You are on page 1of 54

Plasma Spray Torches

Maher I. Boulos, Pierre Fauchais, and Emil Pfender

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Basic Concepts of Plasma Spraying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1 Plasma Spraying and the Thermal Spray Industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Plasma Spray Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Principle of Plasma Spraying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3 DC Plasma Spray Torches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1 Conventional Plasma Spray Torches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Effect of Nozzle Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3 Supersonic Atmospheric Plasma Spraying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4 Mini and Micro Plasma Spray Torches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.5 The Triplex Torch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.6 Delta Plasma Spray Torch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.7 DC Plasma Torches with Axial Particle Injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4 High-Power DC Plasma Spray Torches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.1 Cascade Plasma Torches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.2 PlazJet Torch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.3 Water-Stabilized Plasma Spray Torch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4 CACT Plasma Torch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5 Plasma Spray System Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.1 Atmospheric Plasma Spraying (APS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Emil Pfender: Deceased.


M. I. Boulos (*)
Department of Chemical Engineering, University of Sherbrooke and Tekna Plasma Systems Inc.,
Sherbrooke, QC, Canada
e-mail: maher.boulos@tekna.com
P. Fauchais
European Center of Ceramics, University of Limoges, Limoges, France
e-mail: pfauchais@gmail.com
E. Pfender
Department of Mechanical Engineering, University of Minnesota, Minneapolis, MN, USA

# Springer International Publishing AG 2018 1


M. Boulos et al. (eds.), Handbook of Thermal Plasmas,
https://doi.org/10.1007/978-3-319-12183-3_49-1
2 M. I. Boulos et al.

5.2 Controlled Atmosphere Plasma Spraying (CAPS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39


5.3 Vacuum Plasma Spraying (VPS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.4 Plasma Spraying-Physical Vapor Deposition (PS-PVD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

Abbreviations
APS Atmospheric plasma spraying
CAPS Controlled atmosphere plasma spraying
CARS Coherent anti-Stokes Raman spectroscopy
CBL Cold boundary layer
CGDS Cold gas dynamic spraying
CNTs Carbon nanotubes
CS Cold Spray
DC Direct current
D-gun Detonation gun
EBCs Environmental barrier coatings
EB-PVD Electron beam-physical vapor deposition
ERW Electric Resistance Welding
FS Flame spraying
HVOF High-velocity oxy-fuel flame
i.d. Internal diameter
ISPC International Symposium on Plasma Chemistry
ITSC International Thermal Spray Conference
LIF Laser-induced fluorescence
LPG Liquefied petroleum gas
LPPS Low-pressure plasma spraying
NTSC National Thermal Spray Conference
PS-PVD Plasma spraying-physical vapor deposition
PTA Plasma Transferred Arc
RF Radio frequency
SHS Self-propagating high-temperature synthesis
SOFCs Solid oxide fuel cells
SWAS Single-wire arc spraying
TBCs Thermal barrier coatings
TGO Thermally grown oxide
UTSC United Thermal Spray Conference
VPS Vacuum plasma spraying
WPS Water plasma spraying
YSZ Yttria-stabilized zirconia
Plasma Spray Torches 3

1 Introduction

Plasma spraying belongs to a wide range of thermal spray processes in which finely
divided metallic or nonmetallic materials are deposited on the substrate in a molten
or semi-molten condition to form a coating. The coating material may be in the form
of powder, ceramic rod, or wire which is introduced into the plasma jet, melted, and
projected toward the surface of the substrate forming a coating of thickness that can
vary between a few microns to several millimeters. The rapid development and
acceptance of the technology as a reliable material processing technique is based on
a solid understanding of the fundamental processes involved. According to
(Fauchais et al. 2014), the motivation for coating structural parts can be summarized
by the following needs:

• Decouple surface properties from the bulk properties of the materials by modi-
fying their surface properties such as tribological, wear, and corrosion resistance
• Improve functional performance of the materials by, e.g., allowing higher tem-
perature exposure using thermal barrier coatings
• Improve component life by reducing wear due to abrasion, erosion, and corrosion
• Recycling of worn parts through rebuilding them to their original dimensions
avoiding the need for replacing the entire component
• Improved economics using low-cost bulk materials combined with high func-
tionality coatings

The wide spread use of the technology has created an ever-increasing demand on
the functionality and performance of the of plasma spray coating system compo-
nents. The plasma torch design which is at the heart of such systems has evolved
considerably over the past two decades to meet such requirements as:

• Compact lightweight torch design for ease of handling by a robotic arm for the
coating of intricate complex-shaped parts including internal surfaces of engine
cylinder bores
• Operation under atmospheric pressure and vacuum conditions for improved
coating quality
• Extended electrode life for improved coating reliability and process economics
• High-power operation for increased material throughput and coating rates and
improved process economics

In this chapter, a review is presented of different designs and performance of DC


plasma spraying torches commonly used in the plasma spray industry. The content
builds on discussions of the fundamental phenomena involved which can be found in
earlier chapter “▶ Basic Concepts of Plasma Generation”; chapter “▶ Thermal
Arcs”; chapter “▶ Electrode Phenomena in Plasma Sources”; and chapter “▶ DC
Plasma Torch Design and Performance.” Alternate designs of RF inductively
coupled plasma torches used in powder processing and for certain thermal spray
applications are discussed in chapter “▶ Inductively Coupled Radio Frequency
Plasma Torches.”
4 M. I. Boulos et al.

2 Basic Concepts of Plasma Spraying

2.1 Plasma Spraying and the Thermal Spray Industry

An overall display of the different subgroups in the family of “thermal spray


technologies” is given in Fig 1. This comprises essentially two broad groups,
combustion-based processes and electric-arc (plasma)-based processes. Plasma
includes atmospheric spraying, vacuum DC plasma spraying, RF induction plasma
spraying, Wire Arc Spraying (WAS), and Plasma Transferred Arc (PTA) deposition.
The general concept of thermal spray is illustrated in Fig. 2. The central part of the
plasma spray system is the plasma torch, which converts the supplied electrical
energy into a stream of high-temperature gases. The coating material, introduced into

Fig. 1 Classification of thermal spray technologies (Fauchais et al. 2014)

Fig. 2 Schematic of the thermal spray concept except Cold Spray and weld coating (Fauchais et al.
2014)
Plasma Spray Torches 5

the plasma jet in powder form, is heated, melted, and accelerated toward the
substrate, where the molten particle droplet deforms on impact to generate a splat.
Multiple layers of splats form the coating. In the special case of wire arc spraying,
the material to be sprayed is introduced into the spray torch in the form of a
continuous wire which is heated, melted, and atomized by a high-velocity gas jet.
The coating formation step is the same as that of standard plasma spraying.
Figure 3 presents on a temperature-velocity diagram the range of values that can
be expected for gases heated by the different thermal spray processes. This shows
that the plasma-based processes offer high-temperature capabilities (mostly above
7000 K) with DC plasma sources offering considerably higher plasma jet velocities
(above 900 m/s) compared to RF induction plasma sources or PTA and Wire arc.
While the invention of thermal spray process is credited to M.U. Schoop in 1910,
the technology was essentially limited to flame spraying up to the 1950s. In the
mid-1950s, the first plasma spray torch was developed by Thermal Dynamics Corp.
followed by Metco and Plasmadyne Corps., which form the basis of many of the
torch designs used today. Figure 4 (courtesy of Öerlikon Metco) illustrates the
history of thermal spray torch designs and the improvement of their performance
which reflects better understanding of the complexity of the fundamental processes
involved and technology needs.
Throughout this period, thermal spray has been a truly enabling technology since
the significant progress in the aircraft engines and high-efficiency gas turbine
engines would not have been possible without thermal spray coatings. Thermal
spray technology has also made economically possible the substitution of steel
with lighter weight materials such as Al or Mg and their alloys. The use of aluminum
engine blocks in automobile industry to reduce its weight is another example for

Fig. 3 Gas temperatures and velocities obtained with different thermal spray systems (Fauchais
et al. 2014)
6 M. I. Boulos et al.

Fig. 4 Milestones in the development of the thermal spray industry. (Reproduced with kind
permission of Öerlikon Metco, Switzerland)

how thermal spray technology has respond to industrial and social needs. The
development of miniaturized plasma spray torches has been critical for the coating
of cylinder walls in aluminum engine blocks with materials withstanding the high
temperatures and corrosive environment of the internal combustion chamber
(Barbezat 2001). The need for replacement of wear-resistant hard chrome coatings
has also been a strong driver for expanding thermal spray technology.
An overview of industrial applications of plasma spray technologies in Europe in
2002 is presented in Fig. 5 (Ducos and Durand 2001). This shows that the early
domination of aerospace applications, which was around 50% of industrial thermal
spray applications in the 1980s, has dropped to 28% due to new applications in the
automotive and chemical process industries. A gradual shift, from high added-value
products to high-volume production items, is also an indication of the increased
maturity and reliability of the technology. While the diversity of applications is
likely to increase, the main pillars of users are still expected to remain the aerospace,
automotive, power, and chemical process industries.
An excellent and broad review of the range of applications of thermal spray
technology has been reported in the volume edited by Tucker (2013). These include:

• Thermal and environmental barrier coatings (TBCs/EBCs) for turbine engines


• Gas turbine applications
• Abradable thermal spray applications and technology
• Replacement of hard chrome plating on aircraft landing gear
• Automotive coatings and applications
• Biomedical coatings for orthopedic joints
• Corrosion and wear control for industrial applications
Plasma Spray Torches 7

Fig. 5 Industrial applications of thermal spray technology in Europe in 2001 (Ducos and Durand
2001)

• Corrosion control for marine- and land-based infrastructure applications


• Wear-resistant coatings
• Protective overlays and coatings used in oil sands
• Renewable energy applications
• Nuclear industry applications
• Electrical and electronic application
• Applications in the steel industry
• Applications in the paper industry
• Printing industry applications
• Other applications in metal processing, textile and plastics industry, forming
process, and ceramic and glass industry that are also in various stages of
development

In the following a summary is presented of the different plasma spray techniques


used. It is important to point out that the quality of the coating obtained by each of
these techniques depends to a large extent on the nature of the powder or precursor
used, powder morphology, particle size distribution (PSD), and the characteristics of
the spraying device. The different processes are generally not competitive but
complementary. The choice between them depends mainly on the coating properties
that are desired, the material to be sprayed, as well as economic considerations,
which are often the determining factor. For example, to spray a ceramic material,
plasma spraying is the most widely adopted process. However, if a high-porosity
coating is acceptable, flame spraying of ceramic rods will be a more economical
option.
8 M. I. Boulos et al.

2.2 Plasma Spray Sources

2.2.1 Direct Current (DC) Plasma Torches


DC plasma torches are used to generate a plasma jet through the heating of a
continuously flowing gas by an electric arc inside a nozzle (described in detail in
this chapter). When operated under atmospheric pressure, the process is commonly
known as atmospheric plasma spraying (APS), while under soft vacuum conditions
(10 < p < 70 kPa), it is commonly referred to as vacuum plasma spraying (VPS). In
a few cases, supersonic plasma jets have been generated in APS when using torches
with special design of the anode nozzle (Fauchais 2004). Most of the applications
use powders, but more recently solution and suspension precursors were used.
Controlled atmosphere chambers filled with an oxygen-free gas at pressures between
80 and 300 kPa are often used for spraying metals, alloys, and non-oxide ceramics
with an excellent adhesion to the substrate.

2.2.2 Radio-Frequency (RF) Inductively Coupled Plasma Torches


The plasma is generated through the electromagnetic coupling of the energy into the
discharge cavity. The energy density is significantly lower than in DC spray torches
resulting in plasma temperatures (6000–9000 K) and lower plasma velocities (less
than 100 m/s). Through their central powder injection capabilities and long particle
residence time, RF induction plasma sources are best suited for the melting and
deposition of relatively large particle size powders (50–200 μm) with deposition
efficiencies above 85–90%. They are also commonly used for the spraying of high
purity materials, solution and suspension plasma spraying, as well as for powder
densification and spheroidization (for more details see chapter “▶ Inductively
Coupled Radio Frequency Plasma Torches”).

