04.01.2013 Views

Thermal Spray Tips - Swinburne University of Technology

Thermal Spray Tips - Swinburne University of Technology

Thermal Spray Tips - Swinburne University of Technology

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

<strong>Thermal</strong> <strong>Spray</strong> <strong>Tips</strong><br />

by <strong>Thermal</strong> <strong>Spray</strong> Society (TSS)<br />

This document has been compiled from data and information acquired from ASM <strong>Thermal</strong> <strong>Spray</strong> Society<br />

(TSS) website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/ for easy reference.<br />

This compilation serves as a contribution from <strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> (SUT) <strong>Thermal</strong> <strong>Spray</strong><br />

Group (SwinTS).<br />

The tips have been categorized into several topics (thermal spray processes, feedstock materials, surface<br />

preparation, coating processing and operation, post coating operation, coating characterization and<br />

testing as well as applications) where individual articles can be accessed easily through the hyperlinked<br />

Table <strong>of</strong> Contents (TOC) or the overview <strong>of</strong> the thermal spray (TS) process. The overview highlights the<br />

different components <strong>of</strong> the TS process and related tips are attached to each segment. Users may return<br />

easily to the TOC or overview anytime by clicking the link located at bottom right <strong>of</strong> each page.<br />

Please note that a list <strong>of</strong> glossary for <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong> terminology can be accessed at the<br />

following website: http://www.asminternational.org/tss/glossary/tssgloss.htm<br />

Welcome to the wonderful world <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong>!<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

1


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

Table <strong>of</strong> Contents<br />

1. OVERVIEW OF THERMAL SPRAY PROCESS 5<br />

2. THERMAL SPRAY PROCESSES 6<br />

2.1. COMPARISON OF MAJOR COATING METHODS 6<br />

2.2. CLASSIFICATION OF THERMAL SPRAY PROCESSES 8<br />

2.3. CHARACTERISTICS OF THERMAL SPRAY 9<br />

2.4. THERMAL SPRAY PROCESSING CHARACTERISTICS 10<br />

2.5. THERMAL SPRAY PROCESS VARIATIONS 11<br />

2.6. EFFECT OF TORCH HARDWARE ON PARTICLE TEMPERATURE AND VELOCITY 12<br />

2.7. FLAME SPRAY PROCESS 13<br />

2.8. FLAME SPRAY GUNS 14<br />

2.9. FLAME SPRAY USES COMBUSTIBLE GAS AS HEAT SOURCE TO MELT COATING MATERIAL 16<br />

2.10. HYPERSONIC FLAME SPRAYING 18<br />

2.11. ARC AND FLAME SPRAYING PROCESSES 19<br />

2.12. ELECTRIC ARC WIRE SPRAYING 21<br />

2.13. OXYFUEL WIRE SPRAY 22<br />

2.14. REACTIVE PLASMA SPRAY FORMING 23<br />

2.15. REACTIVE PLASMA SPRAY 24<br />

2.16. TRANSFERRED PLASMA-ARC PROCESS 25<br />

2.17. HIGH-VELOCITY OXY FUEL SPRAY PROCESS 26<br />

2.18. DETONATION GUN THERMAL SPRAY PROCESS FOR CERAMICS 27<br />

2.19. SPRAY TABLES FOR THERMAL SPRAY PARAMETERS 28<br />

2.20. COLD-SPRAY PROCESS PARAMETERS 29<br />

3. FEEDSTOCK MATERIALS 30<br />

3.1. PRODUCTION OF THERMAL SPRAY POWDERS 30<br />

3.2. GAS-PHASE PRODUCTION METHODS FOR ULTRAFINE POWDERS 31<br />

3.3. ADVANTAGES AND DISADVANTAGES OF POWDER PRODUCTION PROCESSES 32<br />

3.4. WATER ATOMIZATION 33<br />

3.5. COMPOSITE MATERIALS 34<br />

3.6. POWDER TESTING AND CHARACTERIZATION: IMAGE ANALYSIS 36<br />

3.7. POWDER CHARACTERISTICS: MORPHOLOGY 37<br />

3.8. PARTICLE SIZE DISTRIBUTION PLOTS 38<br />

3.9. WIRE AND ROD FEEDERS 39<br />

3.10. FLUIDIZED-BED POWDER FEEDERS 40<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

2


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.11. GRAVITY-BASED POWDER FEEDING 41<br />

3.12. POWDER FEEDING SYSTEMS: ROTATING WHEEL DEVICES 42<br />

4. SURFACE PREPARATION 43<br />

4.1. DRY ABRASIVE GRIT BLASTING 43<br />

4.2. ALUMINA BLASTING ABRASIVES 44<br />

4.3. MACHINING AND MACROROUGHENING 45<br />

5. COATING PROCESSING AND OPERATION 46<br />

5.1. PART CONFIGURATION AND COATING LOCATION CONSIDERATIONS 46<br />

5.2. CONTACT AND SHADOW MASKING 47<br />

5.3. THERMAL SPRAY PATTERN 48<br />

5.4. RELATIONSHIP BETWEEN PARTICLE MELTING AND COATING STRUCTURE 49<br />

5.5. AIR-COOLING DEVICES TO MINIMIZE DEBRIS 51<br />

5.6. INFLUENCE OF TEMPERATURE CONTROL ON SUBSTRATE/COATING 52<br />

6. COATING STRUCTURES, PROPERTIES AND MATERIALS 53<br />

6.1. MICROSTRUCTURES OF THERMAL SPRAY COATINGS 53<br />

6.2. STRUCTURE OF THERMAL SPRAY COATINGS 54<br />

6.3. SOURCES OF COATING POROSITY 55<br />

6.4. INFLUENCE OF POROSITY ON COATING PROPERTIES 56<br />

6.5. COAT BONDING: MECHANICAL INTERLOCKING 57<br />

7. POST COATING OPERATION 58<br />

7.1. SEALING 58<br />

7.2. RECOMMENDED METALLOGRAPHIC PRACTICE: SECTIONING 59<br />

8. COATING CHARACTERIZATION AND TESTING 60<br />

8.1. MEASURING DEPOSITION EFFICIENCY 60<br />

8.2. TENSILE ADHESION TESTING OF THERMAL SPRAY COATINGS 62<br />

8.3. BOND PULL TESTING OF THERMAL SPRAY COATINGS 64<br />

8.4. ROLLING CONTACT FATIGUE FAILURE MODES IN THERMAL SPRAY COATINGS 65<br />

8.5. RESIDUAL STRESS DETERMINATION IN THERMAL SPRAY COATINGS 66<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

3


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8.6. THERMAL FATIGUE TESTING OF TBCS 67<br />

9. APPLICATIONS 68<br />

9.1. COATINGS FOR CORROSIVE ENVIRONMENT: CONVENTIONAL COAL-FIRED BOILERS 68<br />

9.2. ELEVATED-TEMPERATURE CORROSION APPLICATIONS 69<br />

9.3. OXIDATION PROTECTION 70<br />

9.4. THERMAL BARRIER COATINGS 71<br />

9.5. FUNCTIONALLY GRADIENT MATERIALS 72<br />

9.6. WEAR AND ABRASION RESISTANCE OF THERMAL SPRAY COATINGS 73<br />

9.7. METALLOGRAPHY OF NICRAL/BENTONITE ABRADABLE COATINGS 74<br />

9.8. ABRADABLE SEALS 75<br />

9.9. PROCESSING ELECTRONIC DEVICES: BERYLLIA 77<br />

9.10. ALUMINUM COATINGS AND ZINC COATINGS 78<br />

9.11. THERMAL SPRAY POLYMER COATINGS 79<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

4


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

Tip 2.1 Tip 2.11<br />

Tip 2.2 Tip 2.12<br />

Tip 2.3 Tip 2.13<br />

Tip 2.4 Tip 2.14<br />

Tip 2.5 Tip 2.15<br />

Tip 2.6 Tip 2.16<br />

Tip 2.7 Tip 2.17<br />

Tip 2.8 Tip 2.18<br />

Tip 2.9 Tip 2.19<br />

Tip 2.10 Tip 2.20<br />

Image adapted from J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, 2004, p 43<br />

5<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

1. Overview <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Process<br />

Tip 3.1 Tip 3.2 Tip 3.3 Tip 3.4<br />

Tip 3.5 Tip 3.6 Tip 3.7 Tip 3.8<br />

Tip 3.9 Tip 3.10 Tip 3.11 Tip 3.12<br />

Back to TOC<br />

Back to Overview<br />

Tip 4.1 Tip 4.2 Tip 4.3<br />

Tip 5.1 Tip 5.2<br />

Tip 5.3 Tip 5.4<br />

Tip 5.5 Tip 5.6<br />

Tip 6.1<br />

Tip 6.2<br />

Tip 6.3<br />

Tip 6.4<br />

Tip 6.5<br />

Tip 7.1<br />

Tip 7.2<br />

Tip 8.1 Tip 8.2<br />

Tip 8.3 Tip 8.4<br />

Tip 8.5 Tip 8.6


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2. <strong>Thermal</strong> <strong>Spray</strong> Processes<br />

2.1. Comparison <strong>of</strong> Major Coating Methods<br />

Characteristic Electro/electroless<br />

plating<br />

General Characteristics <strong>of</strong> Major Coating Methods<br />

<strong>Thermal</strong> spray Chemical vapor<br />

deposition<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Physical vapor<br />

deposition<br />

Equipment cost Low Low to moderate Moderate Moderate to high<br />

Operating cost Low Low to moderate Low to moderate Moderate to high<br />

Process<br />

environment<br />

Aqueous solution Atmospheric to<br />

s<strong>of</strong>t vacuum<br />

Atmospheric to<br />

medium vacuum<br />

Hard vacuum<br />

Coating geometry Omnidirectional Line <strong>of</strong> sight Omnidirectional Line <strong>of</strong> sight<br />

Coating thickness Moderate to thick,<br />

10 μm-mm<br />

Substrate<br />

temperature<br />

Adherence Moderate<br />

mechanical bond to<br />

very good chemical<br />

bond<br />

Surface finish Moderately coarse<br />

to glossy<br />

Thick, 50 μm-cm Thin to thick, 0.1<br />

μm-mm<br />

Low Low to moderate Moderate to high Low<br />

Good mechanical<br />

bond<br />

Coating materials Metals Powder/wire,<br />

polymers,<br />

metals/ceramics<br />

Very good<br />

chemical bond to<br />

excellent diffusion<br />

bond<br />

Very thin to<br />

moderate<br />

Moderate<br />

mechanical bond<br />

to good chemical<br />

bond<br />

Coarse to smooth Smooth to glossy Smooth to high<br />

gloss<br />

Metals, ceramics,<br />

polymers<br />

Metals, ceramics,<br />

polymers<br />

Major coating methods include electro/electroless plating (EP/ElsP), chemical vapor deposition (CVD),<br />

physical vapor deposition (PVD), and thermal spray (TS). Both the initial equipment costs and the<br />

operating costs for EP/ElsP coatings are relatively low. However, the by-products are considered highly<br />

toxic and are subject to increasingly stricter government regulation, thereby making storage, reclamation,<br />

and disposal major economic concerns. <strong>Thermal</strong> spray equipment, and therefore its cost, varies widely<br />

from simple combustion devices to computer-controlled low-pressure plasma spray systems.<br />

The operating costs are very much dependent on the cost <strong>of</strong> consumables such as powder, wire, or rod<br />

materials as well as the quantities and types <strong>of</strong> gases used. Equipment costs for CVD are moderate, with<br />

the understanding that neutralization <strong>of</strong> the output gases is an integral part <strong>of</strong> any system. The operating<br />

costs <strong>of</strong> CVD are dominated by precursor gas costs and the frequent need to clean the systems. The cost<br />

<strong>of</strong> PVD coating equipment is very high, due to the need to maintain high vacuums in chambers <strong>of</strong><br />

sufficient volume to make the process cost-effective. Operating costs are associated with the degree <strong>of</strong><br />

surface cleanliness necessary for coating adhesion, as well as the target costs.<br />

Back to TOC<br />

Back to Overview<br />

6


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

The EP/ElsP and CVD are omnidirectional coating processes, while TS and PVD are line-<strong>of</strong>-sight<br />

processes. Adherence for TS coatings, in general, is provided by mechanical bonding. Adherence <strong>of</strong><br />

