.
Longer life span ions are below the detection limit of the
spectrometer. The peak of doubly charged ions consists of fast
ions reaching the mass spectrometer shortly after the end of the
pulse. The average energies of doubly charged ions are plotted
as a function of time on Fig. 14. The average energies are lower

IEEE TRANSACTIONS ON PLASMA SCIENCE, VOL. 39, NO. 4, APRIL 2011

Fig. 15. Temporal evolution of the Ar ion flux at HIPIMS discharge of various
elements at pressure of 3 Pa.

compared to low-pressure measurements since the energy of


the metal ions is lost in elastic collisions with the surrounding
thermalized gas particles due to short mean free path of 0.5 cm.
The shoulder of Nb2+ and re-appearance of Cr2+ ions area
result of thermalized plasma reaching the mass spectrometer.
As shown on Fig. 14, after 400 s, all doubly charged ions
are thermalized with surrounding gas. After 1.2 ms, no doubly
charged ion flux is detected at high pressure probably as a result
of doubly charged ions being lost to the walls.
3) Singly Charged Ar Ions: The measurement of Ar ion
flux for various target materials at high pressure is shown on
Fig. 15. The ion flux of Ar ions has a peak shortly after the
end of the pulse at 100 s. After the peak, in HIPIMS plasma
discharges of Cu and Cr, Ar ion flux decays exponentially in
time. For other materials, the ion flux decays exponentially for
a short period of time until 300 s when the decay is halted.
The first peak observed during the time discharge comprises
of Ar ions ionized by electrons produced during the HIPIMS
pulse. In previously published work, it has been shown that Ar
ions are detected at all distances early in the pulse [26]. These
ions are created by electron impact ionization of Ar atoms at
distances of up to 20 cm from the target by secondary electrons
created at the target during the pulse. The Ar ion flux is almost
constant from 300 s until 2.5 ms after which the exponential
decay continues. This continuous influx of Ar ions arriving to
the mass spectrometer between 300 s and 2.5 ms coincides in
time with second peak of the ion flux of various elements shown
on Fig. 10.
The highest density of Ar ions after the pulse is detected
for HIPIMS plasma discharge of C. The high amounts of gas
ions in Carbon discharge could be explained by high ionization
potential of C atoms. As C atoms are sputtered from the target
into the plasma, the plasma electrons are cooled less compared
to metals and the ionization of Ar is more efficient producing
higher number of Ar ions. Second and third highest density of
Ar ions detected after the pulse is for the HIPIMS discharges of
Al and Nb. During Al and Nb discharges, the high amount of
energy is introduced into the system during the pulse, calculated
as integral of power during the pulse and presented in Table I.
This high energy is transferred to all particles in the system and
increased energy and number of electrons could lead to higher

HECIMOVIC AND EHIASARIAN: TEMPORAL EVOLUTION OF ION FLUX IN HIPIMS DISCHARGE

1161

TABLE III
R ATIO OF THE F IRST P EAK AND S HOULDER I NTENSITY OF THE Ar I ON
F LUX IN A R ELATION TO A S PUTTER Y IELD OF E LEMENT