2.2.3 Wire Arc Spray Torches


In these types of plasma spray torches, the arc is struck either between two contin-
uously advancing metal or alloy wires, one of which acting as cathode and the other
as anode, or between an advancing metallic wire and a fixed refractory metal
electrode. The melted tips of the wire(s) are fragmented into tiny droplets (a few
tens of μm in diameter) by the atomizing gas blown between both wires. The wires
are necessarily made of ductile material or of a ductile envelope filled with a
non-ductile material such as a ceramic powder (commonly referred to as “cored
wires”). For more details see chapter “▶ Wire Arc Spraying Torches.”

2.2.4 Plasma Transferred Arc (PTA)


In this type of plasma torches, the plasma is generated using the same principle as
that of the blown arc in a DC plasma torch with the exception that the arc is
transferred between the cathode of the torch, surrounded by a floating nozzle, and
the substrate, necessarily metallic, which, in most cases, becomes the anode. The
transferred arc induces local melting of the substrate, while the particles injected into
the plasma column are heated and captured into the molten metal pool where their
melting is completed by the arc column. In certain cases, the particles injected into
Plasma Spray Torches 9

the arc do not melt and are simply captured into the molten metal bath forming, on
solidification, a metal matrix composite. This process is described in chapter
“▶ Plasma Cutting, Welding, and PTA Torches.”

2.3 Principle of Plasma Spraying

2.3.1 Principal System Components


The principal components of a typical plasma spray setup are shown in Fig. 6. The
central piece of equipment is the plasma torch. A gas and power control console
allows the adjustment of the operating parameters, i.e., the control of arc ignition, arc
current, plasma gas flow rates, and powder and carrier gas flow rates. The control
console also houses safety interlocks to avoid starting the arc under unsafe condi-
tions such as lack of cooling water or primary plasma gas flows. The auxiliary
system components are the power supply system including the high-frequency
starter unit, the plasma gas supply system, the high-pressure cooling water system,
the powder feed system, and the exhaust gas evacuation.
Typical cooling water flow rates required vary between 15 and 35 l/min (4–10
USGPM) with pressures around 1.5 MPa (220 psig). Typical operating conditions
for plasma spray torches are given in Table 1.

Fig. 6 Schematic of a plasma spray system with plasma torch, power supply, high-frequency
starter, cooling water supply, gas supply, spray powder supply, and control unit (Fauchais et al.
2014)
10 M. I. Boulos et al.

Table 1 Typical operating conditions for conventional DC plasma spray torches


Plasma gas composition Gas flow rate (slm) Torch current (A)
Ar-He (up to 60 vol.% He) 30–120 300–1200
Ar-H2 (up to 25 vol.%H2) 30–90 300–700
N2-H2 (up to 25 vol.% H2) 30–90 300–500
Ar-H2-He 30–90 300–700

Fig. 7 Schematic of the different elements in the DC plasma spray process with the coating powder
injected radially into the plasma jet and structure representation of typical coating (Fauchais et al.
2014)

In plasma spraying, an electric spark is used to ignite a plasma within the torch.
The plasma is sustained by an arc struck between a cathode (usually a rod- or button-
type design) and a cylindrical anode nozzle. The plasma gas injected at the base of
the cathode is heated by the arc and exits the torch through a water-cooled nozzle as a
high-temperature, high-velocity jet. A schematic of a typical setup used in DC
plasma spraying in an open atmosphere is presented in Fig. 7. Peak plasma temper-
atures and velocities at the nozzle exit can be, respectively, as high as 12,000 to
15,000 K and 500 to 2500 m/s.
Pfender et al. (1991) showed that as the plasma jet exits the nozzle, it encounters a
steep laminar shear at the outer edge of the jet. The large velocity difference causes
rolling up of the shear layer flow around the nozzle exit into a ring vortex, which is
pulled downstream by the flow, allowing the process to repeat itself again at the
nozzle exit. Adjacently formed vortex rings at the outer edge of the jet have the
tendency to coalesce, forming larger vortices with perturbations to these vortices
leading to wave instabilities growing around the entire vortex ring. The distorted
vortex rings start entangling themselves with the adjacent rings, finally resulting in
total breakdown of the vortex structure into large-scale eddies and the onset of
turbulent flow as sketched in Fig. 7. This entanglement process of adjacent unstable
Plasma Spray Torches 11

vortices results in large-scale engulfment of ambient air, or environment gas when


spraying in a chamber, although some entrainment also takes place during the roll-up
process of the jet’s shear layer.
Maximum arc currents depend on the i.d. of the anode nozzle, the composition of
the plasma-forming gas and its flow rate. The time-averaged arc voltage, its standard
deviation, and the temperature rise of the cooling water mainly characterize the
operation of a plasma torch. These parameters allow determining the torch energy
efficiency, the energy transferred into the plasma jet, and the average specific
enthalpy of the plasma gas, which is typically in the range from 10 to 40 MJ/kg.
The corresponding average plasma temperature and velocity and the axial and radial
temperature and velocity profiles of the plasma jet are strongly dependent on the
composition of the plasma-forming gas and its mass flow rate, the way the gas is
injected into the torch, the cathode and anode nozzle design, the arc current, and the
ambient gas composition. In the following a separate discussion is presented on the
influence of each of these parameters on the plasma torch performance and the
characteristics of the plasma jet.
All power sources deliver an adjustable constant current intensity. The
corresponding voltage depends on the composition of the plasma-forming gas,
cathode-to-anode distance and anode nozzle internal diameter. The arc root fluctu-
ates at the anode with two characteristic modes: take-over obtained with mono-
atomic gases (Ar or Ar-He) and restrike one with plasma-forming gases containing
diatomic gases (Table 1).
The restrike mode is characterized by large amplitude voltage fluctuations of a
saw-tooth shape reflecting corresponding fluctuations of the length of the high-
temperature core of the plasma jet generally defined by the 8000 K temperature
isotherm. Typically, the length of the plasma core can vary between 50 mm to 65 mm
causing significant variations of arc voltage and torch power which will have in turn
a critical impact on plasma jet stability and coating quality.
Solid particles to be melted are most commonly injected into the plasma stream in
the radial direction either outside of the nozzle as shown in Fig. 7 or inside the nozzle
but close to its exit. To achieve near optimum heat and momentum transfer in spray
conditions, an angle of 3.5–4 between the particle mean trajectory and the torch axis
seems to give the best results (Vardelle et al. 2001). Particle penetration into the jet
depends on the injector internal diameter and shape (straight or curved), its position,
and inclination relative to the plasma jet for external injection, while internal
injection can induce a high deviation of the plasma jet as soon as the carrier gas
exceeds 10% of the plasma mass flow rate (Vardelle et al. 2001).
To illustrate the importance of the carrier gas on particle trajectories, a photodiode
®
sensor, Spray and Deposit Control (SDC ) was used to observe the plasma plume
emission of a standard DC plasma torch operated using an Ar/H2 mixture (45 slm
Ar + 15 slm H2) as plasma gas and a torch power of 20 kW. A filter with 3 nm
bandwidth allows eliminating the most important background emission of the
plasma plume. Alumina particles were injected into the plume using a straight
cylindrical injector 1.75 mm i.d. placed at 3 mm downstream of exit plane of the
torch nozzle with the injector tip 8 mm from the jet axis. Schematic representation of
12 M. I. Boulos et al.

Fig. 8 SDC measurement of alumina particles heat flux 60 mm downstream of the nozzle exit of an
Ar-H2 plasma with different argon carrier gas flow rates: (a) mean trajectories, (b) radial distribution
of emitted light intensity (Vardelle et al. 2001)

the different mean particle trajectories injected with different carrier gas flow rates is
given in Fig. 8a. The corresponding radial profiles of the light intensities emitted by
the particles in-flight at a location 60 mm downstream of the nozzle exit for different
carrier gases flow rates are given in Fig. 8b. The position of the maxima observed for
the different carrier gas flow rates is shifted to the right or the left of the optimal one
(corresponding for these spray conditions to an argon carrier gas flow rate of
4.5 slm). It is important to point out that any change in the spray parameters requires
adjusting the carrier gas flow rate. Moreover, the lower the dispersion of the particle
sizes, the narrower will be their trajectory distribution.

2.3.2 Splat Formation


First each melted particle upon impact flattens and solidifies as a splat. Splat shapes
depend on many factors, including the particle velocity and temperatures at impact,
the degree of the superheating of the particle, the in-flight oxidation, the partial
solidification of the particle in-flight resulting in the formation of a solid shell around
the liquid central part (this occurs especially for low-thermal conductivity materials),
the surface roughness of the substrate and presence of contaminants, and the impact
angle of the droplet relative to the substrate surface. According to recent work
(Fauchais et al. 2014) where the flattening of droplets has been followed on flat
substrates, the main types of splats observed are presented in Fig. 9.
Since the beginning of the 1980s, much work was devoted to the study of the
impact and deformation of single molten particles on their impact on a substrate to
form a solid splat (Fauchais 2004; Chandra and Fauchais 2009). These studies
emphasized the difficulty of monitoring the impact of particles in a molten state
Plasma Spray Torches 13

Fig. 9 Main types of molten splats: (a) and (b) extensively fragmented, alumina on cold 304 stain-
less steel (SS); (c) and (d) close to disk shaped, c. Ni on 304 SS pre-oxidized at 150  C, d. alumina
on low carbon steel preheated at 200  C; (e) and (f) fingered splats, (e) alumina on SS preheated at
400  C, (f) Ni on pre-oxidized SS at 650  C (Fauchais et al. 2014)

with sizes between 20 and 60 μm and velocities ranging between 50 and 600 m/s.
The flattening time of such molten particles is only a few microseconds, and
solidification starts before flattening is completed. State-of-the-art optical techniques
used to monitor splat formations have shown that the substrate temperature had a
significant influence on the shape of the formed splats (Goutier et al. 2011, 2013;
Fauchais et al. 2016). Fragmented splats were observed to be formed at relatively
low temperatures, while disk-shaped splats (see Fig. 9d) were obtained at higher
temperature. The transition temperature, Tt, between these two regimes was found to
depend on the thermophysical properties of the deposited material and that of the
substrate. To promote the formation of disc-type splats and to get rid of adsorbates
and condensates, which on evaporation lifts the flattening particle resulting in
extensively fragmented splats, it was recommended to preheat the substrate to a
temperature just above the transition temperature Tt. Such preheating must be
carefully controlled to limit as much as possible the development of the oxide
layer at the surface of metal or alloy substrates. Consequently, it can be stated that
coating adhesion depends, among other factors, on substrate preheating over Tt, but
not too much over, and its roughening adapted to the mean size of impacting
particles. The adhesion of the same coating deposited on the same cold or hot
(over Tt) substrate is up to ten times higher.
14 M. I. Boulos et al.

2.3.3 Beads and Coating Formation


Coating formation takes place through the piling up of successive layers of particle
splats on the substrate (Fauchais et al. 2014). Schematic representations of the
process are given in Figs. 7 and 10. The real contact between successive splat layers
determines the coating properties such as thermal conductivity and Young’s modu-
lus. Interlamellar pores, often called globular, are observed between layered splats,
or first splats, and the substrate. For clarity, the size of these pores is
disproportionally enlarged in Fig 10a.The interlamellar flat pores are thin voids,
orthogonal to the spray direction, filling spaces between splats. These interlamellar
pores are part of coating porosity, and their thickness is a few hundredths to a few
tenths of micrometers. The real contact between splats increases with particle impact
velocities, if they are not either excessively superheated or below their melting
temperature. Splashing of the melted particles during flattening upon impact can
significantly affect the coating properties (Fauchais et al. 2014).
The lower adhesion of splats deposited on splashed material, as described sche-
matically in Fig. 9b, can become more important when spraying metals because
splashed material is oxidized rather fast due to the small droplet sizes. Pores are
formed during coating generation, as illustrated in Fig. 11. They are due to the
shadowing effect Fig. 11a, narrow holes in valleys between splats, which are not
filled Fig. 11b, due to unmelted or partially melted particles that create the worst
defects within the coatings Fig. 11c, or due to exploded particles Fig. 11d.
As schematically illustrated in Fig. 12, the formation of a coating over a large
surface is typically achieved through the linear motion of the torch at a fixed
spraying distance from the substrate. The formation of a bead, illustrated in
Fig. 12a, is the result of single path of the torch on the substrate with the width of
the bead, wb, dependent on the divergence angle of the spraying jet and the spraying
distance, while the height of the bead, hb, depends on the powder feed rate m p into the
plasma jet, the deposition efficiency, ηd, defined as the fraction of the powder
injected into the plasma jet that is deposited on the substrate, and the relative velocity
vr of the torch with respect to the substrate. The weak part of the bead is its wings,

Fig. 10 Schematic
representation of coating
formation during plasma
spraying showing the globular
pores within the coatings due
to (a) shadow effect, (b)
exploded particles (Fauchais
et al. 2014)
Plasma Spray Torches 15

Fig. 11 Schematic
representation of globular
pores within the coatings: (a)
shadow effect, (b) narrow
holes in valleys between
splats, (c) unmelted or
partially melted particles, (d)
exploded particles

Fig. 12 (a) Schematic representation of a bead, (b) typical spray pattern on a flat substrate
(Fauchais et al. 2014)

because they result from particles traveling in the plasma jet fringes and contain the
highest quantity of poorly treated particles. To reduce them the injection conditions
must be optimized and combined with the use of a powder with a narrow particle size
distribution. Once the conditions for bead deposition are optimized, the complete
coating of the surface is achieved through the overlapping of successive beads as
illustrated in Fig. 12b. The distance dy between successive passes has to be carefully
evaluated as a function of the bead profile in order to insure a reasonable uniformity
of the surface of the coating.
While the coating of flat plates or cylindrical parts will require relatively simple
control and programming of the torch movement, more sophisticated computerized
robots are needed as soon the shape of the part is complex. Complex geometries will
generally require optimized torch operating conditions coupled with precise control
of the torch movement. It is also important to maintain during spraying process a
16 M. I. Boulos et al.