EP/ElsP and PVD coatings is, in general, provided by mechanical and/or chemical bonding. Adherence<br />

for CVD coatings is provided by either chemical and/or diffusion bonding.<br />

With the exception <strong>of</strong> EP/ElsP, which is limited to metals and some alloys, the TS, CVD, and PVD coating<br />

processes can apply a variety <strong>of</strong> metals, ceramics, cermets, and polymers.<br />

Source: Surface Science, Thomas Bernecki, BIRL, Northwestern <strong>University</strong>, as published in Handbook <strong>of</strong><br />

<strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 35, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

7


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.2. Classification <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Processes<br />

<strong>Thermal</strong> spray processes and subsets<br />

Members <strong>of</strong> the thermal spray family <strong>of</strong> processes typically are grouped into three major categories:<br />

flame, plasma arc, and electric arc. (Kinetic energy-driven spray, or cold spray, is a recent addition to the<br />

family.) Each <strong>of</strong> these processes encompasses many more subsets, and each has its own characteristic<br />

range <strong>of</strong> temperature, enthalpy, and velocity. These attributes, in turn, develop coating characteristics that<br />

are unique to each process, which in simplest terms include coating bond strength, porosity, inclusions<br />

(usually oxides), and hardness. Selection <strong>of</strong> the appropriate thermal spray method typically is determined<br />

by:<br />

• Desired coating material<br />

• Coating performance requirements<br />

• Economics<br />

• Part size and portability<br />

Source: J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, 2004, p 44<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

8


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.3. Characteristics <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong><br />

Schematic <strong>of</strong> a typical thermal spray powder process<br />

<strong>Thermal</strong> spray is a generic term for a group <strong>of</strong> coating processes used to apply metallic or nonmetallic<br />

coatings. Three major categories <strong>of</strong> the processes are flame spray, electric arc spray, and plasma arc<br />

spray, with subsets under each category. (Cold spray is a recent addition that uses modest preheating,<br />

but is largely a kinetic energy process.) The energy sources heat the coating material (powder, wire, and<br />

rod form) to a molten or semimolten state. Heated particles are propelled toward a prepared surface via<br />

process gases or atomization jets, forming a bond with the surface upon impact. Subsequent particles<br />

cause thickness buildup and form a lamellar structure. Advantages <strong>of</strong> thermal spray processes include<br />

the ability to use a wide variety <strong>of</strong> coating materials, the ability <strong>of</strong> most thermal spray processes to apply<br />

coatings without significant heat input, and the ability, in most cases, to strip <strong>of</strong>f and recoat worn or<br />

damaged coatings without changing part properties or dimensions.<br />

Source: J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, p 4, 2004<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview<br />

9


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.4. <strong>Thermal</strong> <strong>Spray</strong> Processing Characteristics<br />

Typical thermal spray process parameters and variables<br />

From a simplistic point <strong>of</strong> view, the powder spray process can be viewed as high-speed heat treating<br />

(dealing with time, temperature, and mass) where the objective is to bring the powder mass to the<br />

required temperature in a given time period. The time a particle spends in the process jet is called dwell<br />

time, which is governed by gas velocity and powder particle characteristics. Gas velocity, in turn, is<br />

determined by the total gas flow through the nozzle, gas characteristics, and energy acting on or resulting<br />

from the process. Particle velocity is a function <strong>of</strong> jet velocity coupled with particle characteristics; i.e.,<br />

particle size, morphology, and mass. Particle temperature is a function <strong>of</strong> enthalpy, velocity, trajectory,<br />

and its own physical and thermal properties. Gas temperatures in the spray stream vary greatly with the<br />

process.<br />

Source: J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, 2004, p 45<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

10<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.5. <strong>Thermal</strong> <strong>Spray</strong> Process Variations<br />

Heat energy input and particle velocity for common thermal spray processes<br />

Process Input heat energy to<br />

particle<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Output particle velocity (a)<br />

High Low Highest High Medium high Medium Low<br />

Combustion wire X X<br />

Combustion powder X X<br />

Standard plasma X X<br />

High-velocity plasma X X<br />

Vacuum plasma X X<br />

Standard wire arc X X<br />

Vacuum arc X X<br />

High-velocity oxyfuel X X<br />

Detonation gun X X<br />

(a) Particle speed ranges from a high <strong>of</strong> approximately 1000 m/s (3000 ft/s) to a low <strong>of</strong> 25 m/s (80 ft/s). Further variations within<br />

each process depends on the particle size, material type and gas velocity<br />

Source: ASM Handbook: Metallography and Microstructures, Vol.9, p.1038, 2004<br />

The controllable parameters for each <strong>of</strong> the major thermal spray types are different, but all share some<br />

common fundamentals. Converting feedstock material (whether powder, wire, or rod) into a coating<br />

requires heat energy for melting and a gas flow for atomization and/or particle acceleration to propel the<br />

particles toward the substrate. Input heat energy and gas flow, in turn, directly influence process variables<br />

including flame/plasma/jet temperature and velocity; particle temperature, speed, and trajectory; and<br />

deposit temperature. Each <strong>of</strong> these variables affects spray-pattern formation and particle heating<br />

differently.<br />

In general, combustion-flame, combustion-wire, and twin-wire arc spray coatings are generated from high<br />

heat input plus low particle velocity, producing coatings having high oxide content, a large number <strong>of</strong><br />

rounded particles, and relatively high porosity. Due to higher particle velocity, superheated particles in<br />

plasma spray coatings are flattened more than in combustion spray coatings, producing denser coatings<br />

with finer porosity. Very high-velocity systems such as detonation gun and HVOF, combined with lower<br />

heat inputs, typically produce very dense coatings not achievable by the other processes.<br />

11<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.6. Effect <strong>of</strong> Torch Hardware on Particle Temperature and Velocity<br />

Gun Cooling, Pinch, and Flame Cooling air caps air caps were fitted onto nozzles, and the particle<br />

temperature (Tp) and particle velocity (Vp) were measured. The gun cooling air cap produced the highest<br />

Tp and lowest Vp because it minimized the interaction between the cooling air and the particulate loaded<br />

flame. The pinch air cap strongly directed the cooling air into the flame, and produced the highest Vp and<br />

lowest Tp values. The flame cooling air cap also directed air into the flame, but not as strongly as the<br />

pinch air cap, and as a result, the flame cooling air cap produced Tp and Vp values between those <strong>of</strong> the<br />

gun cooling and pinch air caps.<br />

Source: Journal <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, Volume 19(4) June 2010, p. 824-827.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

12<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.7. Flame <strong>Spray</strong> Process<br />

Powder flame spray system<br />

Conventional flame spray was the first thermal spray process developed (~1910) and is still in common<br />

use. Modern torches have changed little since the 1950s. Flame spray uses the chemical energy <strong>of</strong><br />

combusting fuel gases to generate heat. Oxyacetylene torches are the most common, using acetylene as<br />

the main fuel in combustion with oxygen to generate the highest combustion temperatures. Powder, wire,<br />

or rod is introduced axially through the rear <strong>of</strong> the nozzle into the flame at the nozzle exit. The feedstock<br />

material is melted and the particles/droplets accelerated toward the surface by the expanding gas flow<br />

and air jets. An advantage <strong>of</strong> wire and rod over powder is that the degree <strong>of</strong> melting is significantly higher,<br />

producing denser coatings. In addition, the atomizing air produces finer droplets, which in turn produce<br />

finer, smoother coatings. In flame spray processes, fuel/oxygen ratio and total gas flow rates are adjusted<br />

to produce the desired thermal output. Optional air jets, downstream <strong>of</strong> the combustion zone, can further<br />

adjust the thermal pr<strong>of</strong>ile <strong>of</strong> the flame. Flame spray is capable <strong>of</strong> depositing a wide range <strong>of</strong> materials,<br />

ranging from polymers to ceramics and refractory metals.<br />

Source: <strong>Thermal</strong> <strong>Spray</strong> Processes, revised by Daryl E. Crawmer, <strong>Thermal</strong> <strong>Spray</strong> Technologies Inc., as<br />

published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 55.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

13<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.8. Flame <strong>Spray</strong> Guns<br />

Combustible gas serves as the heat source to melt the coating material in flame spray guns. Materials are<br />

in the form <strong>of</strong> rod, wire, or powder, most flame spray guns can be adapted to several combinations <strong>of</strong><br />

gases to balance operating cost and coating properties. Acetylene, propane, methyl-acetylenepropadiene<br />

(MAPP) gas, and hydrogen, along with oxygen, are typical flame spray gases.<br />

In general, changing the nozzle and/or air cap is all that is required to adapt the gun to different alloys,<br />

wire sizes, or gases. The diagrams depict powder and wire flame spray guns. For all practical purposes,<br />

the rod and wire guns are similar.<br />

Cross sections <strong>of</strong> typical flame spray guns. (a) Wire or rod. (b) Powder.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

14<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

Flame temperatures and characteristics depend on the oxygen-to-fuel gas ratio and pressure. The flame<br />

spray process is characterized by low capital investment, high deposition rates and efficiencies, and<br />

relative ease <strong>of</strong> operation and cost <strong>of</strong> equipment maintenance.<br />

In general, as-deposited (or cold spray) flame-sprayed coatings exhibit lower bond strengths, higher<br />

porosity, a narrower working temperature range, and higher heat transmittal to the substrate than most<br />

other thermal spray processes. The flame spray process is frequently selected for the reclamation <strong>of</strong> worn<br />

or out-<strong>of</strong>-tolerance parts, frequently with nickel-base alloys. Bronze alloys may be used for some bearings<br />

and seal areas. Blends <strong>of</strong> tungsten carbide and nickel-base alloys work well for wear resistance. Zinc is<br />

commonly sprayed for corrosion resistance on bridges and other structures.<br />

Source: R.C. Tucker, Jr., <strong>Thermal</strong> <strong>Spray</strong> Coatings, Surface Engineering, Vol 5, ASM Handbook, ASM<br />

International, 1994, p 498.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

15<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.9. Flame <strong>Spray</strong> Uses Combustible Gas as Heat Source to Melt Coating Material<br />

Flame spray uses combustible gas as a heat source to melt the coating material. Flame spray guns are<br />

available to spray materials in either rod, wire, or powder form. Most flame spray guns can be adapted to<br />

use several combinations <strong>of</strong> gases to balance operating cost and coating properties. Acetylene, propane,<br />

methyl-acetylene-propadiene (MAPP) gas, and hydrogen, along with oxygen, are commonly used flame<br />

spray gases. In general, changing the nozzle and/or air cap is all that is required to adapt the gun to<br />

different alloys, wire sizes, or gases. Figures 3(a) and 3(b) depict powder and wire flame spray guns. For<br />

all practical purposes, the rod and wire guns are similar.<br />

Cross sections <strong>of</strong> typical flame spray guns. (a) Wire or rod. (b) Powder.<br />

Source: R.C. Tucker, Jr., <strong>Thermal</strong> <strong>Spray</strong> Coatings, Surface Engineering, Vol 5, ASM Handbook, ASM<br />

International, 1994, p 498.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

16<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

Flame temperatures and characteristics depend on the oxygen-to-fuel gas ratio and pressure. The<br />

approximate temperatures for stoichiometric combustion at 1 atm for some oxyfuel combinations are<br />

shown in Table 1.<br />

The flame spray process is characterized by low capital investment, high deposition rates and<br />

efficiencies, and relative ease <strong>of</strong> operation and cost <strong>of</strong> equipment maintenance. In general, as-deposited<br />

(or cold spray) flame-sprayed coatings exhibit lower bond strengths, higher porosity, a narrower working<br />

temperature range, and higher heat transmittal to the substrate than most other thermal spray processes.<br />

The flame spray process is widely used for the reclamation <strong>of</strong> worn or out-<strong>of</strong>-tolerance parts, frequently<br />

using nickel-base alloys. Bronze alloys may be used for some bearings and seal areas. Blends <strong>of</strong><br />

tungsten carbide and nickel-base alloys may be used for wear resistance. Zinc is commonly applied for<br />

corrosion resistance on bridges and other structures.<br />

Table 1: Maximum temperature <strong>of</strong> heat sources<br />

Heat source<br />

Approximate temperature<br />

(stoichiometric combustion)<br />

Propane-oxygen 2526 °C (4579 °F)<br />

Natural gas-oxygen 2538 °C (4600 °F)<br />

Hydrogen-oxygen 2660 °C (4820 °F)<br />

Propylene-oxygen 2843 °C (5240 °F)<br />

Methylacetylene/propadiene- 2927 °C (5301 °F)<br />

oxygen<br />

Acetylene-oxygen 3087 °C (5589 °F)<br />

Plasma arc 2200 to 28,000 °C<br />

(4000 to 50,000 °F)<br />

Source: Adapted Publication 1G191, National Association <strong>of</strong> Corrosion Engineers<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