amounts of Ar ions detected in the discharges with Al and Nb


target.
The shoulder on the Ar ion flux in the postdischarge comprises of Ar ions arriving to the mass spectrometer together with
the bulk metal plasma. After 2.5 ms, the density of ions in the
bulk plasma reduces due to losses to the wall and the density
of Ar ions is reduced accordingly. In the HIPIMS plasma discharges of Cu and Cr, the shoulder of Ar ion flux does not exist.
Cu and Cr are materials with high sputter yield, 2.59 (Ar incident energy of 600 eV) and 1.55 (Ar incident energy of 530 eV),
respectively, and the gas rarefaction in the target vicinity is
emphasized. The bulk plasma reaching the mass spectrometer
at 1.2 ms is mostly composed of metal ions, 80% and 92%
of the ions flux are metal ions for discharge of Cr and Cu,
respectively. For this reason, the life span of Ar ion flux in Cu
and Cr HIPIMS plasma discharges is up to 5 ms and for other
materials, it is up to 9.9 ms before the start of consequent pulse.
In Table III, the ratio of the first peak and shoulder intensity
of the Ar ion flux in a relation to a sputter yield of element
is presented. For high sputter yield elements Cu and Cr, the
shoulder has below 8% of intensity while for elements with
lower sputter yield Ti, Al, and Nb, the ratio is between 0.18 and
0. 26. For elements with lower sputter yield, gas rarefaction is
less emphasized and both Ar and metal ions can be found in
bulk plasma arriving around 1.2 ms to the mass spectrometer.
C has a lowest sputter yield of 0.12 and the ratio is only 0.11;
however, it can be excluded from this discussion due to high
intensity of Ar ion flux created during the pulse due to higher
efficiency of Ar ionization as mentioned in previous section.
In Fig. 16, the average energy of Ar ion flux is presented. The
average energy for all elements is below 1.2 eV indicating high
thermalization of Ar ions in the plasma.
IV. D ISCUSSION
A. Influence of Gas Pressure on the Plasma Properties
The results have demonstrated that pressure has a significant
effect on the energy and life span of metal and gas ions.
Temporal evolution of the metal ion flux at low pressure of
0.3 Pa exhibits a strong peak detected at 150 s followed by
exponential decay. Intensity of the peak strongly depends on
the energy delivered into the system during the pulse. Life- pan
of the metal ions shows dependence on the mass of the element
such that ions with heavy mass have longer life spans compared
to light elements.
The flux of Ar ions has similar temporal evolution as metal
ions with peak intensity being detected shortly after the end of
the pulse, followed by exponential decay with average energies

Fig. 16. Average energies of Ar ions for HIPIMS plasma discharges of various
elements at pressure of 3 Pa.

of Ar ions, having strong correlation on the mass of the element


used in HIPIMS plasma discharge. Elements that have bigger
mass than Ar ions transfer more kinetic energy to the gas ions
in the elastic collisions, increasing the average energy of Ar
ions during the pulse.
Temporal evolution of metal ions at high pressure of 3 Pa
has two peaks, first one comprises of ballistic ions reaching the
mass spectrometer and second peak comprises of thermalized
bulk plasma. Comparing the peak intensity of first and second
peak, it has been shown that majority of the sputtered particles
of the heavy elements arrives to the mass spectrometer shortly
after the pulse whereas for light elements the bulk of the
sputtered metal arrives between 1 and 2 ms from the start of
the pulse. Furthermore, the average energy of metal ions was
lower for ions of the elements that have mass comparable to
the mass of Ar ions and it was higher for light C and heavy Cu
and Nb, showing that kinetic energy transfer is most effective
when mass of sputtered particles and mass of the working gas
are comparable.
Temporal evolution of the Ar ion flux at high pressure of
3 Pa is similar to the metal ion flux at the same pressure with
the exception of second peak, that is in case of Ar ion flux
rather than a shoulder. It has been shown that the intensity of
the shoulder is dependent on the sputter yield of the element
since ratio of the first and second peak showed lower second
peak for the elements with higher sputter yield since gas
rarefaction in discharge of high-sputter-yield elements, i.e., Cu,
is emphasized, therefore density of Ar ions is reduced in the
target vicinity, resulting in lower ion flux at 1 ms.
B. Influence of Properties of Sputtered Elements on the
Plasma Properties
The nature of the sputtered material seems to be an overbearing factor that determines the metal-to-gas ion ratio as well
as the ion flux. The trends are preserved regardless of the gas
pressure.
Fig. 17 shows the metal-to-gas-ion ratio extracted from the
integral values of the metal and ion fluxes. At both low and high

1162

IEEE TRANSACTIONS ON PLASMA SCIENCE, VOL. 39, NO. 4, APRIL 2011

Fig. 17. Metal-to-gas-ion ratio as a function of sputtering yield. The measurements were done for two pressures 0.3 and 3 Pa.