Fig. 13 Micrographs of (a) non-optimized and (b) optimized yttria-stabilized zirconia (YSZ)
coatings (Fauchais et al. 2014)

close monitoring and control of the surface temperature of the substrate and coating
using appropriate pyrometers or infrared camera coupled with external cooling
systems using gas jets strategically positioned and directed toward the substrate
and the coated surfaces. These are necessary to avoid the generation of thermal
stresses in the coating caused by the heat released from the impacting molten
droplets and the hot plasma plume.
Finally, in an industrial production environment operating 24/7 it is important to
take the necessary precautions to follow closely electrode wear which can be
responsible for the gradual decrease of the arc voltage while other spray parameters
such as plasma gas flow rates and arc current remain constant. This could result in
the gradual drifting of the operating conditions and the deterioration of the quality of
the coating unless appropriate corrective action is taken. Leblanc and Moreau (2002)
demonstrated that in-flight measurements of particle parameters such as mean
velocity and temperature can be used as an effective technique to monitor and
control the spraying conditions and insure a uniform and constant coating quality.
To illustrate the link between coating conditions and coating quality, typical
micrographs of yttria-stabilized zirconia (YSZ) coatings obtained using the same
plasma spray installations, spraying YSZ powder with a particle size range of
5–22 μm, under two sets of conditions are given in Fig. 13a and b. The coating on
the LHS (Fig. 13a) was obtained under non-optimized conditions, while the one on
the RHS (Fig. 13b) was obtained under optimal conditions. For non-optimized spray
conditions, the porosity of the coating was measured as 10.9%, and its ionic
conductivity was 0.019 S/m. For the coating obtained under optimized conditions,
the porosity was down to 6.1% and the ionic conductivity up to 0.5 S/m which is
considerably closer to 1.0 S/m, the value for sintered YSZ, reflecting a superior
contact between successive layers of splats in the coating.

2.3.4 Finishing and Posttreatment of Coatings


After spraying, coatings are generally posttreated to achieve the final dimension and
obtain a smooth surface. Machining of coatings can be troublesome with materials
that are abrasion resistant, porous, and with poorly bonded particles or splats that can
Plasma Spray Torches 17

pull out easily. Consequently, generally accepted techniques used to machine


wrought material do not necessarily apply to the same material when sprayed
(Davis 2004; Tucker 2013). Posttreatments of coatings can also be performed to
enhance the bond strength; relax partially or totally residual stress; create compres-
sive residual stress; close or reduce porosity, especially the interconnected one;
create a barrier to corrosive or oxidizing products; improve coating homogeneity;
improve splat or particle cohesion; obtain hard phase precipitation; or induce
chemical modifications. For details see (Davis 2004; Fauchais et al. 2014; Tucker
2013).

3 DC Plasma Spray Torches

3.1 Conventional Plasma Spray Torches

The first plasma spray torches designed in the 1960s involved all stick-type cathodes
and cylindrical anodes as schematically represented in Fig. 7. They differed in the
details of the design of the plasma-forming gas injection ring, arc chamber config-
uration, and the details of the design of the cathode and anode and their
corresponding cooling-water circuits. The schematic drawing presented in Fig. 14
represents a typical commercial DC spray torch design (SG100 by Praxair-TAFA)
showing its central stick-type hot cathode and the annular, water-cooled cold anode.

Fig. 14 Schematic representation of the Praxair-TAFA, model SG100, DC plasma spraying torch
with internal powder injection into the anode nozzle. (Reproduced with kind permission of Praxair-
TAFA, USA)
18 M. I. Boulos et al.

In most of these designs, the cooling-water circuits of the cathode and anode are
interconnected in series. The high-pressure cooling-water, typically at a pressure of 1
to 1.18 MPa (145–180 psig), usually enters the torch on the side of the component
with the lowest pressure drop (anode) and exits from the side of the higher pressure
drop component (cathode). This insures the highest back pressure in the cooling
water circuit in the torch to avoid water boiling on the surfaces of the electrodes. The
revers flow configuration could also be used depending on the pressure drop in each
of these two critical components. In many of these designs, the cooling-water hoses
also carry the power cables for each of the two electrodes. As shown in the figures,
the plasma gas is injected separately into the discharge cavity through a gas
distribution ring, which serves to control the flow pattern in the discharge. The
powder to be sprayed is injected into the plasma stream using a carrier gas, either
internally, into the anode nozzle as shown in Fig 14, or externally at the exit level of
the plasma jet, as shown in Figs 15 and 16. The latter two show the design of one of
the most widely used commercial plasma spray torches manufactured by Öerlikon
Metco, models PTF4 and F4MB-XL.

Fig. 15 Sectional view of an


Öerlikon Metco model
PTF-4DC plasma spray torch
with external powder injection
at the exit level of the anode
nozzle. (Reproduced with
kind permission of Öerlikon
Metco AG, Switzerland)

Fig. 16 Photograph of an
Öerlikon Metco model
F4MB-XLDC plasma spray
torch with double port
external powder injection at
the exit level of the anode
nozzle. (Reproduced with
kind permission of Öerlikon
Metco AG, Switzerland)
Plasma Spray Torches 19

This type of torches represents more than 70% of the currently used spray torches.
They are typically designed for power levels between 40 and 80 kW. In most cases
their weight is below 3 kg without the connecting cables, which allows mounting
them on a robot arm. Their external diameter is typically between 50 and 90 mm and
their lengths between 100 and 180 mm. The anode nozzle i.d. is typically between
5 and 8 mm and its length between 1 and 5 times the nozzle i.d. The gas injection can
be axial, conical, radial, or as vortex. Most plasma torches operate in a current
control mode, i.e., the current is kept constant. The current set point and the plasma
gas flow rate and composition are among the most important independent parameters
and are chosen by the operator.
Maximum arc currents depend on the anode nozzle i.d., the composition of the
plasma-forming gas, and its flow rate. The time-averaged arc voltage, its standard
deviation, and the temperature rise of the cooling water mainly characterize the
operation of a plasma torch. These parameters allow determining the torch energy
efficiency, the energy coupled into the plasma jet, and the average specific enthalpy
of the plasma gas, which is typically in the range from 10 to 40 MJ/kg. The
corresponding average plasma temperature and velocity, as well as the axial and
radial temperature and velocity profiles of the plasma jet, are strongly dependent on
the composition of the plasma-forming gas and its mass flow rate, the way the gas is
injected into the torch, the cathode and anode nozzle design, the arc current, and the
ambient gas composition.
Plasma spray torches work mainly with Ar as primary gas or sometimes diatomic
nitrogen, alone or in combination with other inert or molecular gases, such as for Ar,
He, or H2. Ternary mixtures such as Ar-He-H2 can also be used. The mass of primary
gas is important to accelerate particles which allows them to acquire a high momen-
tum before their impact on the substrate. The secondary gases are important for the
heat transfer to particles. Helium increases the viscosity of the mixture at high
temperatures (over 10,000 K) as well as its thermal conductivity, while hydrogen,
even at concentrations as low as 10–20 vol.% H2, increases significantly the thermal
conductivity of the gas mixture at temperatures above 4000 K. Increasing the
concentration of H2 beyond 30 vol.% H2 is more challenging, requiring a significant
increase of the total plasma power and gas flow rate to hundreds slm. The addition of
even small concentrations of hydrogen, or helium, to the argon plasma gas, for a
given arc current (anode nozzle i.d. 6 mm and I = 500 A) gives rise to a sharp
increase of the arc voltage and the corresponding torch efficiency (Roumilhac et al.
1990a). This enhanced cooling of the arc column in the presence of hydrogen is due
to the fast diffusion of the smaller hydrogen atoms into the fringes of the arc column,
thus increasing significantly the thermal conductivity. It should be mentioned that
larger hydrogen concentration in the plasma gas is not recommended since it will
also significantly increase electrode erosion. The reason for the voltage increase is
the much higher specific heat of hydrogen and its higher thermal conductivity, both
requiring higher electric fields to sustain an arc. The higher efficiency can be
explained by the fact that the anode heat losses are primarily determined by the
current flow, and, for constant current, increased power dissipation due to higher
20 M. I. Boulos et al.

voltages will reduce the fraction of the power transferred to the anode. Also,
hydrogen addition will result in smaller arc diameters and lower radiation losses.
Plasmas of pure heavy gases such as Ar and N2 behave differently because of
their transport properties. Both have about the same electrical conductivity above
8000 K, though the thermal conductivity of nitrogen, with its dissociation at about
7000 K, is much higher than that of argon (see Part I chapter “▶ Transport Properties
of Gases Under Plasma Conditions”). The arc column of pure nitrogen plasma is
consequently constricted at temperatures where the electrical conductivity, σ N, is
sufficiently high to run the arc, which is not the case for pure argon arc. When adding
hydrogen, which diffuses fast in both gases, the constriction is mainly due to the
increase of the thermal conductivity of the plasma caused by the hydrogen addition.
However, the arc column in Ar-H2 is relatively less constricted than that in N2-H2. In
the latter case, for similar working conditions, the anode diameter is generally smaller
for N2-H2 plasma jets. This is illustrated in Fig. 17 representing the internal anode
nozzle designs of two commercial plasma torches from Öerlikon Metco, the PTF-4
and the 3MB torches. Figure 17a illustrates the PTF4 working with [33 slm (Ar) +
10 slm (H2)] (23.3 vol.% H2) for arc currents up to 600 A and Fig. 17b for the 3 MB
working with [5 slm (N2) + 4.5 slm (H2)] (11.4 vol.% H2) and arc currents up to 500 A.
The plasma jets obtained also depend strongly on the plasma-forming gas injec-
tors’ designs. For example, (Roumilhac et al. 1990a, b) compared two identical
torches working with the same arc current (500 A) and anode nozzle with radial
and axial injections of the plasma-forming gas. The corresponding arc voltages and
powers were 57 Vand 28.5 kW for the radial mode of injection and 67 Vand 33.5 kW
for the axial injection. The corresponding jets are presented in Fig. 18. These data
represent time-averaged integrated measurement since they were obtained using
emission spectroscopy with recording time longer than those corresponding to
plasma jet fluctuations occurring at frequencies between 10 and 100 kHz.

Fig. 17 Internal anode nozzle designs of two commercial plasma torches made by Öerlikon Metco:
(a) PTF4 for operation with Ar-H2 and (b) 3 MB for operation with N2-H2 (Nogues et al. 2008)
Plasma Spray Torches 21

Fig. 18 Temperature contours for a plasma jet [30 slm (Ar) + 12 slm (H2)] with an arc current of
500 A, at two different gas injection modes: (a) radial injection, (b) axial injection (Roumilhac et al.
1990b)

Fig. 19: Plasma jet temperature contours for two different nozzle shapes showing the longer and
broader jet with a (a) cylindrical nozzle and (b) divergent nozzle (Coudert et al. 1995)

3.2 Effect of Nozzle Design

The design of the anode nozzle also affects the length of the plasma jet. Numerous
nozzle designs exist, giving specific relationships between operating parameters and
plasma jet velocity and temperature distributions (Rahmane et al. 1998). Figure 19
(Coudert et al. 1995; Roumilhac et al. 1990a, b) shows, as an example, how a small
change in the nozzle design corresponding to a slight divergence of the nozzle
downstream of the arc attachment changes the jet temperature distributions. The
22 M. I. Boulos et al.

effect of a divergent anode nozzle or a Laval-type anode nozzle has been shown to
result in less cold gas entrainment (Henne et al. 2001) and higher deposition
efficiencies (Schwenk et al. 2004.)