17<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.10. Hypersonic Flame <strong>Spray</strong>ing<br />

Cross-sectional view <strong>of</strong> a flame powder spraying system showing powder feed material being transported<br />

by the carrier gas and then melted by the oxyfuel mixture.<br />

In recent years, the field <strong>of</strong> thermal spraying has seen the introduction <strong>of</strong> hypersonic combustion flame<br />

spray (HCFS) guns. In the Jet-Kote method (the first gun <strong>of</strong> this type was commercially available in 1983),<br />

a high-velocity combustion process (chemical energy) accelerates and melts the particulates via<br />

convective heat transfer from the hot flame.<br />

The system consists basically <strong>of</strong> three parts: a main console, a pressurized powder feeder, and a<br />

portable torch. An internal combustion flame is produced in a water-cooled combustion chamber with a<br />

continuous flow <strong>of</strong> oxygen and fuel gases.<br />

This flame makes a right-angle bend and passes through four vortices that focus the hot gases into a<br />

narrow beam. This beam is accelerated through an extended-length water-cooled nozzle. The<br />

combustion process creates a high back pressure, thus requiring a pressurized powder feeder to propel<br />

particulates axially from the rear <strong>of</strong> the nozzle into the stream <strong>of</strong> hot gases.<br />

Source: Engineered Materials Handbook -> Densification and Sintering <strong>of</strong> Ceramics -> Nontraditional<br />

Densification Processes<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

18<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.11. Arc and Flame <strong>Spray</strong>ing Processes<br />

Electric arc spraying and flame spraying are the most suitable methods for applying coatings <strong>of</strong> zinc,<br />

aluminum, and their alloys to steel structures. These methods are simple, they can be used on-site, and<br />

they <strong>of</strong>fer relatively high coverage rates compared to other thermal spray processes. Wire consumables<br />

are also more economical and easier to handle than powders.<br />

Electric arc spraying involves feeding two electrically conducting metal wires toward<br />

each other. An electric arc is produced at the wire tips, melting the metal. A highpressure<br />

air line atomizes and then sprays the fine droplets onto the steel surface.<br />

In flame spraying, a metal wire is melted in a gas flame and then air sprayed onto the steel surface.<br />

Arc and flame spray coatings usually contain numerous pores, some <strong>of</strong> which are closed and the rest<br />

connected from the coating surface to the substrate. Natural sealing can be achieved by oxidation <strong>of</strong> the<br />

metallic coating under normal environmental exposure conditions, when the resulting oxides, hydroxides,<br />

and/or basic salts are not soluble in this environment.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

19<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

It is <strong>of</strong>ten beneficial to apply an organic sealant to close these pores. However, penetration <strong>of</strong> sealant into<br />

the coating becomes more difficult with time after spraying. This is due to the adsorption <strong>of</strong> moisture on<br />

the pore surface, which can lead to the formation <strong>of</strong> corrosion product within the pores. As a<br />

consequence, it is <strong>of</strong>ten recommended that any post-treatment be completed within the same day as the<br />

coating application.<br />

Source: <strong>Thermal</strong> <strong>Spray</strong> Application Methods for TSA and TSZ Coatings, <strong>Thermal</strong> <strong>Spray</strong> Coatings for<br />

Corrosion Protection in Atmospheric and Aqueous Environments, Vol 13B, ASM Handbook, ASM<br />

International<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

20<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.12. Electric arc wire spraying<br />

Electric arc wire spraying applies coatings <strong>of</strong> selected metals in wire form. Push-pull motors feed two<br />

electrically charged wires through the arc gun to contact tips at the gun head. An arc is created that melts<br />

the wires at temperatures above 5500 °C (10,000 °F). Compressed air atomizes the molten metal and<br />

projects it onto a prepared surface. The diagram shows a schematic <strong>of</strong> the EAW spray process.<br />

Schematic <strong>of</strong> the electric arc wire spray process.<br />

The EAW process is excellent for applications that require heavy coating buildup or have large surfaces<br />

to be sprayed. The arc system can produce a spray pattern ranging from 50 to 300 mm (2 to 12 in.) and<br />

can spray at high speeds. It has built-in flexibility, allowing coating characteristics such as hardness or<br />

surface texture to be tailored to specific applications.<br />

The EAW method is characterized by strong coating adhesion because <strong>of</strong> the high particle temperatures<br />

produced. Because the process uses only electricity and compressed air, it allows equipment to be<br />

moved relatively easily from one installation to another, and it eliminates the need to stock oxygen and<br />

fuel gas supplies. Typical spray materials include austenitic and martensitic stainless steels, nickel<br />

aluminide, nickel-chromium alloys, bronze, Monel, babbitt, aluminum, zinc, and molybdenum.<br />

Source: ASM Handbooks Online, Volume 6, Welding, Brazing, and Soldering -> Hardfacing, Weld<br />

Cladding, and Dissimilar Metal Joining -> <strong>Thermal</strong> <strong>Spray</strong>ing<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

21<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.13. Oxyfuel Wire <strong>Spray</strong><br />

Schematic <strong>of</strong> the oxyfuel wire spray process<br />

The oxyfuel wire spray process (also called wire flame spraying or the combustion wire process) is the<br />

oldest <strong>of</strong> the thermal spray coating methods and among the lowest in capital investment. The process<br />

utilizes an oxygen-fuel gas flame as a heating source and coating material in wire form. Any material in<br />

the form <strong>of</strong> wire and capable <strong>of</strong> being melted below 2480°C (4500°F) can be flame-sprayed.<br />

During operation, the wire is drawn into the flame by drive rolls that are powered by an adjustable air<br />

turbine or electric motor. The tip <strong>of</strong> the wire is melted as it enters the flame, atomized into particles by a<br />

surrounding jet <strong>of</strong> compressed air, and propelled to the workpiece. The diagram shows a schematic <strong>of</strong> the<br />

OFW spray process.<br />

<strong>Spray</strong> rates for this process range from 0.5 to 10 kg/h (1 to 20 lb/h) and are dictated by the melting point<br />

<strong>of</strong> the material and the choice <strong>of</strong> fuel gas. Common fuel gases are acetylene, MAPP gas, propane,<br />

propylene, and natural gas, each combined with oxygen.<br />

Source: J.R. Davis, Hardfacing, Weld Cladding, and Dissimilar Metal Joining, ASM Handbook, Vol 6,<br />

Welding, Brazing, and Soldering, ASM International, 1993, p 805.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

22<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.14. Reactive Plasma <strong>Spray</strong> Forming<br />

Reactive plasma spray process for the synthesis <strong>of</strong> composite materials<br />

Controlled-atmosphere plasma spraying, adapted for reactive plasma spraying, has recently been<br />

developed to combine controlled dissociation and reactions in thermal plasma jets, for the in situ forming<br />

<strong>of</strong> new materials or to produce new phases in sprayed deposits. Reactive plasma spray forming is<br />

consequently emerging as a viable method for producing a wide range <strong>of</strong> advanced materials.<br />

The process, a logical evolution from conventional plasma spraying, allows reactive precursors to be<br />

injected into the particulate and/or hot gas streams. These reactive precursors may be liquids, gases, or<br />

mixtures <strong>of</strong> solid reactants. On contact with the high-temperature plasma jet, they decompose or<br />

dissociate to form highly reactive and ionic species that can then react with other heated materials within<br />

the plasma jet to form new compounds.<br />

The diagram illustrates the reactive plasma spray concept. Chemical reactions rely on plasma-induced<br />

dissociation <strong>of</strong> the injected precursors. These species can react with other elements to produce carbides<br />

such as TiC and WC, or under certain conditions, diamond or diamondlike carbon (DLC) films. The<br />

primary requirements are that the precursors dissociate into reactive species and that the reaction times<br />

and temperatures are sufficiently long for the desired products or phases to form.<br />

Source: R. Knight and R.W. Smith, <strong>Thermal</strong> <strong>Spray</strong> Forming <strong>of</strong> Materials, ASM Handbook, Vol 7, Powder<br />

Metal Technologies and Applications, ASM International, p 415.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

23<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.15. Reactive Plasma <strong>Spray</strong><br />

Reactive plasma spray applications include the synthesis <strong>of</strong> composite materials, shaped brittle<br />

intermetallic alloys, reinforced or toughened ceramics, and tribological coatings with in situ-formed hard or<br />

lubricating phases. Reactive plasma spray forming enables a wide range <strong>of</strong> materials to be produced,<br />

such as aluminium with AlN, Al2O3, or SiC; NiCrTi alloys with TiC or TiN; intermetallics such as TiAl, Ti3Al,<br />

MoSi2, and other ceramics with oxides, nitrides, borides, and/or carbides. All <strong>of</strong> these have been<br />

produced in situ in reactive thermal plasma jets.<br />

The process allows "reactive" precursors to be injected into the particulate and/or hot gas streams. These<br />

reactive precursors may be liquids, gases, or mixtures <strong>of</strong> solid reactants which, on contact with the hightemperature<br />

plasma jet, decompose or dissociate to form highly reactive and ionic species that can then<br />

react with other heated materials within the plasma jet to form new compounds.<br />

Reactive plasma spray process for the synthesis <strong>of</strong> composite materials<br />

Chemical reactions rely on plasma-induced dissociation <strong>of</strong> the injected precursors, for example gaseous<br />

methane (CH4), which decomposes into elemental or ionic species such as CH3, C4- or even atomic<br />

carbon. These species can react with other elements to produce carbides such as TiC and WC; or, under<br />

certain conditions, diamond or diamond like carbon (DLC) films. The diagram shows the plasma heating<br />

<strong>of</strong> gases, using either a non-transferred electric arc or an inductively coupled plasma (ICP) radio<br />

frequency discharge; injection <strong>of</strong> reactive precursors into the plasma jet; and injection <strong>of</strong> powders in a<br />

carrier gas stream.<br />

Source: R. Knight and R.W. Smith, <strong>Thermal</strong> <strong>Spray</strong> Forming <strong>of</strong> Materials, ASM Handbook, vol 7, Powder<br />

Metal Technologies and Applications, ASM International, 1998, p 415<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

24<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.16. Transferred plasma-arc process<br />

The transferred plasma-arc process adds to plasma spray the capability <strong>of</strong> substrate surface<br />

heating and melting. The diagram shows a schematic representation <strong>of</strong> the process.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

A secondary arc current that controls surface<br />

melting and depth <strong>of</strong> penetration is established<br />

through the plasma and substrate. Several<br />

advantages result from this direct heating:<br />

metallurgical bonding, high density, high<br />

deposition rates, and high thickness per pass.<br />

Coating thickness <strong>of</strong> 0.50 to 6.35 mm (0.020 to<br />

0.250 in.) and widths up to 32 mm (1.25 in.) can<br />

be made in a single pass at powder feed rates <strong>of</strong><br />

9 kg/h (20 lb/h).<br />

In addition, less electrical power is required than<br />

with nontransferred arc processes. For example,<br />

for an 88% tungsten carbide, 12% cobalt material,<br />

plasma spray deposition 0.30 mm (0.012 in.) thick<br />

and 9.50 mm (0.375 in.) in width might require 24<br />

passes at 40 to 60 kW to achieve maximum<br />

coating properties. However, the transferred<br />

plasma-arc process needs only one pass at approximately 2.5 kW for the same thickness.<br />

The method <strong>of</strong> heating and heat transfer in the transferred plasma-arc process eliminates many <strong>of</strong> the<br />

problems related to powders with wide particle size distributions or large particle sizes. Larger-particles,<br />

for example in the 50-mesh range, tend to be less expensive than closely classified 325-mesh powders.<br />

Some limitations <strong>of</strong> the process should be considered for any potential application. Because substrate<br />

heating is a part <strong>of</strong> the process, some alteration <strong>of</strong> its microstructure is inevitable. Applications are also<br />

limited to substrates that are electrically conductive and can withstand some melting. The transferred<br />

plasma-arc process is used in hardfacing applications such as valve seats, plowshares, oilfield<br />

components, and mining machinery.<br />

Source: ASM Handbooks Online, Volume 5, Surface Engineering -> <strong>Thermal</strong> <strong>Spray</strong> Coatings -><br />

Processes<br />

25<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.17. High-Velocity Oxy Fuel <strong>Spray</strong> Process<br />