pressure, the metal-ion-to-gas-ion ratio increases with sputter


yield of the element. Increase in metal-to-gas-ion ratio with increase of the sputter yield can be explained by two effects. First
effect is higher density of sputtered material during HIPIMS
plasma discharge of elements with higher sputter yield. Second
effect is higher gas rarefaction in the target vicinity due to
higher density of material sputtered from the target for elements
with high sputter yield. First effect increases metal ion density
and second effect reduces gas ion density resulting in increased
metal to gas ion ratio. In Fig. 17, the increase is constant
for high-pressure measurements; however, at low pressure, the
metal-to-gas-ion ratio increases for Cr (sputter yield is 1.18)
and then is reduced to a value of 4 for Cu (sputter yield is 2.35).
The reason for such behavior could be an associated decrease
in power for Cu and high intensity of Cr ions proportional to
higher energy introduced into the system during the pulse.
In Section III-A1, it has been argued that high average
energy of elements singly and doubly charged ions in plasma is
increased due to the Thompson distribution of the particles. On
Fig. 18, temporal evolution of Ti+ , Ti2+ , and Ar+ ion average
energy and plasma potential measured using the Langmuir
probe at distance of 15 cm in Ti HIPIMS discharge is showed.
Fig. 18(a) shows the data measured at pressure of 0.3 Pa,
showing that during the pulse, the average energy has value
similar to the value of the plasma potential. After the end of
the pulse, plasma potential has a sharp drop to values around
1 V whereas average energy decays at slower rate reaching
the value of 1 eV after 300 s after the start of the pulse.
The peak of the IEDF is shifted on the energy scale according
to change of the plasma potential. Nevertheless, the average
energy has values higher than the plasma potential due to nonnegligible number of ions with energy higher than peak value,
comprising of high-energy tail of the IEDF originating from
Thompson distribution of sputtered atoms. Fig. 18(b) shows the
data measured at pressure of 3 Pa, showing similar temporal
evolution of plasma potential and average energy. The particle
reaching the mass spectrometer orifice at pressure of 3 Pa will
collide, in average, more than 20 times, therefore majority of
ions will be thermalized.
The average energy of ions for different elements from
time-averaged measurements as a function of atomic mass is

Fig. 18. Temporal evolution of Ti+ , Ti2+ , and Ar+ ion average energy and
plasma potential measured using the Langmuir probe at pressure of (a) 0.3 and
(b) 3 Pa.

Fig. 19. Average energy of singly charged ions of the various elements as a
function of element atomic mass number at pressure of 0.3 and 3 Pa.

presented on Fig. 19 at pressure of 0.3 Pa. The increase in


average energy of ions with increase in atomic mass of the
element shows that the effect of thermalization is stronger for
lighter particles.
Fig. 19 shows the average energy of ions for different elements as a function of atomic mass at high pressure of 3 Pa.
On this figure, the average energy is decreasing from 0.8 to
0.58 eV as the atomic mass increases from 12 amu for C to
93 amu for Nb. The decrease in average energy for heavier
elements could be due to shorter mean free path for larger
particles. The mean free path is reversely proportional to the
product of gas pressure and elastic collision cross section =
1/(ng ) [27]. The elastic collision cross section is proportional

HECIMOVIC AND EHIASARIAN: TEMPORAL EVOLUTION OF ION FLUX IN HIPIMS DISCHARGE

Fig. 20. Average energy of singly charged ions of the various elements as a
function of ion radius at pressure of 3 Pa.