3.3 Supersonic Atmospheric Plasma Spraying

By adapting the anode nozzle, it is possible to produce supersonic plasma jets at


atmospheric pressure in air. This is the case of the Praxair-TAFA model SG-100,
80 kW, multimode plasma spray torch, which can produce, with the adapted anode
nozzle either subsonic, Mach I (M = 1) and Mach II (M = 1.6) gas velocities. The
design of the anode nozzle depends on the type of flow that is desired. For example,
for subsonic flow, the SG-100 torch uses the nozzle shown in Fig. 20a, while for
sonic or supersonic flows, the designs are those shown in Fig. 20b and c.

Fig. 20 Praxair-TAFA model SG-100 plasma torch anode for (a) subsonic mode, (b) Mach I mode,
(c) Mach II mode. (Courtesy of Praxair-TAFA)
Plasma Spray Torches 23

To achieve a sonic flow under isentropic conditions (in a first approximation), the
ratio of the pressure inside the arc chamber, pch, to that of the ambient atmosphere,
pa, into which the jet emanates must be
    γ
Pch γ þ 1 ðγ1Þ
Rc ¼ ¼ (1)
Pa 2

where γ is the isentropic coefficient (γ = cp/cv) which varies with the plasma jet
temperature (Pateyron et al. 1996). For example, with hydrogen at room tempera-
ture, γ = 1.4; at 4000 K γ = 1.8; at 9000 K γ = 1.57; and at 14,000 K, γ = 1.2. Rc
varies from 1.77 to 2, and it is, thus, slightly easier to achieve sonic velocity with a
H2 plasma at 9000 K than at 14,000 K. The Mach II torch presents an arc chamber
with a step decrease in diameter allowing increasing the arc chamber pressure. Its
divergent nozzle allows achieving a supersonic jet. These Mach number values were
derived from measurements of the pressure drop across the nozzle. Figure 21 shows
the corresponding particle velocity, temperature, size, and flux distributions for YSZ

Fig. 21 Measurements of the lateral distribution of average values of (a) the spray particle velocity,
(b) the spray particle temperature, (c) the spray particle diameter, and (d) the relative particle flux for
two different anode nozzles shown in Fig. 20 (Fauchais et al. 2014)
24 M. I. Boulos et al.

particles. These measurements were achieved with DPV-2000 instrument (Tecnar


Automation, Montreal, Canada), at an axial distance of 80 mm from the anode
nozzle exit. Torch operating conditions at 800 A, 48 slm Ar + 12 slm He, carrier gas
flow 2.5 slm Ar, and powder YSZ (20%) at a feed rate of 5.8 g/min
Fincke et al. (1993a) have shown that the cold air entrainment is lower by a factor
of 2 in the supersonic jet compared to that for subsonic jets. At 100 mm from the
torch, WC-Co particle velocities at the axis are 380 m/s for the supersonic nozzle
vs. 200 m/s for the sonic nozzle. Particle temperatures are given as 3200 K for the
supersonic nozzle versus 2300 K for the sonic nozzle. It is interesting to note that the
plasma temperatures at the nozzle exit are higher with the sonic nozzle, but the
stronger entrainment leads to a faster reduction in temperatures, and at an axial
position 100 mm from the nozzle, the gas temperatures are about 1600 K for the
supersonic jet and about 1100 K for the sonic jet. Results presented in Fig. 21 for
both torches show particle characteristics in a plane given by the direction of particle
injection (y-direction) and perpendicular to the torch axis, as a function of the
distance from the torch axis, at an axial location 8 cm from the nozzle exit. Shown
are particle velocities, temperatures, diameters, and fluxes, all averaged over 1000
particles. The normalized flux shown is the ratio of particles detected at a specific
y-location in particles per second, divided by the particle flux at all y-locations. All
the data have been obtained with the DPV-2000 instrument. The following obser-
vations can be made:

• Highest velocities and temperatures are for the particles on the axis.
• Highest particle fluxes are below the axis for particle injection above the axis.
• Largest particles are farthest from the jet axis and have lower temperatures and
velocities.
• Supersonic anode provides higher velocities but lower temperatures.

3.4 Mini and Micro Plasma Spray Torches

A recent development in internal surface spraying considers the rotation of the


plasma torch, while the substrate remains stationary (Nicoll 1994; Barbezat 2006;
Roumilhac et al. 1991; Yushchenko et al. 1998). This implies, for example, that for
spraying the interior of cylinders of an engine block, the plasma torch with its
support structure must fit into the cylinder with typical diameters between 25 and
55 mm. The power levels of such torches range from 10 to 25 kW, which limits the
arc currents to 500 A. Minimum argon flow rates are around 15 slm with Ar-H2 or
Ar-He mixtures. An example of such a torch design is given in Fig. 22 (Roumilhac
et al. 1991). The overall dimensions of the torch are 54 mm diameter and 38.5 mm
length. The nozzle i.d. is between 5 and 8 mm and its length is 15 mm.
Micro-torches have been developed to coat small parts and components including
parts with fine details (Yushchenko et al. 1998). Such torches can spray narrow strips
of 1 to 3 mm in width and 0.2 to 0.5 mm in thickness on stainless steel sheets 0.5 mm
thick or on aluminum alloy sheets 1 mm thick. The torch nozzle i.d. is between 1 and
Plasma Spray Torches 25

Fig. 22 Schematic of the


PSA-44 mini plasma spray
torch (Roumilhac et al. 1991)

2 mm and the anode is external. The torch works with Ar at flow rates of 1.5–3 slm,
with arc currents of 3.5–50 A (power <2 kW). Laminar plasma jets are generated
resulting in low noise levels (30–50 dB against 100–130 dB for conventional torches).
A significant advance in this area was achieved by Öerlikon- Metco (Barbezat
2001; Barbezat and Landes 2000) as a result of their development of the RotaPlasma
spray torch shown in Fig. 23 for the coating of inside surface of the cylinders in an
internal combustion engine. The technology represented a breakthrough for automo-
bile engine manufacturers allowing for the significant reduction of the weight of the
engine using aluminum engine blocks, with the plasma spray coating of the internal
surfaces of the cylinders to improve their tribological properties. The most notable
feature of this torch design is the location of the spray nozzle with its axis perpen-
dicular to the axis of the torch body and the axis of rotation. The gas and cooling-
water connections are made through a rotating feed-through shaft. The torch can be
rotated at speeds up to 200 rpm inside cylinders with diameters of as small as 40 mm
and as large as 500 mm over a length up to 500 mm. Figure 24 shows a typical setup
for the coating of a four-cylinder automobile engine block in which two plasma
spraying systems are used simultaneously to reduce the total treatment time by half.

3.5 The Triplex Torch

As pointed out earlier, in the restrike mode, the length of the arc column is
continuously fluctuating in conventional plasma torches especially when hydrogen
is added to the heavy gas (Ar or N2) to improve the heat transfer to particles. These
fluctuations give rise to voltage fluctuations (ΔV) which can be as high as the mean
26 M. I. Boulos et al.

Fig. 23 Different views of the Öerlikon Metco RotaPlasma torch for coating inside surfaces of
cylinders. (Reproduced with kind permission of Öerlikon Metco AG, Switzerland)

Fig. 24 Öerlikon Metco


RotaPlasma torches used for
coating inside surfaces of a
four-cylinder block.
(Reproduced with kind
permission of Öerlikon Metco
AG, Switzerland)

arc voltage Vm with corresponding power fluctuations which affect the heating and
melting of the powder injected into the plasma stream. It has also been reported that
that electrode erosion increases with the increase of the arc current. A joint devel-
opment by the Universität der Bundeswehr, Munich, and Öerlikon Metco succeeded
in solving these two problems in their novel torch design commercially identified as
the “TriplexProTM-200” shown in Figs. 25 and 26 (Zierhut et al. 1998; Barbezat and
Landes 2000).
Plasma Spray Torches 27

Fig. 25 Öerlikon Metco TriplexProTM-200 plasma torch. (Reproduced with kind permission of
Öerlikon Metco AG, Switzerland)

Fig. 26 Representation of the


arc interaction in the Öerlikon
Metco TriplexProTM-200
plasma torch. (Reproduced
with kind permission of
Öerlikon Metco AG,
Switzerland)

The problem of voltage fluctuation was significantly reduced in the “Triplex”


torch using a “cascade arc chamber” design illustrated in Fig. 25a. This was achieved
through the introduction of several electrically insulated and water-cooled segments
(called neutrons) which form the wall of the arcing channel and force the arc to
maintain a longer and relatively constant arc length. This translates into a higher and
more stable torch voltage and a very stable operation especially with argon or argon-
helium mixtures, considering arc root fluctuations at the anode as being the same as
those obtained at the anode of a conventional torch with a stick-type cathode, voltage
fluctuations of about 10–15 V for Ar-He plasma. Considering that in a conven-
tional torch, the mean arc voltage Vm would be around 35 V, the ratio ΔV/Vm would
correspond to about (30–50%). Considering that with a Triplex torch, the mean arc
voltage is around 100 V; this ratio becomes (10–15%).
28 M. I. Boulos et al.

While the increase of the arc voltage allowed for the reduction of the arc current
for a given torch power, the problem of anode erosion was further reduced by
dividing the arc current into three separate arcs with three cathodes (Fig. 25a and
b) and one anode with three separate anode attachments. In order to produce a
centered high-temperature region, the arcs are forced toward each other by
narrowing the nozzle internal diameter while still avoiding joining them together
by the Lorentz force as shown in Fig. 26 (Marqués et al. 2009). However the
problem is that the self-magnetic field (attractive Lorentz force) favors merging of
the three individual arcs with a single anode attachment. Marqués et al. (2009) show
that the individual azimuthal fixation of the anodic arc roots can only be achieved
through Steenbeck’s minimum principle postulating a minimum arc length.
This torch also offers a significant increase in the powder flow rate. However,
powder injection is external to the plasma torch immediately downstream of the

anode, and three injectors are adjusted such that they are separated by 120 on the
circumference, preferably lining up with the direction between the anode attach-
ments of the three arcs. The improved arc stability, lower turbulence in the jet, and
the use of three injection systems allow not only better process control and longer
operating times; they also provide higher deposition rates. Anode nozzle i.ds. are
between 6 and 9 mm. Observing the plasma jet for small anode nozzle diameters
reveals that the jet consists of three lobes as shown in Fig. 27. Thus particles can be
injected either into the lobes or between the lobes, the latter injection being called
cage effect. (Mauer et al. 2011a, b) have studied the resulting plasma jets by emission
computer tomography using three CCD cameras rotating around the jet axis and the
interaction of the injection conditions on particle properties by DPV-2000 measure-
ments. The tomographic reconstruction of the gas temperature distribution allowed
them to relate the injector direction with the regions of high gas temperature and thus
of high gas velocity and viscosity. In all cases investigated, the most efficient particle
heating corresponded to an injection direction between two high-temperature lobes,
thus confirming the so-called cage effect (Mauer et al. 2011b). The lobe opposite to
the injector kept the larger particles properly in the jet center and prevented them

Fig. 27 Representation of the plasma jet with its three lobes in the Öerlikon Metco TriplexProTM-
200 plasma torch. (Reproduced with kind permission of Öerlikon Metco AG, Switzerland)
Plasma Spray Torches 29

from passing through the plume due to their too high momentum. As particle
velocities were found to vary only moderately, the particle temperatures were
more meaningful when the injection conditions were to be optimized. The results
showed that particle injection between the hot lobes, combined with an appropriate
carrier gas mass flow rate, allowed optimizing the process effectiveness as a function
of the size of the injected particles (Mauer et al. 2011b).
Recently Öerlikon Metco has launched the SimplexProTM series, which is of
similar design as the TriplexProTM torch except for the use of a single cathode
limiting it to a power rating of 60 kW.