Schematic <strong>of</strong> HVOF spray process. Photo courtesy <strong>of</strong> Sulzer-Metco Ltd.<br />

The HVOF (high velocity oxy-fuel) process efficiently uses high kinetic energy and controlled thermal<br />

output to produce dense, low-porosity coatings that exhibit high bond strengths (some <strong>of</strong> which exceed 83<br />

MPa, or 12,000 psi), low oxides, and extremely fine as-sprayed finishes. The coatings have low residual<br />

internal stresses, and therefore, can be sprayed to a thickness not normally associated with dense,<br />

thermal spray coatings. This process uses an oxygen-fuel mixture. Depending on user requirements,<br />

propylene, propane, hydrogen, or natural gas can be used as the fuel in gas-fueled spray systems and<br />

kerosene as the fuel in liquid-fueled systems. The coating material in powder form is fed axially through<br />

the gun, generally using nitrogen as a carrier gas. The fuel is thoroughly mixed with oxygen within the gun<br />

and the mixture is then ejected from a nozzle and ignited outside the gun. The ignited gases surround<br />

and uniformly heat the powder spray material as it exits the gun and is propelled to the workpiece<br />

surface. As a result <strong>of</strong> the high kinetic energy transferred to the particles through the HVOF process, the<br />

coating material generally does not need to be fully melted. Instead, the powder particles are in a molten<br />

state and flatten plastically as they impact the workpiece surface. The resulting coatings have very<br />

predictable chemistries that are homogeneous and have a fine granular structure.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

26<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.18. Detonation Gun <strong>Thermal</strong> <strong>Spray</strong> Process for Ceramics<br />

The detonation gun, or D-gun, consists <strong>of</strong> a water-cooled barrel about 1 m (3 ft) in length with an inside<br />

diameter <strong>of</strong> approximately 25 mm (1 inch) and associated gas and powder metering equipment.<br />

Detonation gun coatings have some <strong>of</strong> the highest bond strengths and lowest porosities <strong>of</strong> the thermal<br />

spray coatings.<br />

Detonation gun process. Courtesy <strong>of</strong> Praxair Surface Technologies Inc.<br />

Initially, a mixture <strong>of</strong> oxygen and acetylene is fed into the barrel along with a charge <strong>of</strong> powder. The gas is<br />

then ignited. It accelerates the powder to a velocity <strong>of</strong> 800 m/s ( 2600 ft/s). The maximum free burning<br />

temperature <strong>of</strong> oxygen-acetylene mixtures occurs with 45 vol% C2H2 and is 3140 C ( 5680°F). It is<br />

estimated that under detonation conditions, the temperature reaches 4000°C (7230°F). Even higher<br />

maximum free burning temperatures are reached in the super D-gun. The great success <strong>of</strong> this process<br />

has been in decomposition-sensitive materials, such as cermets that require the metal matrix to be<br />

slightly melted to avoid carbide dilution.<br />

The extremely high velocities and consequent kinetic energy <strong>of</strong> the particles in the Super D-Gun process<br />

allow most <strong>of</strong> the coatings to be deposited with residual compressive stress, rather than tensile stress as<br />

is typical <strong>of</strong> most <strong>of</strong> the other thermal spray coatings.<br />

Source: Engineered Materials Handbook -> Densification and Sintering <strong>of</strong> Ceramics -> Nontraditional<br />

Densification Processes<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

27<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.19. <strong>Spray</strong> Tables for <strong>Thermal</strong> <strong>Spray</strong> Parameters<br />

Checklist for a plasma spray process<br />

Plasma processing equipment<br />

• Torch type (anode, cathode, injector ring), mm<br />

• Volts and amps = powder, kW<br />

• Primary gas and flow rate, slpm<br />

• Secondary gas and flow rate, slpm<br />

• Feed gas and flow rate, slpm<br />

• Stand<strong>of</strong>f distance, mm<br />

• Powder injection, mm<br />

• Powder feed rate, g/min<br />

Hardware<br />

• Traverse speed <strong>of</strong> torch, m/s<br />

• Substrate cooling, m 3 /s<br />

• <strong>Spray</strong> footprint, mm 2<br />

Checklists, or spray tables, can be defined as a list <strong>of</strong> primary thermal spray parameters that need to be<br />

specified and controlled during a specific coating operation. All <strong>of</strong> these parameters influence process<br />

economics and coating quality. In many instances, manufacturers indicate the physical properties, such<br />

as surface roughness, microstructure porosity, and adhesion, that can be expected if coatings are<br />

produced using their recommended spray parameters. These values should be used only as a rough<br />

guideline for coating properties. The physical properties <strong>of</strong> the substrate also are very important, being<br />

essential to the production <strong>of</strong> an appropriate surface pr<strong>of</strong>ile.<br />

Even with similar thermal spray processes, the spray parameters are likely to change from process to<br />

process and between different equipment. Powders that are presumably identical may exhibit different<br />

spray parameters for each set <strong>of</strong> equipment.<br />

Source: C.C. Berndt, Material Categories for <strong>Thermal</strong> <strong>Spray</strong> Coatings, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong><br />

<strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 143, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

28<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

2.20. Cold-<strong>Spray</strong> Process Parameters<br />

Typical range <strong>of</strong> gas-jet parameters for cold spray coating<br />

Stagnation jet pressure, MPa (psi) 1 – 3 (145 – 435)<br />

Stagnation jet temperature, °C (°F) 0 – 700 (32 – 1290)<br />

Gas flow rate, m 3 /min (ft 3 /min) 1 – 2 (35 – 70)<br />

Powder feed rate, kg/h (lb/h) 2 – 8 (4 – 18)<br />

<strong>Spray</strong> distance, mm (in.) 10 – 50 (0.4 – 2)<br />

Power consumption, kW (for heating gas) 5 – 25<br />

Particle size, μm 1 - 50<br />

Process gases include nitrogen, helium, air, and mixtures <strong>of</strong> these gases. Nitrogen is a favored process<br />

gas because it can be used to spray some materials without promoting oxidation and because is much<br />

less expensive than helium. Helium is capable <strong>of</strong> providing the highest gas velocities, and therefore can<br />

be used in depositing the widest possible range <strong>of</strong> materials. Helium may also be diluted with nitrogen to<br />

improve the economics <strong>of</strong> the process, while providing particle velocities in excess <strong>of</strong> that achieved with<br />

nitrogen alone.<br />

Source: A. Papyrin, Cold <strong>Spray</strong> <strong>Technology</strong>, Adv. Mater. Process., Sept 2001, p 49-51; from Cold <strong>Spray</strong><br />

Process, as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p<br />

78, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

29<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3. Feedstock Materials<br />

3.1. Production <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Powders<br />

Particle size ranges for common industrial processes including flame spray and plasma<br />

spray thermal spray powders<br />

<strong>Thermal</strong> spray powders have evolved principally from powder metallurgy procedures. The most recent<br />

developments have arisen from ceramics and composites processing techniques. The production <strong>of</strong><br />

polymeric powders that are suitable for thermal spray is an emerging engineering field. The powder<br />

production technique has a marked influence on the nature <strong>of</strong> the powder that is produced. The generic<br />

terms that describe the manufacture <strong>of</strong> powders by mechanical processes are comminution and attrition<br />

and denote the breaking up and size reduction <strong>of</strong> solid materials. The normal range <strong>of</strong> particle sizes<br />

required for thermal spray is between 5 µm to about 120 µm.<br />

Source: A. Bose, Advances in Particulate Materials, Butterworth-Heinemann, 1995, in Material<br />

Production Processes, Christopher C. Berndt, Stoney Brook <strong>University</strong> (at the time <strong>of</strong> publishing), as<br />

published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 147, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

30<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.2. Gas-phase production methods for ultrafine powders<br />

Several different gas-phase methods are available for the production <strong>of</strong> ultrafine powders with submicron<br />

particle size. Four such methods are inert gas condensation, flame reduction, plasma reduction, and<br />

chemical vapor reaction.<br />

With inert gas condensation (IGC), high-quality powders can be produced with low chemical impurities<br />

from precursor materials, and low amounts <strong>of</strong> oxides or nitrides from the production process.<br />

Flame reduction and plasma reduction are based on the decomposition and reduction <strong>of</strong> metal salts in a<br />

gas flame or plasma.<br />

In chemical vapor reactions (CVR), metal chlorides and hydrogen are chemically reacted in a hot-wall<br />

reactor.<br />

Comparison <strong>of</strong> the most frequently used gas phase methods for production <strong>of</strong> metal<br />

nanopowders<br />

Method Advantage Disadvantage Typical capacity<br />

Flame reaction Large output Broad distribution <strong>of</strong> >1 tonne/day<br />

particle<br />

impurities<br />

size, ionic<br />

Plasma reaction Large output Broad distribution <strong>of</strong> >1 tonne/day<br />

particle<br />

impurities<br />

size, ionic<br />

Chemical vapour<br />

reaction<br />

Narrow distribution <strong>of</strong><br />

particle size<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Ionic impurities 200 kg/day<br />

Inert gas condensation No impurities Small output Ultrafine<br />

and Nanophase Powders -> Metallic Nanopowders -> Production Methods<br />

31<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.3. Advantages and Disadvantages <strong>of</strong> Powder Production Processes<br />

Advantages and disadvantages <strong>of</strong> selected powder preparation techniques<br />

Powder preparation technique Advantages Disadvantages<br />

Conventional method<br />

Conventional comminution Inexpensive<br />

Unaggregated powders<br />

Wide applicability<br />

Chemical preparation<br />

Solution methods<br />

High purity<br />

Solvent vaporization<br />

Small particles<br />

Simple evaporation<br />

Compositional control<br />

<strong>Spray</strong> drying<br />

Chemical homogeneity<br />

<strong>Spray</strong> roasting<br />

Fluid bed drying<br />

Emulsion drying<br />

Sol-gel or glass drying<br />

Freeze drying<br />

Precipitation or coprecipitation<br />

Vapour-phase methods<br />

High purity<br />

Vaporization-condensation Very small particles<br />

Vapor decomposition<br />

Vapor-vapor reaction<br />

Salt decomposition Used for solution techniques<br />

Simple apparatus<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Limited purity<br />

Limited homogeneity<br />

Large particles<br />

Expensive<br />

Aggregation <strong>of</strong> particles<br />

Poor for nonoxides<br />

Expensive<br />

Difficult for multi-component<br />

materials<br />

Aggregation <strong>of</strong> particles<br />

<strong>Thermal</strong> <strong>Spray</strong> powders have evolved principally from powder metallurgy procedures. The most recent<br />

developments have arisen from ceramics and composites processing techniques. The production <strong>of</strong><br />

polymeric powders suitable for thermal spray is an emerging field. The powder-production technique has<br />

a marked influence on the nature <strong>of</strong> the powder that is produced. Feedstock production is done using<br />

mechanical and chemical routes. The generic terms that describe the manufacture <strong>of</strong> powders by<br />

mechanical processes are comminution and attrition and denote the breaking up and size reduction <strong>of</strong><br />

solid materials. Atomization can be classified as a mechanical powder production method, but it differs<br />

from conventional comminution in that it involves the dispersion <strong>of</strong> melts rather than the dispersion <strong>of</strong><br />

solids. The powder preparation subcategory that is used extensively for the production <strong>of</strong> ceramic<br />

feedstock lies within the solution methods <strong>of</strong> chemical preparation. The normal range <strong>of</strong> particle sizes<br />

required for thermal spray is 5 to about 120 µm.<br />

Source: Material Production Processes, Christopher C. Berndt, Stony Brook <strong>University</strong> (at the time <strong>of</strong> this<br />

publication), as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM<br />

International, p 147, 2005.<br />

32<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.4. Water Atomization<br />

In general, water atomization is less expensive than the other methods <strong>of</strong> atomization because <strong>of</strong> the low<br />

cost <strong>of</strong> the medium (water), low energy use for pressurization compared with gas or air, and the very high<br />

productivity that can be achieved (up to 30 tons/hour or about 500 kg/min). The primary limitations <strong>of</strong><br />

water atomization are powder purity and particle shape, particularly with more reactive metals and alloys.<br />

These generally give irregular powders <strong>of</strong> (relatively) high oxygen content.<br />

Schematic <strong>of</strong> water atomization process. The major components <strong>of</strong> a typical installation include<br />

a melting facility, an atomizing chamber, water pumping/recycling system, and powder<br />

dewatering and drying equipment. Melting <strong>of</strong> metals follows standard procedures. Air induction<br />

melting, arc melting, and fuel heating are suitable procedures.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