to the radius of the ion, therefore bigger ions have shorter mean
free path and thermalization is enhanced.
The average energy as a function of ion radius is presented on
Fig. 20 and it shows a decrease in average energy with increase
in ion radius. This effect has less impact on the ions at the low
pressure since at low pressure, the mean free path is ten times
longer and it is around 7 cm, therefore, on average, an ion will
collide twothree times before entering the mass spectrometer
at 17 cm away from the target. At high pressure of 3 Pa, the
ion has an average of 2030 collisions before entering the mass
spectrometer. At low pressure due to few collisions, lighter
elements will lose more kinetic energy than heavy particles
due to their cross section, whereas at high pressure, cross
section of the ion plays a bigger role than at low pressure and
consequently heavy ions lose more energy.
The effect of the binding energy of the element on the ion
density is shown on Fig. 21 for the low pressure measurements.
The high-pressure measurements are presented on Fig. 22. The
ion density is calculated as an integral of the IEDF measured in
time averaged mode. The particles are sputtered from the target
with Thompson distribution [25]. Thompson distribution has a
dependence on the binding energy Eb as fT 1/Eb2 + 1/Eb .
The shape of the curves on Figs. 21 and 22 match the dependence of the Thompson distribution on the binding energy.
The elements with low binding energy Cu and Al have higher
intensities and the elements with higher binding energy such as
C and Nb have lower intensities. In Figs. 21 and 22, calculated
integral of the Thomson distribution as a function of binding
energy is plotted for comparison.

1163

Fig. 21. Ion flux of elements as a function of binding energy measured at low
pressure of 0.3 Pa.

Fig. 22. Ion flux of elements as a function of binding energy measured at high
pressure of 3 Pa.

Metal-to-gas ion ratio increased for elements with higher


sputter yield. The average energy measured in time-resolved
mode increased for elements that have significantly different
mass than Ar, indicating that energy losses are highest when
masses are similar. At high pressure, the average energy of
singly charged ions reduced with heavier mass due to shorter
mean free path that is reversely proportional to the size of the
ions.
The metal ion flux showed dependence on binding energy of
the element as predicted by Thompson distribution function.

V. C ONCLUSION

R EFERENCES

The ion energy and life span of ions were found to depend
strongly on collisions of the sputtered flux with the surrounding
gas and as such depended on the target material. Life spans are
longest for ions of the heavy mass elements due to rarefaction. Furthermore, doubly charged ions are detected only for
elements with mass heavier than Ar at both pressures.
At high pressure, a second peak for singly charged metal ions
around 1.2 ms has been detected that could be the signal of bulk
plasma slowly diffusing away from the target.

[1] V. Kouznetsov, K. Macak, J. M. Schneider, U. Helmersson, and I. Petrov,


A novel pulsed magnetron sputter technique utilizing very high target
power densities, Surf. Coat. Technol., vol. 122, no. 2/3, pp. 290293,
Dec. 1999.
[2] A. P. Ehiasarian, P. E. Hovsepian, L. Hultman, and U. Helmersson,
Comparison of microstructure and mechanical properties of chromium
nitride-based coatings deposited by high power impulse magnetron sputtering and by the combined steered cathodic arc/unbalanced magnetron
technique, Thin Solid Films, vol. 457, no. 2, pp. 270277, Jun. 2004.
[3] A. P. Ehiasarian, W.-. Munz, L. Hultman, U. Helmersson, and
I. Petrov, High power pulsed magnetron sputtered CrNx films, Surf.
Coat. Technol., vol. 163/164, pp. 267272, Jan. 2003.