3.6 Delta Plasma Spray Torch

As emphasized by (Marqués et al. 2009), GTV GmbH produces the so-called Delta
Gun plasma torch with fixed anode attachment points, originally developed in
Munich. The principle of this torch is illustrated in Fig. 28 where the arc is ignited
between a single cathode and three anodes, with an extended insulating area placed
between the cathode and the anodes. Anodes are obtained by dividing the anode ring
into three insulated pie-shaped pieces (Schwenk et al. 2007). When switching the
anode potential to the three anodes, a single arc is established inside the insulating
region, which splits up into three anode attachments near the exit. As with the
TriplexProTM, a triangular structure of the arc is expected, with the cage effect
providing a path for the particles to be injected into the plasma stream.

3.7 DC Plasma Torches with Axial Particle Injection

Since the early stages of development of plasma spray technology, it was recognized
that one of its challenges was the need for precise control of the injection of the
powders into the plasma stream to achieve optimum penetration into the jet and

Fig. 28 Principle of Delta


Gun (Marqués et al. 2009)
30 M. I. Boulos et al.

maximize the powder feed rate that could be tolerated without compromising the
melting of the powder or the quality of the coating. Most DC plasma torches use a
radial powder injection into the plasma, either internal in the anode nozzle or
external a few millimeters downstream of the nozzle exit plane. Either approach
requires precise control of the injection velocity. A too low velocity of injection
would limit the penetration of the powder stream into the plasma, with the
low-momentum particles (lower mass or lower velocity) remaining in the fringes
on the injection side and not being able to penetrate the plasma jet. On the other
hand, the use of an excessively high powder injection velocity would result in the
particles with higher mass and a higher injection velocity penetrating the plasma jet,
traversing it, and having most of their trajectories in the low-temperature fringes of
the jet on the side opposite of the jet with a minimal energy exchange between the
particles and the plasma. Obviously neither one of these situations will have the
optimal particle temperature and velocity for the coating formation. A detailed
discussion of the particle trajectory calculation in plasma flows can be found in
Part III, chapter “▶ Particles Trajectories and Plasma-Particle Interactions.”
Several efforts have, therefore, been undertaken to provide uniform heating of the
particles through the injection of the particles along the axis of the plasma stream.
While such a condition is easily achieved in RF inductively coupled plasma spraying
(see Part II, chapter “▶ Inductively Coupled RF Plasma Torches”), its use with DC
plasma sources requires specially designed plasma torches with multi-cathode
arrangement (Marqués et al. 2009). One of the early designs of such DC plasma
torches capable of axial powder injection was proposed by Fukanuma (1988). The
proposed design, schematically illustrated in Fig. 29, incorporates three individual
cathodes, each with axial plasma-forming gas injection, arranged around a central
powder feeding tube. The three arcs strike a common anode nozzle coaxial to the
feeding tube.
The same concept was further developed by Marantz and Herman (1992) and
Marantz et al. (1991) who coalesced the three plasma jets, resulting from the three
cathodes, into a single plasma stream before transferring the arc to a common
cylindrical anode nozzle coaxial with a central powder feeding probe as illustrated
in Fig. 30.

Fig. 29 Multielectrode torch


with three separate arcs
(Fukanuma 1988)
Plasma Spray Torches 31

Fig. 30 Multiple cathode


attachments converging into a
single anode (Marqués et al.
2009)

Fig. 31 Plasma torch with


three cathodes, common
anode embodiment (Marqués
et al. 2009)

Fig. 32 Schematic of Northwest Mettech Axial III central injection torch. (Courtesy of Northwest
Mettech)

Muehlberger et al. (1994) from Electro-Plasma Inc. used a similar three-plasma


torch system illustrated in Fig. 31, converging toward a common anode followed by
a diverging nozzle adapted for low-pressure plasma spraying.
One of the most commercially successful designs in this area is the Axial III
plasma torch developed by Northwest Mettech (Canada) (Moreau et al. 1995;
Burgess 2002). This torch design, illustrated by a sectional view in Fig. 32, consists
essentially of three mini-torches arranged in a converging pattern around a central
powder feeding tube. Each torch has its own cathode and anode. The plasma jets
32 M. I. Boulos et al.

from each of these torches merged into a single plasma plume transporting the
axially injected powder feed toward a common interchangeable water-cooled cylin-
drical nozzle with its axis aligned with the central powder injector. From the central
plasma mixing chamber, the plasma and particle flow are directed toward a nozzle
extension from which the plasma jet immerges uniformly loaded by the particles
which are heated, melted, and accelerated toward the substrate on which they are
deposited. Each plasma torch works with arc currents up to 250 A and a maximum
gas flow rate of 250 slm (e.g., Ar-N2-He). The total power of the three converging
torches is typically in the range 100–120 kW. The voltage fluctuations of the three
plasma torches are quite independent and have a low effect on the axially injected
particle trajectories or thermal history. The nozzle extension delays the mixing of the
plasma flow with ambient air, resulting in a slow decrease of the temperatures and
velocities of the plasma flow. The axial injection of the particles surrounded by the
three plasma jets reduces the particle sticking on the extension wall. Particle impact
velocities on the substrate can be almost as high as those obtained with HVOF guns
(Oberste Berghaus et al. 2006). The major advantage of such a torch is that the
coating quality is less sensitive to the morphology and particle size distribution of the
feed powder material used.
Another axial injection system was developed in Japan by Aeroplasma Corp.
(Suzuki et al. 2014). As schematically illustrated in Fig. 33, the torch comprises
essentially three plasma torches working exclusively with Ar as plasma gas. The
main “P-torch,” where P refers to positive polarity, has its nozzle as the cathode,
while the powder or suspension is injected through a hollow anode, of the stick-type
design with rounded extremity. Two cathode plasma torches, called “N-torches,” N
for negative polarity, are located downstream the “P-torch” with their jets blowing
orthogonally to its axis, forming the cathode of the whole setup. Depending on the
length of the plasma column, the overall torch voltage is relatively high and
accordingly the specific enthalpy of the plasma which allows for the melting of
ceramic particles. This torch has been mainly used to spray YSZ dry powders and
suspensions.

Fig. 33 Principle of the twin-cathode spray gun developed by Aeroplasma Corp., Japan
Plasma Spray Torches 33

4 High-Power DC Plasma Spray Torches

With the exception of Axial III torch, the DC plasma torches described in the
previous section are generally operated at power levels in the 50–60 kW range,
with powder feed rate of 4–6 kg/h. When large surfaces must be coated, for example,
in the paper and printing industry, torches with power levels up to 200–250 kW are
needed to allow powder feed rates as high as 15–20 kg/h in order to reduce the
coating time required to economically viable levels. Three of such torches are
described in this section.

4.1 Cascade Plasma Torches

As pointed out by Marquès et al. (2009), the cascade torch design, which can be
traced back to 1964, was one of the first spray torches with more power and less
fluctuations than conventional DC plasma spray torches. As schematically illustrated
in Fig. 34, the arc channel in the Advanced Plasma Gun (APG) consists of a series of
coaxial water-cooled segments, electrically insulated from each other and from the
anode (called neutrodes) forming a cascaded arc with a stick-type cathode on one
end and an annular ring anode serving as an exit nozzle on the opposite side of the
plasma torch. The arc can be initiated by having an initial breakdown between the
cathode and the segment closest to it (auxiliary anode), creating a plasma flow inside
the channel and then transferring the arc to the regular downstream anode.
With these torches, a much higher voltage between cathode and anode is obtained
with significantly reduced arc voltage fluctuations due to the limited movement of
the arc in the discharge cavity. Typically, mean arc voltages Vm can be higher than
150 V with voltage fluctuations ΔV of 40–50 V with molecular gases and especially
hydrogen. As the erosion of electrodes depends essentially on the arc current and
little on the voltage, power levels three to four times higher than with conventional
torches can be achieved with of course longer plasma jets.

Fig. 34 Schematic of the


cascade-design Advanced
Plasma Gun (APG) (Marqués
et al. 2009)
34 M. I. Boulos et al.

For spraying two torches of this type are used. The 100HE™ high power plasma
torch by Progressive Surface (USA) is a typical example of such torch design with a
stick-type cathode, and cylindrical anode separated by water-cooled neutrodes. It
works with voltages higher than 200 V and electric powers up to 105 kW, with
high-enthalpy ternary plasma gas mixtures (Ar, N2, H2, or He). Powder feed rates
up to 300 g/min (18 kg/h) can be achieved, and the deposition efficiency is between
50 and 80%.

4.2 PlazJet Torch

This torch developed by (Morishita (1991) is based on a concept which can be traced
back to the 1960s ( Zhukov (1977). As shown in Fig. 35, its design makes use of a
button-type cathode and a coaxial long tubular anode nozzle with a length-to-
diameter ratio of 5 up to 12. The plasma gas is introduced into the discharge cavity
with a strong swirl component. The anode nozzle shape may be either of a constant
diameter cylindrical bore, a conical bore, or a cylindrical bore with a step change in
internal diameter from 14 mm diameter inlet to 8 mm diameter exit. This latter
design corresponds to a fixed arc length torch design. The choice of different anode
designs allows tailoring the particle velocities which are highest with the conical
design, while the particle temperatures were highest with the step change design.
Gas flow rates can be varied over a wide range to up to 250–300 slm of nitrogen and
up to 150 slm of hydrogen. This range allows some adjustment of the specific gas
enthalpy vs. gas velocity and the residence time of the powders in the high-
temperature zone. The high gas flow rates result in a very long arc (of the order of
125 mm) with a voltage of 400 V at 500 A. While the plasma temperatures at these
flow rates are not very high (below 10,000 K), the plasma velocities are very high.
This high power allows spray rates of typically in the range of 10–20 kg/h of Cr2O3.

Fig. 35 Schematic of the Plazjet DC plasma torch (Morishita (1991)


Plasma Spray Torches 35

As the vortex of the plasma-forming gases is important, the particle injector must be
pointed slightly out of the torch axis.

4.3 Water-Stabilized Plasma Spray Torch

Water as a stabilizing medium for an electric arc has been used already in the 1920s
(Gerdien and Lotz 1922) mainly for reaching high plasma temperatures. The high
specific heat of water, steam, and hydrogen/oxygen plasma lead to very constricted
arcs, high arc voltages, and high temperatures. A water-stabilized plasma torch was
developed for plasma spraying applications by the Institute for Plasma Physics of the
Czech Academy of Sciences (Harabovsky et al. 1995, 1997). A schematic of the
basic torch design and photograph of the torch are given in Fig. 36. The torch design
is based on striking an arc between a central carbon electrode which acts as cathode
and an external, water-cooled rotating copper anode. Water is introduced into the arc
chamber with a strong swirl velocity component in several sections and exits the

Fig. 36 (a) Schematic and (b) picture of water swirl-stabilized arc torch with external rotating
anode. (With kind permission of Milan Harabovsky and the Institute of Plasma Physics of the Czech
Academy of Sciences)
36 M. I. Boulos et al.

torch at the anode nozzle level which allows a return of the unused water. The plasma
gas is provided by the evaporation of the water, and the plasma exits the torch
through a nozzle. The anode is rotating at high speed to distribute the high heat load
associated with the hydrogen-oxygen plasma. The graphite cathode is consumable
needing to be continuously fed during operation. The strong cooling of the arc
results in very high arc voltages (260–300 V) and high torch powers (80–200 kW)
at moderate arc currents (300–600 A). The plasma mean bulk temperature at the
torch exit can reach 28,000 K, with peak plasma velocities of 7000 m/s (Harabovsky
et al. 1995). These values allow spray rates in the range of 25–45 kg/h for ceramics
and 80–100 kg/h for metals, which are an order of magnitude higher than those for
conventional plasma torches operating at 40–50 kW and still significantly higher
than that of a high-power gas torch operating at 200 kW with a nitrogen-hydrogen
mixture. The early models of the torch were relatively cumbersome and not designed
to spray small parts with complex shapes. Newer designs illustrated in Fig. 37 use a
fixed nonconsumable cathode with an argon gas sheath in the upstream end of the
torch (hybrid water torch) which significantly increased operation times
(Harabovsky 2002).
A comparison between the operating power range of water-stabilized and
gas-stabilized conventional plasma torches, as function of the mass flow rate of the
water vapor or plasma gas, is given in Fig. 38. This shows that for an equivalent
power rating, considerable lower mass flow rates of water vapor are needed in a
water-stabilized torch compared to the plasma gas flow rates needed with conven-
tional torches. This has the direct consequence of achieving higher powers and
higher specific enthalpy levels with water-stabilized torches. For hybrid torches
the Ar flow rate needed for the sheathing of the cathode is of the order of 2 g/s
which is not negligible. The high deposition rates achieved with water-stabilized
torches make them particularly useful for manufacturing of freestanding parts or for
coating large areas with ceramics such as alumina or YSZ.