33<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.5. Composite Materials<br />

<strong>Thermal</strong> spraying, either as a coating or as a bulk structural consolidation process, has clearly<br />

demonstrated advantages for the production <strong>of</strong> composites. Difficult-to-process composites can be readily<br />

produced by thermal spray forming, with vacuum plasma spray the process <strong>of</strong> choice for the most<br />

reactive matrix materials. Particulate-, fiber-, and whisker-reinforced composites have all been produced<br />

for various applications. Particulate-reinforced wear-resistant coatings such as WC/Co, Cr3C2/NiCr, and<br />

TiC/NiCr are the most common applications and comprise one <strong>of</strong> the largest single thermal spray<br />

application areas.<br />

Fig. 1: Schematic microstructures <strong>of</strong> possible thermally spray-formed composites. (a) Deposit with a<br />

particulate-reinforced second phase. (b) Deposit with a whisker-reinforced second phase. (c) Deposit with<br />

a continuous-fiber-reinforced second phase<br />

Figure 1 shows schematically the diverse forms <strong>of</strong> composites that can be thermally spray formed.<br />

Whiskers <strong>of</strong> particles can be incorporated using so-called "engineered" powders, mechanical blending,<br />

and by co-injecting different materials into a single spray jet. Mechanical blends and co-injection, although<br />

useful, have been found to result in segregation <strong>of</strong> the reinforcing phase and, in many cases, degradation<br />

<strong>of</strong> the second-phase whiskers or particles. <strong>Thermal</strong> spray composite materials can have reinforcingphase<br />

contents ranging from 10 to 90% by volume, where the metal matrix acts as a binder, supporting<br />

the reinforcing phase. The ability to consolidate such fine-grained, high reinforcing phase content<br />

materials is a major advantage <strong>of</strong> thermal spray over other methods.<br />

<strong>Thermal</strong> spraying <strong>of</strong> composite materials with discontinuous reinforcements, such as particulates or short<br />

fibers, is usually accomplished by spraying powders or powder blends. Investigators have developed<br />

techniques for the production <strong>of</strong> continuous fiber-reinforced materials that overcome the "line-<strong>of</strong>-sight"<br />

limitations <strong>of</strong> thermal spray processes. This includes "monotape" fabrication techniques, where<br />

continuous fibers are prewrapped around a mandrel and a thin layer <strong>of</strong> a metal, ceramic, or intermetallic<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

34<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

matrix material is sprayed. Plasma, HVOF, and wire arc spray have been used, although in the cases <strong>of</strong><br />

intermetallics and high-temperature alloys, controlled atmosphere plasma spray (VPS) has generally<br />

been used. The fibers are thus encapsulated within thin monolayer tapes and are subsequently removed<br />

for consolidation to full density by hot pressing with preferred fiber orientations, producing continuously<br />

reinforced bulk composites.<br />

Powders for <strong>Spray</strong>ed Composites<br />

Discontinuously reinforced composites produced using thermal spray techniques use either composite<br />

powders or direct reactive synthesis approaches, as described below. Powders can be produced<br />

mechanically, chemically, thermomechanically, or by using high-temperature synthesis. "Engineered<br />

powders" defines powders in which different phases are incorporated to produce a "microcomposition" <strong>of</strong><br />

the final desired structure. Typically these powders contain the desired sizes, size distributions, and<br />

morphologies <strong>of</strong> the equilibrium phases. These powders also permit the introduction <strong>of</strong> higher<br />

concentrations <strong>of</strong> phases than those normally achievable through conventional melt or reaction<br />

processing.<br />

Fig. 2: Schematic representations <strong>of</strong> typical "composite" thermal spray powders.<br />

Figure 2 shows two types <strong>of</strong> powder that could be produced in this way: short ceramic fibers within a<br />

metal matrix and an intermetallic-matrix material reinforced with more ductile phases. The latter has been<br />

found to be a viable approach for increasing the fracture toughness (KIc) <strong>of</strong> intermetallics. Varying<br />

reinforcing phase combinations, compositions, morphologies, and distributions can be produced. The<br />

rapid heating and cooling experienced by these powders during thermal spray forming limits dissolution<br />

and degradation <strong>of</strong> the phases, which remain relatively unchanged after consolidation, although some<br />

microstructural refinement and solutioning can take place. Generally, more significant changes occur<br />

during conventional powder consolidation processes because <strong>of</strong> the longer processing times.<br />

Source: <strong>Thermal</strong> <strong>Spray</strong> Forming <strong>of</strong> Materials, R. Knight and R.W. Smith, Powder Metal Technologies and<br />

Applications, Vol 7, ASM Handbook, ASM International, 1998, p 408-419.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

35<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.6. Powder Testing and Characterization: Image Analysis<br />

Image analysis method <strong>of</strong> determining powder size and shape. (a) Original image. (b) Processed image,<br />

providing shapes that are less ambiguous and easier to measure.<br />

Image analysis techniques are used to digitize an image <strong>of</strong> the powder particles and, using image<br />

enhancement, thresholding, and binary processing routines, yield a size and shape distribution. While<br />

simple in theory, this method is difficult in practice due to the need to separate touching powder particles<br />

and to accumulate a significant number <strong>of</strong> images to properly statistically represent the specimen.<br />

Provided these issues can be overcome, image analysis is a precise method <strong>of</strong> measuring particle size<br />

and shape.<br />

Source: Powder Testing and Characterization, Walter Riggs, TubalCain Co., Inc., as published in<br />

Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 274, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

36<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.7. Powder Characteristics: Morphology<br />

Micrograph <strong>of</strong> irregularly shaped water-atomized powders having high surface areas (top), and<br />

micrograph <strong>of</strong> spherical gas-atomized powder (bottom).<br />

Manufacturing process variations can effect changes in particle morphology, or shape. Morphology<br />

affects various parameters including feedability, apparent density, sprayability, unmelted particles in the<br />

coating, porosity, and vaporization. Water-atomized powders are one example <strong>of</strong> this phenomenon,<br />

where the morphology may range from almost perfect spheres to highly irregular, convoluted shapes,<br />

depending on the manufacturing process parameters. Irregularly shaped particles do not feed as easily<br />

as spherical or equiaxed (particles having the same dimension in every direction) powders, because they<br />

pack together more easily. Because a sphere has the lowest surface-to-volume ratio <strong>of</strong> any geometry,<br />

any nonspherical shape can adsorb more surface moisture than a sphere. Gas-atomized powders usually<br />

are more spherical, and, therefore, typically flow very well.<br />

Source: J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, 2004, p 89<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

37<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.8. Particle Size Distribution Plots<br />

Particle classification curves, (a) Cumulative plot (b) Frequency plot<br />

A particle size distribution can be graphically represented as a cumulative plot or a frequency plot. These<br />

graphs are essentially identical since they can be derived from each other; the summation <strong>of</strong> the<br />

individual frequencies at each size enables the cumulative plot to be developed. The key features<br />

conveyed from particle size distribution data are:<br />

• The most commonly occurring particle size (i.e., the modal value) is indicated by the peak in the<br />

frequency plot. The mean value can be determined from the diameter at the 50 wt% value <strong>of</strong> the<br />

cumulative graph.<br />

• The sharpness <strong>of</strong> the plot gives a qualitative indication <strong>of</strong> the spread in particle size. A sharp peak<br />

indicates that the particles are <strong>of</strong> similar size, whereas a flat or broad curve indicates that the<br />

particle sizes are spread over a wide range <strong>of</strong> values.<br />

• The maximum and minimum values in measurable diameters allow the particle size range to be<br />

determined.<br />

• It is important to look for a distribution that does not fluctuate, because that indicates a bimodal<br />

distribution <strong>of</strong> particle sizes. Such powders do not feed reliably.<br />

Source: Particle Characterization, Christopher C. Berndt, Stony Brook <strong>University</strong>, as published in<br />

Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 164.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

38<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.9. Wire and Rod Feeders<br />

Wire feeding principles and devices<br />

Wire and rod feeders are simple mechanical devices consisting <strong>of</strong> electrical or air-driven motors, drive<br />

rolls, and associated speed controls. Wires and rods need to be fed continuously and uniformly into the<br />

flame or arc. The injection point <strong>of</strong> the wire/rod and feed rate are extremely important in controlling the<br />

coating quality. Uniformity <strong>of</strong> feed depends on motor drives having sufficient torque to overcome friction<br />

within the system, and in the case <strong>of</strong> wires, from the wire spools. Drive roll design is another important<br />

aspect <strong>of</strong> uniform feed. Drive rolls must grip the wire/rod sufficiently to prevent slipping while not<br />

deforming the wire or crushing the rod. Drive rolls vary according to the wire properties. Rolls are usually<br />

made <strong>of</strong> metal or fiber-reinforced phenolic. Phenolic rolls are used for ceramic rods, whereas metal rolls<br />

are used for wires. Metal rolls can be knurled for hard wires or grooved for wires; many variations exist.<br />

Source: Process Control Equipment, revised by Daryl E. Crawmer, <strong>Thermal</strong> <strong>Spray</strong> Technologies Inc., as<br />

published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 86.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

39<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.10. Fluidized-Bed Powder Feeders<br />

Fluidized-bed powder feeders are commonly used in thermal spray processes. In these units, a “bed” <strong>of</strong><br />

powder is continuously agitated by a fluidizing gas, in this case the carrier gas. The intent by design is to<br />

suspend a controlled volume <strong>of</strong> powder in a bed using a controlled volume <strong>of</strong> gas such that a sample <strong>of</strong><br />

powder can be drawn <strong>of</strong>f at a uniform rate. To accomplish this, the bed must be in continuous motion—<br />

that is, fluidized—and the volumes <strong>of</strong> gas and powder being removed must be continuously replaced.<br />

Vibrators and/or gravity feed are used to assist and maintain flow from the hopper into the bed. Powder<br />

feed rate in fluidized-bed feeders is affected by powder characteristics and particle size distribution.<br />

Source: Process Control Equipment, Revised by Daryl E. Crawmer, <strong>Thermal</strong> <strong>Spray</strong> Technologies Inc., as<br />

published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 88.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

40<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.11. Gravity-Based Powder Feeding<br />

Gravity- and vibration-based powder feeding system<br />

In a gravity-based powder feeding device, the powder falls into the thermal spray source. Some vibratory<br />

action, normally achieved by using compressed air to drive a ball bearing around a bearing race, may be<br />

supplied to agitate the powder into the gas stream and prevent blockages. The throttle control for the<br />

powder flow is a compressible rubber grommet that can be pinched to either allow or stop powder flowing<br />

into the carrier gas stream. This device relies on a full canister <strong>of</strong> powder so that the gravity head <strong>of</strong><br />

material is approximately constant. There are no powder delivery tubes connecting the powder supply to<br />

the thermal spray source. Uneven and somewhat sporadic powder flow <strong>of</strong>ten occurs when the canister<br />

nears an empty condition. The powder flow rate is controlled by altering the inside diameter <strong>of</strong> a rubber<br />

grommet. This type <strong>of</strong> powder feeding device is simple and robust, but not particularly accurate or<br />

reproducible with respect to powder flow rate.<br />

Source: Feedstock Material Considerations, Christopher C. Berndt, as published in Handbook <strong>of</strong> <strong>Thermal</strong><br />

<strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 138.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

41<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

3.12. Powder Feeding Systems: Rotating Wheel Devices<br />

Rotating wheel (a) and rotating disk (b) powder feeding systems<br />

A rotating wheel device delivers precise parcels <strong>of</strong> powder to the powder feed delivery tube. The wheel<br />

can be mounted either vertically or horizontally. When oriented in the vertical direction (in the geometry <strong>of</strong><br />

a simple water wheel), the powder is dropped into the carrier gas stream. A gas tube is connected<br />

between both sides <strong>of</strong> the rotating wheel to equalized gas pressures on both sides <strong>of</strong> the wheel.<br />

Otherwise, the powder would be blown <strong>of</strong>f the pockets that are machined into the outside rim <strong>of</strong> the<br />

wheel. The horizontally oriented wheel (or disk) has slots that pass over the gas delivery system. In this<br />

fashion, powder is taken out <strong>of</strong> the slots and enters the carrier gas stream.<br />