1164

[4] R. Machunze, A. P. Ehiasarian, F. D. Tichelaar, and G. C. A. M. Janssen,


Stress and texture in HIPIMS TiN thin films, Thin Solid Films, vol. 518,
no. 5, pp. 15611565, Dec. 2009.
[5] S. Konstantinidis, A. Hemberg, J. P. Dauchot, and M. Hecq, Deposition
of zinc oxide layers by high-power impulse magnetron sputtering,
J. Vac. Sci. Technol. B, Microelectron. Nanometer Struct., vol. 25, no. 3,
pp. 1921, May 2007.
[6] A. P. Ehiasarian, R. New, W. Munz, L. Hultman, U. Helmersson, and
V. Kouznetsov, Influence of high power densities on the composition
of pulsed magnetron plasmas, Vacuum, vol. 65, no. 2, pp. 147154,
Apr. 2002.
[7] J. Bohlmark, J. T. Gudmundsson, J. Alami, M. Latteman, and
U. Helmersson, Spatial electron density distribution in a high-power
pulsed magnetron discharge, IEEE Trans. Plasma Sci., vol. 33, no. 2,
pp. 346347, Apr. 2005.
[8] J. Bohlmark, M. Lattemann, J. T. Gudmundsson, A. P. Ehiasarian,
Y. A. Gonzalvo, N. Brenning, and U. Helmersson, The ion energy distributions and ion flux composition from a high power impulse magnetron
sputtering discharge, Thin Solid Films, vol. 515, no. 4, pp. 15221526,
Dec. 2006.
[9] A. Hecimovic, K. Burcalova, and A. P. Ehiasarian, Origins of ion energy
distribution function (IEDF) in high power impulse magnetron sputtering
(HIPIMS) plasma discharge, J. Phys. D, Appl. Phys., vol. 41, no. 9,
p. 095203, May 2008.
[10] J. Vlcek, A. D. Pajdarova, and J. Musil, Pulsed dc magnetron discharges
and their utilization in plasma surface engineering, Contrib. Plasma
Phys., vol. 44, no. 5/6, pp. 426436, Sep. 2004.
[11] J. Vlcek, P. Kudlacek, K. Burcalova, and J. Musil, Ion flux characteristics
in high-power pulsed magnetron sputtering discharges, Europhys. Lett.,
vol. 77, no. 4, p. 5, Feb. 2007.
[12] J. Bohlmark, A. P. Ehiasarian, M. Lattemann, J. Alami, and
U. Helmersson, The ion energy distributions in a high power impulse
magnetron plasma, in Proc. 48th Annu. Tech. Conf. SVC, Denver, CO,
2005, pp. 470473.
[13] S. Kadlec, C. Quaeyhaegens, G. Knuyt, and L. M. Stals, Energy distribution of ions in an unbalanced magnetron plasma measured with
energy-resolved mass spectrometry, Surf. Coat. Technol., vol. 89, no. 1/2,
pp. 177184, Feb. 1997.
[14] L. de Poucques, J. Imbert, C. Boisse-Laporte, J. Bretagne, M. Ganciu,
L. Teule-Gay, and M. Touzeau, Study of the transport of titanium neutrals
and ions in the postdischarge of a high power pulsed magnetron sputtering device, Plasma Sources Sci. Technol., vol. 15, no. 4, pp. 661669,
Nov. 2006.
[15] J. T. Gudmundsson, J. Alami, and U. Helmersson, Spatial and temporal
behavior of the plasma parameters in a pulsed magnetron discharge, Surf.
Coat. Technol., vol. 161, no. 2/3, pp. 249256, Dec. 2002.
[16] J. Alami, J. T. Gudmundsson, J. Bohlmark, J. Birch, and
U. Helmersson, Plasma dynamics in a highly ionized pulsed magnetron
discharge, Plasma Sources Sci. Technol., vol. 14, no. 3, pp. 525531,
Aug. 2005.
[17] K. Macak, V. Kouznetsov, J. Schneider, U. Helmersson, and I. Petrov,
Ionized sputter deposition using an extremely high plasma density pulsed
magnetron discharge, J. Vac. Sci. Technol. A, Vac. Surf. Films, vol. 18,
no. 4, pp. 15331537, Jul. 2000.
[18] M. B. Hopkins and W. G. Graham, Time-resolved electron energy distribution function measurements in a pulsed magnetic multipole hydrogen
discharge, J. Appl. Phys., vol. 69, no. 6, pp. 34613466, Mar. 1991.
[19] J. T. Gudmundsson, J. Alami, and U. Helmersson, Evolution of the
electron energy distribution and plasma parameters in a pulsed magnetron
discharge, Appl. Phys. Lett., vol. 78, no. 22, pp. 34273429, May 2001.
[20] A. Vetushka and A. P. Ehiasarian, Plasma dynamic in chromium and
titanium HIPIMS discharges, J. Phys. D, Appl. Phys., vol. 41, no. 1,
p. 015204, Jan. 2008.
[21] P. Vasina, M. Mesko, M. Ganciu, J. Bretagne, C. Boisse-Laporte,
L. D. Poucques, and M. Touzeau, Reduction of transient regime in
fast preionized high-power pulsed-magnetron discharge, Europhys. Lett.,
vol. 72, no. 3, pp. 390395, Nov. 2005.