Fig. 37 Schematic of water-stabilized arc torch with fixed cathode surrounded by argon sheath
flow and external rotating anode. (With kind permission of Milan Harabovsky)
Plasma Spray Torches 37

Fig. 38 Torch power as


function of the mass flow rate
of water vapor or plasma gas
for water-stabilized, hybrid,
and gas-stabilized torches.
(With kind permission of
Milan Harabovsky)

Fig. 39 Schematic diagram


of the CACT torch with a
nonconsumable graphite
cathode (Pershin et al. 2013)

4.4 CACT Plasma Torch

A water-cooled DC plasma torch for operation with a mixture of CO2/CH4 as plasma


gas was developed at the Centre for Advanced Coating Technologies (CACT) at the
University of Toronto (Pershin et al. 2013). As schematically illustrated in Fig. 39,
the cathode is made of graphite, which is generally a consumable electrode material
with its tendency to sublime under the high heat flux conditions at the arc attachment
regions. Using a mixture of CO2/CH4 as plasma-forming gas, it was observed that
carbon nanotubes (CNTs) and nanoparticles are deposited on the cathode tip in the
form of clusters which are evenly distributed on the emitting surface with mean
spacing of 2.7 μm between them. These clusters partially compensate for the erosion
of the carbon electrode and considerably extend its lifetime. The movement of the
anode root attachment on the surface of the electrode is achieved using an external
magnetic field coil placed around the anode as shown in Fig. 39. This allows for a
better distribution of heat load on the anode surface, a reduction of anode erosion,
38 M. I. Boulos et al.

and the extension of its lifetime. With CO2/CH4 as plasma-forming gas, the torch
voltage is high, about 250 V for arc currents between 200 and 600 A. Torches based
on this design concept were tested at power levels of 40 and 200 kW. The specific
enthalpy of the plasma in these cases was slightly higher than that obtained with
nitrogen which has similar thermal conductivity as CO2/CH4 mixtures. They also
created favorable conditions for the plasma treatment of refractory materials such as
carbides (Pershin et al. 2013).

5 Plasma Spray System Configurations

5.1 Atmospheric Plasma Spraying (APS)

To limit the risks to operating personnel, plasma spraying operations should be


carried out in a well-designed and well-equipped spray booth such as the one
illustrated in Figs. 40 and 41.
The spray booth by itself is an enclosure that works with the ventilation systems
to provide an effective control of the major hazards described previously. Its size
must be sufficient to allow the operator to prepare the torch and parts to be sprayed
and to maintain the robotic equipment. It must be designed in such a way that during
spraying the operator needs not be inside but controls the spray operation from a
control panel outside the booth as illustrated in Fig. 41. The visual control of the
spray process is achieved through a window with proper radiation filters (which can

Fig. 40 Schematic of a midrange spray booth with automatic handling system


Plasma Spray Torches 39

Fig. 41 Typical spray coating facility. (Reproduced with kind permission of Öerlikon Metco AG,
Switzerland)

include pulldown shades) and soundproofing or with a CCD camera and a TV


screen. To get rid of dusts, the use of laminar flows in the booth is strongly
recommended since turbulent flows often create areas where the dust is trapped. In
most cases the pressure within the booth is slightly negative. The booth should also
have an entry timer to allow the ventilation system to continue to clean the air of
hazardous dust following the completion of the plasma spray operations.

5.2 Controlled Atmosphere Plasma Spraying (CAPS)

DC plasma jets are known for their high-energy density and high plasma temperature
(above 12,000 K) and velocities (in the hundreds of m/s) with steep temperature and
velocity gradients. As the plasma jets immerge into a stagnant ambient atmosphere, a
strong shear layer will develop in the interface between the jet and the ambient gas.
The structure of such a flow has been the subject of intense studies in the 1980s and
1990s (Pfender et al. 1991; Russ et al. 1994a, b; Huang et al. 1995) due to its impact
on the overall jet behavior and the resulting flow and temperature fields. Based on
schlieren photographs, emission spectroscopy, and enthalpy probe concentration
40 M. I. Boulos et al.

Fig. 42 Schematic representation of the main regions of DC plasma jet showing the rolling of the
flow around the nozzle exit, cold eddy engulfment, and breakdown followed by turbulent flow
generation (Pfender et al. 1991)

field measurements for an argon DC plasma jets in air, Pfender et al. (1991)
concluded that the large velocity difference between the jet and the ambient atmo-
sphere causes rolling up of the flow around the nozzle exit into the ring vortex which
is pulled downstream by the flow, allowing the process to repeat itself again at the
nozzle exit. Adjacently formed vortex rings at the outer edge of the jet have the
tendency to coalesce, forming large vortices. As schematically represented in
Fig. 42, the distorted vortex rings start entangling themselves with the adjacent
rings, finally resulting in total breakdown of the vortex structure into large-scale
eddies and the onset of turbulent flow. This process results in the first large-scale
engulfment of the ambient gas into the jet flow. Some entrainment also takes place
during the roll-up process of the jet shear layer. The eddies of cold gas traveling in
the axial direction at lower velocities than the flow continuously break down into
smaller eddies, while diffusion takes place on the molecular level at all eddy
boundaries. The mixing and diffusion process eventually reaches the centerline of
the jet ending its laminar core. The overall consequence of the process is the intense
mixing of the ambient gas into the plasma jet stream as shown in Fig. 43 giving
nitrogen concentration contours for an argon DC plasma jet discharge in ambient air
at two power levels (5 and 10 kW). Such strong interaction between the plasma jet
and the ambient atmosphere results in significant quenching of the jet as shown in
Fig. 44 giving the isotherms of a plasma jet under identical conditions in a chamber
with controlled atmosphere containing, respectively, argon, nitrogen, and air. A
rather long and wide jet is obtained with pure Ar (Fig. 44a) while nitrogen in
between both (Fig. 44b). The jet is rather short and slim when emanating into air
(Fig. 44c).
Plasma Spray Torches 41

Fig. 43 Contour plots of


nitrogen concentration,
normalized with respect to its
value in ambient air, for argon
plasma jets in air at two power
levels: (a) I = 250 A, 5 kW,
(b) I = 500 A, 10 kW (Pfender
et al. 1991)

It may be noted that the degree of cooling of the plasma jet is dependent on the
entrained ambient gas flow  rate and its thermophysical properties. This can be
represented by ( m _ g cp ΔT , where m_ g is the mass flow rate of the ambient gas
entrained by the jet; cp is the specific heat of the ambient gas; and ΔT is the
temperature rise of the ambient gas up to the local mixing gas temperature. In
most cases the cold gas will not penetrate the core of the jet, i.e., it is mixed in
zones where the plasma temperature remains below 12,000 K at most. Thus,
ionization of the cold gas will be negligible. However, when a diatomic gas is
entrained, the temperatures will be sufficient to dissociate it with a significant
increase of the specific heat of the gas in this temperature range. Thus, cooling is
by far more important in the case of molecular gases, especially when air is entrained
(oxygen dissociates at T > 2500 K and nitrogen at T > 7000 K).
The effect of entrained ambient gas on the properties of the coatings is not limited
to thermal effect but can also have chemical effects when dealing with the coating
with reactive metals or non-oxide ceramics (carbides, nitrides, borides, etc.). In these
cases, air plasma spraying is not always acceptable due to the in-flight oxidation of
the surface of the powders and the possible integration of such oxide particles in the
coating. Vacuum plasma spraying, described in the next section, is best adapted for
the spraying of metals of alloys.
Controlled atmosphere plasma spraying (CAPS) developed in the 1960s uses
essentially vacuum-tight chambers with a vacuum pump, for the initial evaluation of
42 M. I. Boulos et al.

Fig. 44 Contour plots of


temperature field in an argon
DC plasma jet (34 kW)
discharged into different
ambient atmospheres: (a)
argon atmosphere, (b)
nitrogen atmosphere, and (c)
air (Roumilhac et al. 1990a, b,
1991)

the chamber, before backfilling it with an inert gas, usually argon, to atmospheric
pressure or higher. The oxygen content can be lower than 7 ppm if a liquid argon
source is used (Freslon 1995). As shown earlier, plasma jets in argon atmosphere are
broader and longer by a factor 1.5 to 2 compared to the plasma jet in air (Roumilhac
et al. 1991). CAPS is normally operated at pressures in the range of 70 to 300 kPa for
the spraying of materials which are very sensitive to oxidation such as carbides,
borides and refractory metals. Few tests have been performed using chambers at
pressures over 300 kPa or 3 atm (Jäger et al. 1992). The plasma jets are shorter than
at atmospheric pressure, and electrode erosion becomes more important due to the
increased radiation from the plasma column. When increasing the pressure, the arc
voltage also increases since the arc column constricts and its temperature rises,
resulting in higher losses to which the arc responds by an increase of the electric
field strength (arc voltage). As the heat losses increase, the specific enthalpy of
plasma jet shows a slight drop. Chemical effects of the plasma gas and/or the
ambient gas, on the quality of the coating, cannot be eliminated using CAPS since
when spraying reactive metals such as titanium with Ar-N2 plasma-forming gas, in
controlled argon or nitrogen atmosphere chamber, at different pressures, the nitrogen
Plasma Spray Torches 43

Table 2 Nitrogen content of titanium plasma coating sprayed using Ar-N2 plasma-forming gas in
an argon or nitrogen CAPS chamber (Jäger et al. 1992)
Chamber atmosphere Ar N2
Chamber pressure (kPa) 20 119 20 119
Coating nitrogen content (wt.%) 2.3 6.7 2.8 12.4

content of the coating in the form of TiN can be significant as shown in Table 2
(Jäger et al. 1992).
Similarly (Guipont et al. 2002, 2010; Espanol et al. 2002) have sprayed hydroxy-
apatite with argon plasma working in argon atmosphere at pressure up to 0.3 MPa.
They have shown that the decomposition level can be tailored without using N2 or
H2. Coatings present highly soluble as well as less soluble (crystalline) characteris-
tics. The nature of the ceramic composite with multiphases can be adjusted through
the chamber pressure. Alumina coatings sprayed in Ar atmosphere at different
pressures exceeding atmospheric pressure have also been studied (Ma et al. 2002).

5.3 Vacuum Plasma Spraying (VPS)

Low-pressure plasma spraying (LPPS) also identified as “vacuum plasma spraying”


(VPS) was developed in 1974 by Müehlberger (Electro-Plasma Inc., now Öerlikon
Metco Inc.) (Muehlberger 1988; Meyer and Hawley 1991). It is mostly performed at
pressures between 5 and 70 kPa, allowing for the coating of small- to medium-sized
parts of complex shapes with high-density coatings of metals and alloys including
refractory metals. It provides for a tight control on the ambient gas composition
during the coating process avoiding in-flight oxidation of the sprayed particles
insuring an oxide-free deposit. In VPS, the substrate can be maintained at high
temperatures (up to 950  C for superalloys) without oxidation. Such temperatures
promote interdiffusion, thus enhancing significantly the coating adhesion. The
principal limitation of VPS is the size of the parts that can be coated and the melting
temperature of the material to be sprayed which excludes ceramic materials partic-
ularly non-oxide ceramics with melting temperatures above 3000 K. The high capital
investment needed for such installations also imposes an important limitation on the
wider used of the technology.
As shown in Fig. 45, vacuum plasma spraying is carried out in large vacuum
chamber with volumes of a few m3 up to 10–20 m3 which houses the substrate, often
on a carousel holding more than one part to be coated, a DC plasma torch, and a
robotic torch manipulator. The chamber is normally water-cooled with a large access
door for ease of servicing, placing the parts to be coated on the carousel substrate
holder and retrieving them at the end of the coating cycle. The gas exit port of the
chamber is to be connected to an efficient dust collection filter followed by a high-
volume pumping station capable of the initial evacuation of the chamber down to
1 kPa or less in a reasonable time (a few minutes) and to maintain the chamber
pressure under the spraying cycle under soft vacuum (5 > p > 70 kPa). It is to be
44 M. I. Boulos et al.