In both cases, the powder delivery rate is controlled by (a) the physical dimensions <strong>of</strong> the slots or pockets<br />

that hold the powder and (b) the rotational speed <strong>of</strong> the wheel. In either geometry, it is easy to overload<br />

the carrier gas with solids and either block the powder feed tube or cause the saltation effect by using too<br />

high a powder feed rate or an insufficient carrier gas flow rate. Of course, the dimensions <strong>of</strong> the powder<br />

feed tube and the powder injection port, as well as the pressure conditions into which the powder and<br />

carrier gas are injected, also influence the powder delivery.<br />

Source: Feedstock Material Considerations, Christopher C. Berndt, Stony Brook <strong>University</strong> (at the time <strong>of</strong><br />

publish date), as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM<br />

International, p 139, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

42<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

4. Surface Preparation<br />

4.1. Dry Abrasive Grit Blasting<br />

Mechanical bonding, showing a properly grit-blasted surface with a large number <strong>of</strong> reverse draft pits into<br />

which sprayed particles flow to achieve positively keyed mechanical interlocking<br />

Dry abrasive grit blasting is the most commonly used surface roughening technique. Dry abrasive<br />

particles are propelled toward the substrate at relatively high speeds. On impact with the substrate, the<br />

sharp angular particles act like chisels, cutting small irregularities into the surface. The amount <strong>of</strong><br />

substrate (plastic) deformation is a function <strong>of</strong> the angularity, size, density, and hardness <strong>of</strong> the particles<br />

and the speed and angle at which the particles are directed toward the substrate. The action <strong>of</strong> the<br />

grit/shot causes irreversible plastic deformation <strong>of</strong> the surface material. The material beneath the<br />

deformed surface material remains elastic and tries to return the deformed surface material to its original<br />

(shorter) length. The resulting balance <strong>of</strong> forces places the outer layer <strong>of</strong> material in residual compressive<br />

stress, while the underlying material is in tension. To avoid reduction <strong>of</strong> fatigue life, titanium substrates<br />

are sometimes shot peened before dry abrasive grit blasting to reduce the underlying tensile stresses,<br />

effectively balancing the compressive/tensile forces. Irregularities on a perfectly grit-blasted surface will<br />

have a large number <strong>of</strong> reverse draft pits, providing a partially hidden space into which sprayed molten<br />

particles can flow. Coatings sprayed to this surface are said to be positively keyed, creating what is<br />

commonly referred to as a mechanical interlocking bond.<br />

Source: Coating Processing, Frank N. Longo, Longo Associates, as published in Handbook <strong>of</strong> <strong>Thermal</strong><br />

<strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 111.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

43<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

4.2. Alumina Blasting Abrasives<br />

Aluminium oxide (Al2O3) particles. 130x. (a) 240 mesh (b) 150 mesh<br />

Alumina (Al2O3) grit works very well in blasting cabinets. The specific gravity <strong>of</strong> alumina is just over half<br />

that <strong>of</strong> chilled iron, so that this type <strong>of</strong> abrasive is readily picked up by the suction feed and is effectively<br />

accelerated through the blasting nozzle. The rate <strong>of</strong> breakdown is not excessive, except when very high<br />

air pressures are used to blast hardened steel or other very hard substrate materials. Alumina, fused and<br />

pure, is extremely hard (>9 on the Mohs scale), and when properly crushed has sharp cutting edges.<br />

There are two grades: C (coarse) and F (fine). A good all-purpose grit for general use in a blasting<br />

cabinet would be to start with a 50/50 mixture <strong>of</strong> C and F, with periodic small additions <strong>of</strong> the C grade to<br />

maintain a balanced grit range <strong>of</strong> the mix. If the -50 mesh fraction is occasionally removed by screening, it<br />

can be set aside and used where a fine, light grit blast is needed.<br />

Source: From Coating Processing, Frank N. Longo, Longo Associates, as published in Handbook <strong>of</strong><br />

<strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 114.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

44<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

4.3. Machining and Macroroughening<br />

Comparison <strong>of</strong> thermal spray coatings deposited on macroroughened and smooth surfaces. (a) <strong>Spray</strong>ed<br />

metal over grooves; shrinkage constrained by grooves (b) <strong>Spray</strong>ed metal on smooth surface; effect <strong>of</strong><br />

shear stress on bond due to shrinkage<br />

Macroroughening is usually accomplished by machining grooves or threads into the surface to be<br />

sprayed. Typically, rough machined surfaces are also grit blasted prior to spraying. Surfaces roughened<br />

to this magnitude are <strong>of</strong>ten used for thick coatings to restrict shrinkage stresses and to disrupt the<br />

lamellar pattern <strong>of</strong> particle deposition in order to break up the shear stresses parallel to the substrate<br />

surface. Groove threads should be considered for coatings used on machine-element repairs for<br />

thicknesses greater than 1.3 mm (0.050 in.) on the radius, for high shrinkage coatings, and for coatings<br />

that may be subject to high shear stresses (i.e., loads parallel to the interface).<br />

Source: From Coating Processing, Frank N. Longo, Longo Associates, adapted from H.S. Ingham and<br />

A.P. Shepard, Eds. Metallizing Handbook, 7th ed., Metallizing Engineering Co. Inc., (Metco), p-A 10,<br />

1959, as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International,<br />

2005, p 115.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

45<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

5. Coating Processing and Operation<br />

5.1. Part Configuration and Coating Location Considerations<br />

<strong>Thermal</strong> spray stream positions. (a) Good (b) Acceptable (c) Minimum acceptable (d)<br />

Formation <strong>of</strong> porosity in sprayed coatings when the spray angle is reduced to approximately<br />

45 degrees. Particles impacting at angles <strong>of</strong> less than 90 degrees create a shadowing effect<br />

that results in increased coating porosity.<br />

After identifying an area to be coated, determine whether the part is symmetrical about the coating<br />

location, which would allow fixturing for rotation and/or translation about an axis perpendicular to the<br />

spray torch. The stream <strong>of</strong> spray particles should impact the target surface as close to normal (90<br />

degrees) as possible. The minimum acceptable impact angle is 45 degrees to the target area. A 45<br />

degree spray angle impact should be used only as a last resort, bearing in mind that some coating<br />

property deviation from the optimum can be expected. Both bond strength and adhesion will be<br />

compromised. Porosity increases dramatically as the angle <strong>of</strong> the spray torch moves from normal to 45<br />

degrees.<br />

Source: Coating Operations, Frank N. Longo, Longo Associates, as published in Handbook <strong>of</strong> <strong>Thermal</strong><br />

<strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 109.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

46<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

5.2. Contact and Shadow Masking<br />

Coatings resulting from contact masks (a) versus shadow masks (b)<br />

Masking protects component areas next to the thermal spray target zone from impact by overspray<br />

particles. Methods include metal “shadow” masks, high temperature tapes, and paint-on masking<br />

compounds.<br />

Metal shadow masks are placed approximately 2-3 times the total coating thickness away from the part to<br />

be sprayed and in front <strong>of</strong> the spray torch. The unwanted spray collects on the mask and is prevented<br />

from reaching the substrate.<br />

Tape, or contact, masks are applied by hand wrapping areas that do not require a coating. After spraying,<br />

the tape is easily removed. Many different tapes are commercially available. The best tapes have<br />

sufficient toughness to resist grit blasting and sufficient heat resistance to withstand the hot gas and<br />

particle impact during spraying.<br />

With a shadow mask, the coating ends in a narrow feathered band rather than a sharply defined edge as<br />

with a contact mask. The sharp edge could lead to chipping if the coating is thick, or it could serve as a<br />

stress raiser that leads to debonding <strong>of</strong> the coating.<br />

Source: F.N. Longo, Coating Processing, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.),<br />

ASM International, 2005, p 117.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

47<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

5.3. <strong>Thermal</strong> <strong>Spray</strong> Pattern<br />

<strong>Spray</strong> pattern illustrating particle deposition and the effect <strong>of</strong> size and debris on thickness and porosity in<br />

cross section.<br />

All thermal spray processes involve molten or semi molten droplets or particles travelling at some velocity<br />

in a gas stream and impacting onto a substrate to form a coating. A spray pattern cross section in planar<br />

view, parallel to the substrate, is circular or oval in shape. The fastest and densest deposits will build up<br />

at the center <strong>of</strong> the jet, where most particles are entrained and where the highest degree <strong>of</strong> melting<br />

occurs. Moving radially out from the center (where fewer particles are entrained and where they tend to<br />

be coarser and perhaps semi molten), combined with impact at angles <strong>of</strong> less than 90 degrees, results in<br />

increased porosity. At the jet periphery, fine particles oxidize readily due to entrained air from the<br />

surrounding atmosphere and deposit as debris. Small particles oxidize rapidly, sometimes completely, to<br />

form the major source <strong>of</strong> oxide inclusions in sprayed coatings.<br />

Source: Coating Operations, Frank N. Longo, Longo Associates, as published in Handbook <strong>of</strong> <strong>Thermal</strong><br />

<strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 121, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

48<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

5.4. Relationship between Particle Melting and Coating Structure<br />

The effect <strong>of</strong> the degree <strong>of</strong> particle melting at or just before impact on shape and coating structure.<br />

(a) Particle is heat s<strong>of</strong>tened or beginning to resolidify. At impact, it does not flow out and begins to<br />

lift at the edges. (b) Properly melted particle impacts and flows to form a well-bonded classical<br />

lamellar splat shape. (c) Superheated particle impacts and splashes, creating debris, satellites, and<br />

dusting.<br />

In addition to the effects <strong>of</strong> peripheral particles stemming from the spray pattern, consideration must be<br />

given to center-stream particles and how variations in the degree <strong>of</strong> melting affect coating structure.<br />

Particles in the center <strong>of</strong> the spray pattern may include one or all <strong>of</strong> the following states on impact:<br />

• Fully molten, or at just above the materials melting temperature<br />

• Superheated, well above the melting temperature and perhaps close to the vaporization point<br />

• Semimolten, with the outside liquid but the core still solid<br />

• Molten then solidified while in-flight before impact<br />

Fully melted particles at or just above the material melting point arrive at the substrate, impact, flow, and<br />

flatten. Particle material spreads and cools rapidly as heat is conducted into the substrate. The classic<br />

49<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

lamellar splat morphology is thus achieved. Lamellar particle thickness depends on particle velocity,<br />

substrate surface tension, starting particle size, and substrate temperature. Superheated particles may<br />

splatter on impact, sending out radial splashes <strong>of</strong> fine droplets—generally spherical—that end up in the<br />

coating as satellites or debris. Debris formed from splatter differs from debris produced in the jet in that<br />

splatter lodges in the coating, building up at the first ridge that stops its radial travel. Air blasting does not<br />

remove splatter debris. Superheating is a condition that generally should be avoided. Particles that<br />

resolidify in flight after having been melted may not deposit; it they do, they may appear as unmelted<br />

particles but with clearly discernable oxide layers on the particle surfaces.<br />

Source: Coating Operations, Frank N. Longo, Longo Associates, as published in Handbook <strong>of</strong> <strong>Thermal</strong><br />

<strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 123, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

50<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

5.5. Air-Cooling Devices to Minimize Debris<br />

Positioning <strong>of</strong> an air-cooling device to blow debris <strong>of</strong>f the surface before it<br />

rotates into the center spray region and is incorporated into the coating.<br />

To spray large target areas, an ancillary air-cooling device can be attached to the spray torch, such that a<br />

stream <strong>of</strong> compressed air blows peripheral debris from the substrate surface in advance <strong>of</strong> the traversing<br />

spray stream. Debris is blown <strong>of</strong>f the surface before it sticks and passes under “good” spray material to<br />

become permanently incorporated into the coating. When spraying flat surfaces, air coolers should be<br />

positioned on both sides <strong>of</strong> the spray stream to help remove debris. Air coolers are usually adjusted to<br />

precede and follow the spray torch along the transverse direction. While air is used to blow loose debris<br />

<strong>of</strong>f the surface, it also helps cool the surface, as the name implies.<br />

Source: Gravity- and vibration-based powder feeding system. Coating Operations, Frank N. Longo,<br />

Longo Associates, as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM<br />

International, 2005, p 122.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

51<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

5.6. Influence <strong>of</strong> Temperature Control on Substrate/Coating<br />

Effect <strong>of</strong> coating shrinkage on interfacial shear stresses. The sprayed metal cools, creating a<br />

tensile stress in the coating and a compressive stress in the underlying substrate material. These<br />

stresses may deform the substrate or weaken the bond between the sprayed coating and the<br />

substrate, leading to debonding <strong>of</strong> the coating. the bottom illustration shows deformation <strong>of</strong> a thinsection<br />

substrate and debonding <strong>of</strong> the coating due to stresses generated by cooling <strong>of</strong> the<br />

sprayed metal.<br />

When substrate preheating has been completed and spraying has begun, it is necessary to control the<br />

temperature <strong>of</strong> both the sprayed coating and the substrate to avoid substrate degradation, oxidation, and<br />

shrinkage/expansion.<br />

Substrate Degradation. For those substrate materials susceptible to physical changes at high<br />

temperatures, irreparable damage can be inflicted when temperature is not controlled. Physical properties<br />