IEEE TRANSACTIONS ON PLASMA SCIENCE, VOL. 39, NO. 4, APRIL 2011

[22] T. Nakano and S. Baba, Simulation of particle transport in high pressure


sputtering, Vacuum, vol. 51, no. 4, pp. 485489, Dec. 1998.
[23] S. Kadlec, Simulation of neutral particle flow during high power
magnetron impulse, Plasma Process. Polym., vol. 4, pp. S419S423,
Apr. 2007.
[24] A. P. Ehiasarian and R. Bugyi, Industrial size high power impulse magnetron sputtering, in Proc. 47th Annu. Tech. Conf. SVC, Denver, CO,
2005, pp. 486490.
[25] M. W. Thompson, The energy spectrum of ejected atoms during the high
energy sputtering of gold, Philos. Mag., vol. 18, no. 152, pp. 377414,
Aug. 1968.
[26] A. Hecimovic and A. P. Ehiasarian, Time evolution of ion energies in
HIPIMS of chromium plasma discharge, J. Phys. D, Appl. Phys., vol. 42,
no. 13, p. 135209, Jul. 2009.
[27] M. A. Liberman and A. J. Lichtenberg, Principles of Plasma Discharges
and Materials Processing. New York: Wiley, 1994.
[28] J. Bretagne, W. G. Graham, and M. B. Hopkins, A comparison of
experimental and theoretical electron energy distribution functions in a
multicusp ion source, J. Phys. D, Appl. Phys., vol. 24, no. 5, pp. 668
671, May 1991.

Ante Hecimovic received the B.Sc. degree from


the University of Zagreb, Zagreb, Croatia, in 2006,
and the Ph.D. degree in the field of a novel Physical Vapour Deposition technology HIPIMS (highpower impulse magnetron sputtering) from Sheffield
Hallam University, Sheffield, U.K., in 2009.
He is a Postdoc in the Institute for the Experimental Physics 2, Ruhr University, Bochum, Germany.
His research interests are in plasma diagnostic of the
pulsed magnetron discharges, with a special focus
on fundamental plasma processes. The center of
attention is on the understanding of the HIPIMS plasma discharge ignition,
discharge development and particle transport.

Arutiun P. Ehiasarian received the Undergraduate


degree from the University of Sofia St Kliment
Ohridski, Sofia, Bulgaria, in 1996, the B.Sc. degree
(Hons.) in applied physics from the University of
Salford, Salford, U.K., in 1998, and the Ph.D. degree
in the field of plasma science and surface engineering from the Nanotechnology Centre for PVD Research, Sheffield Hallam University, Sheffield, U.K.,
in 2002.
He is heading the HIPIMS Technology Centre at
the Materials and Engineering Research Institute,
Sheffield Hallam University, and the Director of the Joint Sheffield Hallam
UniversityFraunhofer IST HIPIMS Research Centre based in Sheffield, U.K.
He is the Chairman of the International Conference on HIPIMS and Vice Chair
of the European COST Action on Highly Ionised Pulsed Plasmas. His research
interests are in development of plasma technology to improve properties and
performance of coatings for a variety of applications. His experience is with
cathodic vacuum arc discharges, dc and pulsed magnetron discharges, and
radio-frequency coil-enhanced magnetron sputtering. Application areas are
high temperature-, corrosion-, and wear-resistant coatings for automotive and
aerospace engines, high aspect ratio (through-silicon) via metallization for
3-D integrated circuits, cryogenic materials for space satellites, and thin-film
photovoltaics for solar cells.

Das könnte Ihnen auch gefallen