Fig. 45 Schematic of vacuum plasma spraying chamber

noted that under such pressure range of operation, the torch nozzle must be adapted
for supersonic jet velocities to reduce or avoid the formation of diamond shock
waves in the jet. A typical non-optimized profile of such nozzles is schematically
illustrated in Fig. 46a. Subsequent photographs given in Fig. 46b–e show images of
the plasma jet under progressively reduced pressure of 100 kPa (atmospheric
pressure), down to 39.4 kPa, 6.6 kPa, and 5.2 kPa. These show a systematic increase
of the length of the plasma jet length with the decrease of the chamber pressure due
to the reduction of turbulence in the jet fringes and ambient gas entrainment into the
flow. The effect is accompanied by the increase of the jet velocity and a substantial
reduction of its temperature (Roumilhac et al. 1990b).
Typical results of modeling studies of plasma jets under VPS conditions are given
in Figs. 47 and 48. These show, respectively, the axial velocity and temperature
distributions along the centerline of an argon plasma jet (8 kW) with 11.6 g/min
argon flow with an internal nozzle diameter of 8 mm issuing into an argon environ-
ment at three different chamber pressures. As noted in Fig. 47, the axial velocity of
the jet at its exit level of the plasma torch nozzle increases with the decrease of the
chamber pressure reaching close to 2300 m/s at a pressure of 8 kPa. Beyond the
initial core of the jet, which is 50–60 mm in length, the axial velocity of the of the
flow drops gradually with the distance from the torch nozzle exit level. The
corresponding temperature distribution along the centerline of the jet is given in
Fig. 48. These show a very rapid drop of the plasma jet temperature at near
atmospheric pressure (100 kPa) with a considerably lower temperature decay
along the centerline of the jet with the decrease of the chamber pressure. The
lower axial velocity and temperature gradients are due to the lower large-scale
turbulence and less cold gas entrainment, a consequence of the smaller density
gradients resulting in less turbulent shear layers. The lower gas densities, however,
result in lower electron-ion recombination rates, and deviations from equilibrium
composition are to be expected.
Plasma Spray Torches 45

Fig. 46 Photographs of (a) DC plasma nozzle adapted for operation under soft vacuum conditions
and (b–e) photographs of plasma jets under different chamber pressures, 5.2 kPa, 6.6 kPa, 39.4 kPa,
and 100 kPa (Lang et al. 2001)

Fig. 47 Axial velocity


profiles for an argon plasma
jet in a VPS chamber at
different pressures
46 M. I. Boulos et al.

Fig. 48 Axial temperature


profiles for an argon plasma
jet in a VPS chamber at
different pressures

For compressible nozzle flows reaching supersonic speeds inside the nozzle,
shock diamonds are usually observed outside the nozzle (see schematic in
Fig. 46a). These shock diamonds occur when the pressure inside the jet is different
from the surrounding pressure, and they lead to sudden velocity reductions and can
result in a dispersion of the particle trajectories. The jet can be overexpanded, i.e.,
with a pressure inside the jet lower than the environment pressure, or under-
expanded with a jet pressure higher than the environment pressure. Therefore,
through the interaction with the environment, the jet is compressed or expanded,
but an overcompensation may occur resulting in another under-expansion/
overexpansion of the flow. A complicated flow structure evolves with shock waves
starting at the nozzle exit being reflected at the discontinuity between the jet and
environment, and these reflected shock waves create the shock diamonds. The more
shock diamonds and the clearer they are visible, the worse the fluid dynamic
condition. The use of a properly designed Laval nozzle either as part of the anode
or as an attachment to the anode can significantly improve the results by avoiding the
shock structures immediately downstream of the anode nozzle exit (Mayr and Henne
1988). A schematic of a torch with a Laval nozzle attachment is shown in Fig. 49
(Mayr and Henne 1988). The Laval nozzle attachment offers an increase of particle
velocities by 30–50% and focuses the particle jet into a direction more parallel to
the torch axis (Mayr and Henne 1988; Henne et al. 1993), resulting in a higher
deposition efficiency (Rahmane et al. 1998). Therefore, the torch current and power
level can be reduced.
VPS offers the following advantages compared to the atmospheric plasma
spraying:

(i) The controlled atmosphere prevents oxidation of the spray powders in-flight
and on the substrate, an important consideration with metal spraying. Finally,
the oxidation in VPS corresponds to oxide content slightly higher than the
oxide content of sprayed powder.
Plasma Spray Torches 47

Fig. 49 Schematic of torch with Laval nozzle attachment for improved pressure equilibration
(Mayr and Henne 1988)

(ii) The higher-velocity plasma jet (2000–3000 m/s) results in higher particle
velocities with the possibility of higher coating densities. Typical particle
velocities have been measured of 800–900 m/s for alumina and of
450–650 m/s for YSZ (Fauchais et al. 2014).
(iii) The lower density gradients between the plasma jet and the surroundings result
in more stable jets and less cold gas entrainment, with the consequence of more
uniform particle heating and acceleration and less divergence of the particle
trajectories.
(iv) The uniformity of the jet temperature and velocity profiles can be further
improved by attaching a Laval nozzle to the anode.
(v) At the lower pressures, the heat transfer rates to the particles are reduced;
however, the longer jets increase the residence time in the plasma; these
reduced heating rates result in less superheating of the particle surface and
lower temperature differences between particle surface and the particle center.
(vi) The cleaning and heating processes of the substrate are relatively simple and
efficient. As illustrated in Fig. 50a, the cleaning of the substrate can be achieved
by striking a low-current transferred arc between the plasma torch in operation
with no powder injected, which acts as anode, and the substrate serving as the
cathode. The formation of cathode spots moving rapidly over the substrate
surface results in evaporation of impurities, in particular of oxides (Itoh et al.
1990). It is also possible, once cleaning is completed, to preheat the substrate by
reversing the polarity, as shown in Fig. 50b., having the substrate positively
biased with respect to the torch which acts then as cathode. Passing a current of
a few tens to hundred amperes, the substrate can be preheated without oxidation
to the desired temperature. Substrate preheating to about 900 oC can promote
diffusion bonding of, e.g., an environmental barrier coating with a superalloy
substrate.
48 M. I. Boulos et al.

Fig. 50 Schematic of transferred arc setting between the torch and the substrate. (a) Substrate
cleaning in reverse polarity, (b) substrate preheating in straight polarity

Of course, these advantages come with a price: besides the large water-cooled
chamber with the associated pumping installation, the system requires filters before
the pumps, heat exchangers for cooling the exhaust, inert gas manifold for back-
filling the chamber, robotic systems adapted to vacuum environment for moving the
part, and/or the torch, shielding of all cables inside the chamber against exposure to
heat and particle fluxes (Meyer and Hawley 1991). These required features raise the
cost of a VPS system significantly, and this increase can be an order of magnitude or
more compared to an APS system with comparable power. One also needs to
consider the fluid dynamics inside the chamber to assure that no cold particle
deposition occurs due to recirculating flows. Moreover, for industrial use a transfer
chamber (load lock) is used that would allow to adjust the pressure, after loading the
part to be coated at atmospheric pressure to the vacuum chamber pressure. This way
the pump-down of the main chamber after every spray process can be avoided.
Clearly, these additional features and the associated cost increases limit the applica-
tions of VPS coatings to high added-value parts compared to APS coated parts.
Torch operating conditions are comparable to APS torches in terms of arc currents,
plasma gas selections and flow rates, power levels, and powder flow rates. However,
the torch substrate distance is typically in the range between 250 and 500 mm.

5.4 Plasma Spraying-Physical Vapor Deposition (PS-PVD)

New developments by Öerlikon Metco show that deposition at very lower pressures,
using a combination of plasma spraying and physical vapor deposition processes,
Plasma Spray Torches 49

results in excellent coating quality. The process known as low-pressure plasma


spraying-Thin Film deposition (ChamPro1 LPPS1-TF) is in direct competition
with the electron beam-physical vapor deposition (EB-PVD) process used for the
manufacturing of high-quality thermal barrier coatings (TBCs) to improve the life
and performance of turbine components. In contrast to conventional plasma spray
technology, the EB-PVD process is operated at much lower pressures (0.1–5 Pa)
with the high-energy electron beam used for the evaporation of the coating material
which is then condensed on the substrate forming the coating. The deposition rates
are typically rather low and the necessary equipment capital intensive.
The Öerlikon Metco (ChamPro1 LPPS1-TF) deposition process is based on
their expertise in low-pressure plasma spraying using a high-power plasma spray
torch (180 kW–3000 A) with gas flow rates up to 200 slm, working at pressures as
low as 0.2 kPa (2 mbar). Under such low-pressure conditions, the plasma jet reaches
more than 2 m in length and up to 0.4 m in diameter as shown in Fig. 51. According
to Öerlikon Metco, the process has the following characteristics:

• Deposition capability from the vapor phase


• Reduced heat load to the substrate
• Very flexible operating conditions

Figure 51 Photograph of a
2 m long plasma plume in a
(ChamPro1 LPPS1-TF)
process. (Reproduced with
kind permission of Öerlikon
Metco AG, Switzerland)
50 M. I. Boulos et al.

Figure 52 (a) SEM image of a vapor-phase deposited, columnar TBC top coat with thickness of
150 μm obtained using (ChamPro1 LPPS1-TF). (b) A magnified view of the columnar structure.
(Reproduced with kind permission of Öerlikon Metco AG, Switzerland)

• Application of thin full-coverage coatings of a few micrometers thick


• Fast uniform coatings over large areas at rates of (10 μm/min/m2)

Electron micrographs of typical thermal barrier coatings (TBDs) obtained using


the LPPS@-TF deposition process are given in Fig. 52. These show the typical
columnar structure normally obtained using EB-PVD processes that are more strain
tolerant at high temperature and stresses compared to standard low-pressure plasma
sprayed (LPPS) coatings. The advantage of the (ChamPro1 LPPS1-TF) process
over the EB-PVD is mostly due to its significantly higher speed of deposition. It
must be kept in mind, however, that both the EB-PVD and (ChamPro1 LPPS1-TF)
coating techniques are of interest for parts of high added value, which limits their
potential to a small fraction of the overall thermal spray market.

Nomenclature
Latin Alphabet
cp Specific heat at constant pressure (J/kg.K)
cv Specific heat at constant volume (J/kg.K)
E Electric field in the arc column (V/m)
Ex Electric field normal to the anode surface (V/m)
ED Dissociation energy (J)
h Specific enthalpy (J/kg)
h Mass mean specific enthalpy (J/kg)
I Arc current (A)
j Electron current density (A/m2)
k Boltzmann constant (k=1.38  1023 J/K)
ṁ p Mass flow rate of the plasma gas (kg/s)
ṁ g Mass flow rate of the entrained atmosphere (kg/s)
ne Electron number density (m3)
pa Ambient pressure (Pa)
Plasma Spray Torches 51

pch Pressure inside the arc chamber (Pa)


Rc Pressure ratio [Rc = pch/pa]
T Heavy species temperature (K)
Te Electron temperature (K)
v Plasma velocity (m/s)
vmax Maximum plasma velocity (m/s)
V Plasma torch voltage (V)
Vm Mean of fluctuating voltage (V)
Greek Alphabet
Δh Increase in average specific enthalpy (J/kg)
ΔT Temperature difference (K)
ΔV Amplitude of voltage fluctuations (V)
ΔZ Length variation (m)
εi Ionization energy (J)
γ Isentropic coefficient (γ = cp/cv)
ρ Mass density (kg/m3)
ηth Plasma torch thermal efficiency

References
Barbezat G (2001) The internal plasma spraying on powerful technology for the aerospace and
automotive industries. In: Berndt CC, Khor KA, Lugscheider EF (eds) Proceedings of ITSC,
Singapore. ASM International, Materials Park, pp 135–140
Barbezat G (2006) Application of thermal spraying in the automobile industry. Surf Coat Technol
201:2028–2031
Barbezat G, Landes K (2000) Plasma technology TRIPLEX for the deposition of ceramic coatings
in the industry. In: Berndt CC (ed) Proceedings of 1st ITSC, Montreal, Canada. ASM Interna-
tional, Materials Park, pp 881–885
Burgess A (2002) Hastelloy C-276 parameter study using the axial III plasma spray system.
In: Lugsheider E (ed) Proceedings of ITSC-2002, Essen, Germany. ASM International,
Materials Park, pp 516–518
Chandra S, Fauchais PL (2009) Formation of solid splats during thermal spray deposition. J Therm
Spray Technol 18:148–180
Coudert JF, Planche MP, Fauchais P (1995) Velocity measurement of D.C. plasma based on arc root
fluctuations. Plasma Chem Plasma Process 15(1):47–70
Davis JR (ed) (2004) Handbook of thermal spray technology. ASM International Materials Park,
Cleveland, OH, USA
Ducos M, Durand JP (2001) Thermal coatings in Europe, a business perspective. In: Berndt CC,
Khor KH, Lugsheider E (eds) Thermal spray 2001. ASM International Materials Park, pp
1267–1276
Espanol M, Guipont V, Khor KA, Jeandin M, Llorca Isern N (2002) Effect of heat treatment on high
pressure plasma sprayed hydroxyapatite coatings. Surf Eng 18(3):213–218
Fauchais PL (2004) Understanding plasma spraying. J Phys D Appl Phys 37:86–108
Fauchais P, Heberlein J, Boulos M (2014) Thermal spray fundamentals, from powder to part.
Springer, New York. 1550 pages
Fauchais P, Vardelle M, Goutier S (2016) Latest researches advances of plasma spraying; from splat
to coating. J Therm Spray Technol 25:1534–1553
52 M. I. Boulos et al.