(e.g., hardness) established by heat treating are most likely to change.<br />

Oxidation. Coating layers deposited and allowed to reach high temperatures during spraying will oxidize<br />

and appear to darken. Surface oxides so formed leave a plane <strong>of</strong> weakness in the coating and could lead<br />

to delamination under applied stresses. The critical oxidizing temperature for each material varies:<br />

nevertheless, keeping surface temperatures below 150 to 205 °C (300 to 400 °F) is a good general rule<br />

when spraying under air conditions.<br />

Shrinkage/Expansion. <strong>Spray</strong>ed coatings shrink as they cool, exerting considerable shear stress on the<br />

substrate and leading to warpage and possible separation. Substrate materials having a high coefficient<br />

<strong>of</strong> thermal expansion that are allowed to expand because <strong>of</strong> high temperatures during the spray process<br />

will experience major debonding at the substrate/coating interface.<br />

Source: H.S. Ingham and A.P. Shepard, Ed., Metallizing Handbook, 7th ed., Metallizing Engineering Co.<br />

Inc., (Metco), 1959, p A-65; as published in Coating Operations, Frank N. Longo, Longo Associates,<br />

Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 125, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

52<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

6. Coating Structures, Properties and Materials<br />

6.1. Microstructures <strong>of</strong> thermal spray coatings<br />

<strong>Thermal</strong> spray coatings consist <strong>of</strong> many layers <strong>of</strong> thin, overlapping, essentially lamellar particles,<br />

frequently called splats. Cross sections <strong>of</strong> detonation gun deposited alumina and titania are shown in the<br />

photomicrograph. Generally, the higher-particle-velocity coating processes produce the most dense and<br />

better bonded coatings, both cohesively (splat-to-splat) and adhesively (coating-to-substrate).<br />

Metallographically estimated porosities for detonation gun coatings and some HVOF coatings are less<br />

than 2%, whereas most plasma sprayed coating porosities are in the range <strong>of</strong> 5 to 15%. The porosities <strong>of</strong><br />

flame sprayed coatings may exceed 15%.<br />

Microstructure <strong>of</strong> detonation gun deposited alumina and titania. As-polished<br />

Most <strong>of</strong> the thermal spray processes lead to very rapid quenching <strong>of</strong> the particles on impact. Quench<br />

rates have been estimated to be 104 to 106 °C/s for ceramics and 106 to 108 °C/s for metallics. As a<br />

result, the materials deposited may be in thermodynamically metastable states, and the grains within the<br />

splats may be submicron-size or even amorphous.<br />

The metastable phases present may not have the expected characteristics, particularly corrosion<br />

characteristics, <strong>of</strong> the material, and this factor should be kept in mind in the selection <strong>of</strong> coating<br />

compositions.<br />

Source: R.C. Tucker, Jr., <strong>Thermal</strong> <strong>Spray</strong> Coatings, ASM Handbook, Vol 5, Surface Engineering, ASM<br />

International, 1994, p 507<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

53<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

6.2. Structure <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Coatings<br />

Schematic showing the buildup <strong>of</strong> a thermal spray coating. Molten particles spread out<br />

and deform (splatter) as they strike the target, at first locking onto irregularities on the<br />

substrate, then interlocking with each other. Voids can occur <strong>of</strong> the growing deposit traps<br />

air. Particles overheated in the flame become oxidized. Unmelted particles may simply be<br />

embedded in the accumulating deposit. Courtesy <strong>of</strong> the Materials Engineering Institute,<br />

ASM International.<br />

The deposited structures <strong>of</strong> thermal spray coatings differ from those <strong>of</strong> the same material in the wrought<br />

form because <strong>of</strong> the incremental nature <strong>of</strong> the coating buildup and because the coating composition is<br />

<strong>of</strong>ten affected by reaction with the process gases and the surrounding atmosphere while the materials are<br />

in the molten state.<br />

For example, where air or oxygen is the process gas, oxides <strong>of</strong> the material applied may be formed and<br />

become part <strong>of</strong> the coating. The as-applied structures <strong>of</strong> all thermal spray coatings are similar in their<br />

lamellar nature; the variations in structure depend on the particular process, the processing parameters<br />

and techniques, and the material applied.<br />

The diagram illustrates the microstructure that results. As shown in this figure, the molten particles spread<br />

out and deform (splatter) as they impact the substrate, at first locking onto irregularities on the roughened<br />

surface, then interlocking with each other. Voids can occur if the growing deposit traps air. Particles<br />

overheated in the flame become oxidized. Unmelted particles may simply be embedded in the<br />

accumulating deposit.<br />

Source: J.R. Davis, Hardfacing, Weld Cladding, and Dissimilar Metal Joining, ASM Handbook, Vol 6,<br />

Welding, Brazing, and Soldering, ASM International, 1993, p. 803.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

54<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

6.3. Sources <strong>of</strong> Coating Porosity<br />

Porosity created by shadowing resulting from <strong>of</strong>f-axis spraying<br />

One <strong>of</strong> several sources <strong>of</strong> porosity is shadowing due to the angle <strong>of</strong> impingement <strong>of</strong> the spray stream.<br />

Shadowing generally is associated with coatings sprayed at angles below 45 degrees from the optimal<br />

“normal” angle <strong>of</strong> incidence. Coating porosity decreases (i.e., density increases) as the angle <strong>of</strong> spray<br />

approaches 90 degrees (normal to the surface being coated). More advanced coating systems are<br />

sprayed using tighter tolerances on fixture alignment. Plasma coatings may be sprayed at ±15 degrees,<br />

whereas low-end coatings may be sprayed at ±30 degrees. In shadowing, surface protrusions build up<br />

and then shadow interstices <strong>of</strong> voids adjacent to and behind the protrusions. These protrusions produce<br />

even more shadowing, particularly as the angle <strong>of</strong> incidence is lower relative to the surface; that is, less<br />

than 90 degrees.<br />

Source: D.E. Crawmer, Coating Structures, Properties, and Materials, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong><br />

<strong>Technology</strong>, J.R. Davis (Ed.), ASM International, 2005, p 50.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

55<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

6.4. Influence <strong>of</strong> Porosity on Coating Properties<br />

Typical thermal spray coating defects<br />

Porosity is an important coating feature that strongly influences coating properties. As with oxide<br />

inclusions, porosity can be a desirable characteristic. This discussion takes the general position that<br />

porosity is undesirable. Porosity creates poor coating cohesion and allows for higher wear and corrosion<br />

rates. Porosity is generally associated with a higher number <strong>of</strong> unmelted or resolidified particles that<br />

become trapped in the coating. Poor splat or particle cohesion leads to premature coating cracking,<br />

delamination, or spalling. Open porosity can interconnect to the coating interface, enabling corroding or<br />

oxidizing elements to attack the base material. Porosity can thus "short-circuit" the inherent corrosion<br />

resistance <strong>of</strong> a coating. For hardface or wear-resistant coatings, porosity lowers coating hardness and<br />

contributes to poor surface finishes, thus decreasing wear resistance. Porosity in wear coatings can also<br />

lead to the generation <strong>of</strong> coating fragments that break away and become abrasive cutting agents,<br />

increasing coating wear rates.<br />

Source: Coating Structures, Properties, and Materials, Revised by Daryl E. Crawmer, <strong>Thermal</strong> <strong>Spray</strong><br />

Technologies Inc., as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM<br />

International, p 49, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

56<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

6.5. Coat Bonding: Mechanical Interlocking<br />

Interlocking <strong>of</strong> coating particles<br />

Bonding <strong>of</strong> the coating to the substrate and cohesion between consecutive splats is affected, in rough<br />

order, by:<br />

• Residual stresses within the coating<br />

• Melting and localized alloying at the contact surfaces between particles and between the<br />

substrate and adjoining particles<br />

• Diffusion <strong>of</strong> elemental species across splat boundaries<br />

• Atomic-level attractive forces (van der Waals forces)<br />

• Mechanical interlocking<br />

Mechanical interlocking has been historically viewed as the main mechanism <strong>of</strong> thermal spray coating<br />

adhesion. Mechanical interlocking can play a part in coating adhesion and cohesion when the surfaces<br />

being coated have features that allow molten material to flow into and fill negative relief or where the part<br />

has negative relief, as with shafts. In this case, the bond between impacting particles and the surface is<br />

established largely through the impact <strong>of</strong> particles that flow and solidify around the substrate surface<br />

asperities. Substrate asperities with negative relief can be formed prior to coating by grit blasting and<br />

other mechanical surface preparation techniques, or by process-induced irregularities on the actual<br />

coating surface.<br />

Source: Coating Structures, Properties, and Materials, Revised by Daryl E. Crawmer, <strong>Thermal</strong> <strong>Spray</strong><br />

Technologies Inc., as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM<br />

International, p 51, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

57<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

7. Post Coating Operation<br />

7.1. Sealing<br />

Common sealants used with thermal spray coatings<br />

Organic material Characteristic<br />

Paints Water and solvent soluble<br />

Waxes Low-temperature melt<br />

Phenolics Air dry or heat cure<br />

Epoxy phenolics Air dry or heat cure<br />

Epoxy resins One-part catalyst<br />

Polyesters Air or heat cure or one-part catalyst<br />

Silicones Heat cure<br />

Polyurethanes Air dry or one-part catalyst<br />

Linseed oil Air dry<br />

Polyimides Heat cure<br />

Coal tars Air dry<br />

Anaerobics Cure in absence <strong>of</strong> air<br />

Source: J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, 2004, p 130<br />

<strong>Thermal</strong> <strong>Spray</strong> coatings, being somewhat porous by nature, <strong>of</strong>ten are sealed with organic materials that<br />

penetrate and fill the pores. The sealant is then cured, which effectively creates a barrier to penetration <strong>of</strong><br />

unwanted materials. Benefits <strong>of</strong> sealing include:<br />

• Preventing corrosive species (liquids and gases) from penetrating the coating and attacking the<br />

substrate<br />

• Preventing grinding debris from lodging in the coating<br />

• Enhancing interparticle cohesion<br />

• Providing special surface properties such as non-stick and release characteristics<br />

The choice <strong>of</strong> sealer dictates the curing temperature and time. The sealer should have a low viscosity to<br />

penetrate the coating to a satisfactory depth. Also, it is important to know the chemical resistance and<br />

service temperature <strong>of</strong> the sealant after curing. Sealants can be applied by brushing, spraying, and<br />

dipping.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

58<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

7.2. Recommended Metallographic Practice: Sectioning<br />

Proper sectioning <strong>of</strong> thermal spray coatings, (a) Sectioning <strong>of</strong> dense, nonfriable coatings, (b) Sectioning<br />

<strong>of</strong> porous, friable coatings.<br />

<strong>Thermal</strong> spray coating specimens should be sectioned perpendicular to their axis using a precision<br />

diamond saw with either a metal-bonded diamond wafering blade or an ultrathin aluminium oxide<br />

abrasive blade. Generally, an abrasive cut <strong>of</strong>f blade selected to cut the substrate effectively will be the<br />

best blade for the combination <strong>of</strong> thermal spray coating and substrate. The specimen should be rigidly<br />

clamped in the vise and positioned so that the cutting blade enters the coated side and exits the substrate<br />

side, thus substantially reducing fracturing <strong>of</strong> the coating. Friable, porous, or brittle coatings to be<br />

sectioned may be vacuum impregnated with epoxy mounting compound prior to sectioning to protect the<br />

specimen. Typical sectioning parameters are:<br />

• Speed: 2,500 to 3,500 rpm<br />

• Applied load: 300 to 500 gf<br />

• Lubricant: water soluble solution<br />

Source: ASM Handbook: Metallography and Microstructures, Vol.9, 2004, p 1040<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

59<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8. Coating Characterization and Testing<br />

8.1. Measuring Deposition Efficiency<br />

Variation <strong>of</strong> cold spray deposition efficiency <strong>of</strong> titanium coatings with powder feed rate<br />