Fincke J, Swank WD, Haggard DC (1993) Atmospheric plasma spraying of WC:Co. In: Harry J
(ed) Proceedings of ISPC, University of Loughborough, Loughborough, UK. pp 145–149
Freslon A (1995) Plasma spraying at a controlled temperature and atmosphere. In: Berndt CC,
Sampath S (eds) Proceedings of NTSC-1995, Houston, TX. ASM International, Materials Park,
pp 57–63
Fukanuma H (1988) Japanese Patent 230,300 JP. 24 Apr 1988
Gerdien H, Lotz A (1922) Wiss Veröff Siemens-Konz. 489–4
Goutier S, Vardelle M, Labbe JC, Fauchais P (2011) Flattening and cooling of millimeter- and
micrometer-sized alumina drops. J Therm Spray Technol 20:59–67
Goutier S, Vardelle M, Fauchais P (2013) Comparison between metallic and ceramic splats:
influence of viscosity and kinetic energy on the particle flattening, surf. Coat Technol 657–668
Guipont V, Espanol M, Borit F, Llorca-Isern N, Jeandin M, Khor KA, Cheang P (2002) High-
pressure plasma spraying of hydroxyapatite powders. Mater Sci Eng A 325(1–2):9–18
Guipont V, Bansard S, Jeandin M, Khor KA, Nivard M, Berthe L, Cuq-Lelandais JP, Boustie M
(2010) Bond strength determination of hydroxyapatite coatings on Ti-6Al-4V substrates using
the LAser Shock Adhesion Test (LASAT). J Biomed Mater Res A 95A(4):1096–1104
Harabovsky M (2002) Generation of thermal plasmas in liquid and hybrid DC arc torches. Pure
Appl Chem 74(3):429–433
Harabovsky M, Kopecky V, Sember V (1995) Water stabilized arc as a source of thermal plasma.
In: Fauchais P (ed) Proceedings of international symposium heat and mass transfer under plasma
conditions, Cesme, Turkey. Begell House, New York, pp 91–98
Harabovsky M, Konrad M, Kopecky V, Semùber V (1997) Processes and properties of electric arc
stabilized by water vortex. IEEE Trans Plasma Sci 25(5):833–839
Henne R, Mayr W, Reusch A (1993) Influence of nozzle geometry on particle behavior and coating
quality in high velocity VPS. In: Proceedings of ITSC TS93, Aachen, Germany. DVS, Germany,
pp 7–11
Henne R, Bouyer E, Borck V, Schiller G (2001) Influence of anode nozzle and external torch
contour on the quality of the atmospheric DC plasma spray process. In: Berndt CC, Khor KA,
Lugsheider E (eds) Proceedings of ITSC-2001, Singapore. ASM International, Materials Park,
pp 471–478
Huang PC, Heberlein J, Pfender E (1995) A two-fluid model of turbulence for a thermal plasma jet.
Plasma Chem Plasma Process 15(1):25–46
Itoh A, Takeda K, Itoh M, Koga M (1990) Pretreatments of substrates by using reversed transferred
arc in low pressure plasma spray. In: Bernecki T (ed) Thermal spray research and application.
ASM International, Materials Park, pp 245–252
Jäger DA, Stöver D, Schlump W (1992) High pressure plasma spraying in controlled atmosphere up
to 2 bars. In: Berndt CC (ed) Proceedings of ITSC-1992., Orlando, FL. ASM International,
Material Park, pp 69–74
Lang M, Henne R, Schaper S, Schiller G (2001) Development and characterization of vacuum
plasma sprayed thin film solid oxide fuel cells. J Therm Spray Technol 10(4):618–625
Leblanc L, Moreau C (2002) The long-term stability of plasma spraying. J Therm Spray Technol
11:380–386
Ma X-Q, Borit F, Guipont V, Jeandin M (2002) Thin alumina coating deposition by using controlled
atmosphere plasma spray system. J Adv Mater 34(4):52–57
Marantz DR, Herman H (1992) Plasma spray gun and method of use. US Patent 5,144,110
Marantz DR, Kowalsky KA, Marantz D (1991) Wire-arc-plasma spray process basic principles and
its versatility. In: Bernecki TF (ed) Proceedings of the fourth NTSC, Pittsburgh, Pennsylvania.
ASM International, Materials Park, pp 381–387
Marqués J-L, Forster G, Schein J (2009) Multi-electrode plasma torches: motivation for develop-
ment and current state-of-the-art. Open Plasma Phys J 2:89–98
Mauer G, Vaßen R, Stöver D, Kirner S, Marqués JL, Zimmermann S, Forster G, Schein J (2011a)
Improving power injection in plasma spraying by optical diagnostics of the plasma and particle
characterization. J Therm Spray Technol 20(1–2):3–11
Plasma Spray Torches 53

Mauer G, Vaßen R, Stöver D (2011b) Plasma and particle temperature measurements in thermal
spray: approaches and applications. J Therm Spray Technol 20(3):391–406
Mayr W, Henne R (1988) Investigation of a VPS burner with Laval nozzel using an automated laser
doppler measuring system. In: Eschnauer H, Huber P, Nicoll AR, Sandmeier S (eds) Pro-
ceedings of 1st plasma-technik symposium, Lucerne, Switzerland. Plasma Technik, Wohlen,
pp 87–97
Meyer PJ, Hawley D (1991) LPPS production systems. In: Bernecki TF (ed) Proceedings of 4th
NTSC-1991., Pittsburgh, PA. ASM International, Materials Park, pp 29–38
Moreau C, Gougeon P, Burgess A, Ross D (1995) Characterization of particle flows in an axial
injection plasma torch. In: Berndt CC, Sampath S (eds) Proceedings of 8th NTSC-1995,
Houston, Texas. ASM International, Materials Park, pp 141–147
Morishita T (1991) Coatings by 250 kW plasma jet spray system. In: Blum-Sandmeier S,
Eschnauer H, Huber P, Nicoll A (eds) Proceedings of 2nd Plasma Technik symposium, Luzern,
Switzerland. Plasma Technik, Wohlen, pp 137–145
Muehlberger E (1988) Industrial plasma processing technology. In: Eschnauer H, Huber P, Nicoll
AR, Sondermeier S (eds) Proceedings of 1st Plasma-Technik Symposium, Lucerne,
Switzerland. Plasma-Technik AG, Wohlen, pp 105–118
Muehlberger E, Muehlberger SE, Sickinger A et al (1994) Modular segmented cathode plasma
generator. US Patent 5,298,835
Nicoll AR (1994) Production plasma spraying in automotive industry: a European viewpoint. In:
Berndt CC, Sampath S (eds) Thermal spray industrial applications, proceedings of the 7th
NTSC-1994. ASM International, Materials Park. 7
Nogues E, Vardelle M, Fauchais P, Granger P (2008) Arc voltage fluctuations: comparison between
two plasma torch types. Surf Coat Technol 202:4387–4393
Oberste Berghaus J, Marple B, Moreau C (2006) Suspension plasma spraying of nanostructured
WC-12Co coatings. J Therm Spray Technol 15(4):676–681
Pateyron B, Elchinger MF, Delluc G, Fauchais P (1996) Sound velocity in different reacting thermal
plasma systems. Plasma Chem Plasma Process 16(1):39–57
Pershin L, Mitrasinovic A, Mostaghimi J (2013) Treatment of refractory powders by a novel, high
enthalpy dc plasma. J Phys D Appl Phys 46:224019
Pfender E (1989) Multiple arc plasma device with continuous gas jet. US patent 4,818,837
Pfender E, Fink J, Spores R (1991) Entrainment of cold gas into thermal plasma jets. Plasma Chem
Plasma Process 11(4):529–543
Rahmane M, Soucy G, Boulos M, Henne R (1998) Fluid dynamic study of direct current plasma jets
for plasma spraying applications. J Therm Spray Technol 7(3):349–356
Roumilhac P, Coudert J-F, Fauchais P (1990a) Designing parameters of spraying plasma torches. In:
Bernecki T (ed) Proceedings of NTSC-1990., Long Beach, CA. ASM International Materials
Park, pp 11–19
Roumilhac P, Coudert J-F, Fauchais P (1990b) Influence of the arc chamber design and the
surrounding atmosphere on the characteristics and temperature distribution of Ar-H2 and
Ar-He spraying plasma jets. In: Apelian D, Szekely J (eds) Plasma processing and synthesis
of materials, vol 190. MRS, Pittsburgh, pp 227–333
Roumilhac P, Fauchais P, Ducos M (1991) Optical and thermal diagnostics regarding the working
conditions of a plasma mini-spray torch. In: First Plasma Technik symposium, vol 1. Plasma
Technik, Wohlen, pp 121–131
Russ S, Strykowski PJ, Pfender E (1994a) Mixing in plasma and low-density jets. Exp Fluids
16:297–307
Russ S, Pfender E, Strykowski PJ (1994b) Unsteadiness and mixing in thermal plasma jets. Plasma
Chem Plasma Process 14(4):425–436
Schwenk A, Gruner H, Zimmermann X, Landes K, Nutsch G (2004) Improved nozzle design of de-
Laval-type nozzles of the atmospheric plasma spraying. In: Ohmori A (ed) Proceedings of
ITSC-2004, Osaka, Japan. ASM International, Materials Park, pp 600–605
54 M. I. Boulos et al.

Schwenk A, Grund T, Wielage B, Zierhut J, Dzulko M, Landes K (2007) Delta gun–an improved
multiple electrode plasma system. In: Marple BR, Hyland MM, Lau Y-C, Li C-J, Lima RS,
Montavon G (eds) Proceedings of ITSC-2007 Beijing. ASM International, Materials Park
Suzuki M, Shahien M, Tsutai Y (2014) Suspension plasma spray by twin cathode type plasma spray
gun, 6th int. In: Meillot E (ed) Workshop on suspension and solution thermal spraying, CEA Le
Ripault (37) France
Tucker RC Jr (2013) Thermal spray technology, vol 5A. ASM International, Materials Park
Vardelle A, Vardelle M, Fauchais P, Li K-I, Dussoubs B, Themelis NJ (2001) Controlling particle
injection in plasma spraying. J Therm Spray Technol 10:267–289
Yushchenko K, Borisov Y, Pereverzev Y, Vojnarovitch S, Darmochval V, Vovric V, Ramaekers P,
Raa G (1998) Microplasma spraying. In: Coddet C (ed) Thermal spray: meeting the challenges
of the 21st century, proceedings of 15th, ITSC-1998, vol 2. ASM International, Materials Park,
pp 1461–1467
Zhukov MF (ed) (1977) Electric arc plasmatrons. The USSR Academy of Science, Siberian
Chapter, Institute of Thermal Physics, Novosibirsk/USSR
Zierhut J, Haslbeck P, Landes KD, Barbezat G, Muller M, Schutz M (1998) TRIPLEX - an
innovative three-cathode plasma torch. In: Coddet C (ed) Proceedings of ITSC-1998, Nice,
France. ASM International, Materials Park, pp 1374–1380

You might also like