Deposition efficiency (DE) is defined as an idealized measure <strong>of</strong> the percentage <strong>of</strong> particles introduced<br />

into a spray jet that actually deposits onto a flat substrate without overspray considerations. Deposition<br />

efficiency can be measured according to the following procedure:<br />

• Determine the mass (X) <strong>of</strong> a clean, grit blasted 4 in. × 6 in. × 0.125 in. (100 × 150 × 3 mm) flat<br />

steel ate<br />

• Place a known mass <strong>of</strong> powder into the powder feeder<br />

• After stabilizing the spray parameters and powder feed, spray material onto the steel plate for a<br />

known period <strong>of</strong> time (e.g., 60 s) at a predetermined feed rate (e.g., 20 g/min) ensuring that the<br />

spray pattern remains on the plate at all times. (The plate may be cooled as necessary using<br />

air/gas jets directed at the rear face <strong>of</strong> the plate.)<br />

• Measure the mass (Y) <strong>of</strong> the plate plus the coating.<br />

• Determine the gain in mass (Z) due to the coating deposited (Z = Y - X).<br />

• Divide the mass deposited in one minute (Z) by the powder feed rate and multiply by 100 to give<br />

DE in %.<br />

• Repeat the measurement a minimum <strong>of</strong> three times and take the average.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

60<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

Related terms: target efficiency, sticking efficiency.<br />

Notes:<br />

1) A scale capable <strong>of</strong> measuring to an accuracy <strong>of</strong> 0.1 g should be used to measure the changes in<br />

mass.<br />

2) Deposition efficiency is only useful ins<strong>of</strong>ar as it provides a measure for optimizing spray<br />

parameters.<br />

3) When substrate geometry, size, and overspray are taken into consideration, the ideal DE<br />

decreases substantially, yielding a true deposition efficiency, properly termed target efficiency<br />

(TE).<br />

Reference: J.R. Davis (Ed.), Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, ASM International, 2004, p 79<br />

Source: Dr. Richard Knight, FASM, Auxiliary Pr<strong>of</strong>essor and CPPM Director, Drexel <strong>University</strong>,<br />

Philadelphia, Pa.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

61<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8.2. Tensile Adhesion Testing <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Coatings<br />

The coating/fixture assembly used for tensile adhesion testing <strong>of</strong> thermal spray coatings. Dimensions are<br />

in inches (1 in. = 25.4 mm)<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

62<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

Adhesion is a property <strong>of</strong> major concern for thermal spray coatings because it is necessary for the coating<br />

to adhere to the substrate throughout the design life <strong>of</strong> the coating system. The standard coating/fixture<br />

assembly used for tensile adhesion testing <strong>of</strong> thermal spray coatings is adapted from ASTM C 633-01,<br />

“Standard Test Method for Adhesion or Cohesion Strength <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Coatings.” The coating is<br />

applied to a 25 mm (1 in.) diameter cylindrical bar that is threaded for gripping at its end. A somewhat<br />

viscous adhesive bonding agent is then applied to the surface <strong>of</strong> the coating, and the coating/bonding<br />

agent is applied downward in a suitable alignment fixture onto another bar. The entire assembly is placed<br />

into an oven for curing at moderate temperatures (120 to 175°C, or 250 to 350°F). Following the cure<br />

cycle, the assembly is pulled apart to determine its strength.<br />

Source: Walter Riggs, Testing <strong>of</strong> Coatings, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.),<br />

ASM International, 2005, p 262.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

63<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8.3. Bond Pull Testing <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Coatings<br />

A bond pull test measures the force required to separate a thermal spray coating from its substrate<br />

Adhesion (bond-strength) testing <strong>of</strong> thermal spray coatings is performed in accordance with ASTM C 633-<br />

01 “Standard Test Method for Adhesion or Cohesion Strength <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Coatings.” The test<br />

measures the force required to separate a thermal spray coating from its substrate. Depending on its<br />

bond strength, the coating can fail at any <strong>of</strong> several locations including the coating-substrate interface,<br />

within the coating, and at the interface between the coating and the bonding agent. For some coatings<br />

(such as HVOF WCCo), the adhesive strength <strong>of</strong> the coating is typically greater than the strength <strong>of</strong> the<br />

bonding agent. The two most common bonding agents are FM-1000 and EC-2214. FM-1000 comes in<br />

the form <strong>of</strong> a wafer, while EC-2214 is a liquid. Both bonding agents are applied between the coating and<br />

a mating metallic slug. After curing, the coating is subjected to increasing tensile load until failure. While<br />

FM-1000 can be used for any coating, EC-2214 should only be used for dense coatings. Because it is a<br />

liquid, it is possible for EC-2214 to penetrate a porous coating and bond to the substrate. As a result,<br />

artificially high bond strength values could be measured.<br />

Source: Doug Puerta, KHA Metallurgical, Portland, Oregon<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

64<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8.4. Rolling Contact Fatigue Failure Modes in <strong>Thermal</strong> <strong>Spray</strong> Coatings<br />

Rolling-contact fatigue failure modes in <strong>Thermal</strong> <strong>Spray</strong> cermet (WC-Co) and ceramic (Al2O3) coatings<br />

have been studied. Investigations relating to the RCF failure modes <strong>of</strong> TS coatings have classified the<br />

fatigue failure modes on the basis <strong>of</strong> surface and subsurface observations in pre- and post-RCF<br />

conditions.<br />

Rolling-contact fatigue failure <strong>of</strong> TS coatings are generally categorized in four main modes:<br />

• Abrasion<br />

• Delamination<br />

• Bulk failure<br />

• Spalling (M1–M4), as indicated in the diagram<br />

Rolling-contact fatigue failure modes <strong>of</strong> thermal spray cermet and ceramic coatings.<br />

Source: Rolling Contact Fatigue, R. Ahmed, Failure Analysis and Prevention, Vol 11, ASM Handbook,<br />

ASM International, 2002, p 941-956.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

65<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8.5. Residual Stress Determination in <strong>Thermal</strong> <strong>Spray</strong> Coatings<br />

Hole-drilling method for measuring residual stresses. (a) Typical three-element strain-gage rosette (b) Inplane<br />

strain components caused by release <strong>of</strong> residual stresses through the introduction <strong>of</strong> a hole<br />

The hole-drilling method for measuring residual stresses in thermal spray coatings involves drilling a<br />

shallow hole in the test specimen to a depth approximately equal to the hole diameter. Typical hole<br />

diameters range from 0.8 to 5.0 mm (0.03 to 0.20 in.). The creation <strong>of</strong> the hole redistributes the stresses<br />

in the material surrounding the hole. A specially designed three-element strain-gage rosette measures the<br />

associated partial strain relief. The in-plane residual stresses that originally existed at the hole location<br />

can then be calculated from the measured strain reliefs using the method described in ASTM E 837,<br />

“Measurements <strong>of</strong> Residual Stresses by Hole-Drilling Strain-Gage Method.” The ASTM standard also<br />

gives details <strong>of</strong> practical drilling procedures.<br />

Source: From Testing <strong>of</strong> Coatings, Walter Riggs, Tubal-Cain Co. Inc., and Ken Couch, Protech Lab<br />

Corp., as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM International,<br />

2005, p 271.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

66<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

8.6. <strong>Thermal</strong> Fatigue Testing <strong>of</strong> TBCs<br />

Schematic illustration <strong>of</strong> oxide surface temperature cycle used for thermal<br />

fatigue testing <strong>of</strong> ZrO3 coated test buttons 1.27 mm (0.050) thick, simulating a<br />

first-stage gas turbine outer air-seal application.<br />

Laboratory testing has been used, particularly in the coating development stage, to thermally cycle a<br />

thermal barrier coating (TBC) specimen, followed by post-test microscopic examination <strong>of</strong> spallation-type<br />

cracks. The thermal cycle for evaluation <strong>of</strong> thermal shock resistance uses rapid heating and cooling rates,<br />

using direct impingement flames or heating jets on the oxide face <strong>of</strong> the specimen, with little hold time at<br />

the maximum temperature. This test principally challenges the oxide layer, because the bond coat<br />

remains at relatively low temperatures due to the insulating nature <strong>of</strong> the zirconia layer and the short time<br />

at high temperatures.<br />

Source: Walter Riggs, Testing <strong>of</strong> Coatings, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.),<br />

ASM International, 2005, p 270.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

67<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9. Applications<br />

9.1. Coatings for Corrosive Environment: Conventional Coal-Fired Boilers<br />

(a) (b)<br />

Economizer tube bends being coated with high-chromium steel alloy using the twin-wire arc process (a).<br />

The tubes after coating are shown in (b). Courtesy <strong>of</strong> ASB Industries Inc.<br />

Utility boilers suffer from high wear and corrosion caused by coal slagging. In recent years, there has<br />

been substantial research showing extended life for such boilers when corrosion/wear zones are coated<br />

by plasma spray, twin-wire arc, and HVOF with Inconel systems or other proprietary alloys. Most are now<br />

coated with wire-arc applications <strong>of</strong> high-chromium alloys or chromium-nickel alloys that effectively<br />

protect them from high-sulfur corrosion for periods <strong>of</strong> up to five years. It is estimated that 4,650 m 2<br />

(50,000 ft 2 ) <strong>of</strong> new surfaces are coated each year.<br />

Source: Selected Applications, as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis<br />

(Ed.), ASM International, p 208, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

68<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.2. Elevated-Temperature Corrosion Applications<br />

<strong>Thermal</strong> spray coatings for elevated-temperature service<br />

Service temperature, °C (°F) Coating metal or alloy Coating thickness<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

µm mils<br />

Up to 550 (1020) Aluminium 175 7<br />

Up to 550 (1020) in the presence <strong>of</strong><br />

sulphurous gases<br />

Ni-43Cr-2Ti 375 15<br />

550 – 900 (1020 – 1650) Aluminium or aluminium-iron 175 7<br />

900 – 1000 (1650 – 1830) Nickel-chromium or MCrAlY 375 15<br />

900 – 1000 (1650 – 1830) in the<br />

presence <strong>of</strong> sulphurous gases<br />

Nickel-chromium (a)<br />

Aluminium (a)<br />

(a) Coating system consists <strong>of</strong> an aluminium layer deposited on top <strong>of</strong> a nickel-chromium layer<br />

Common applications involving high-temperature corrosion include coating exhaust stacks, chimneys,<br />

flues, rotary kilns and dryers, catalytic crackers, and furnace parts. These generally involve the use <strong>of</strong> an<br />

aluminum or nickel-chromium alloy coating. One specific high-temperature corrosion application that has<br />

met with great success in recent years is the coating <strong>of</strong> boilers in paper mills, power plants, and chemical<br />

plants. Water-wall tubes in these applications suffer severe corrosion because <strong>of</strong> the high sulfur content<br />

<strong>of</strong> the burning fuel, the high operating pressures, and abrasion. A 375 μm (15 mil) coating <strong>of</strong> Ni-43Cr-2Ti<br />

alloy <strong>of</strong>fers extremely good protection against this very severe corrosion at temperatures to 550 °C (1020<br />

°F).<br />

Source: M.L. Berndt and C.C. Berndt, <strong>Thermal</strong> <strong>Spray</strong> Coatings, Corrosion: Fundamentals, Testing, and<br />

Protection, Vol 13A, ASM Handbook, ASM International, 2003, p 803-813; as published in Introduction to<br />

Applications for <strong>Thermal</strong> <strong>Spray</strong> Processing, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.),<br />

ASM International, 2005, p 174.<br />

375<br />

100<br />

15<br />

4<br />

69<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.3. Oxidation Protection<br />

<strong>Thermal</strong> spray coatings for elevated-temperature service<br />

Service temperature, °C (°F) Coating metal or alloy Coating thickness<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

µm mils<br />

Up to 550 (1020) Aluminium 175 7<br />

Up to 550 (1020) in the presence <strong>of</strong><br />

sulphurous gases<br />

Ni-43Cr-2Ti 375 15<br />

550 – 900 (1020 – 1650) Aluminium or aluminium-iron 175 7<br />

900 – 1000 (1650 – 1830) Nickel-chromium or MCrAlY 375 15<br />

900 – 1000 (1650 – 1830) in the<br />

presence <strong>of</strong> sulphurous gases<br />

Nickel-chromium (a)<br />

Aluminium (a)<br />

(a) Coating system consists <strong>of</strong> an aluminium layer deposited on top <strong>of</strong> a nickel-chromium layer<br />

<strong>Thermal</strong> spray coatings are extensively used by industry to protect steel components and structures from<br />

heat oxidation at surface temperatures to 1095 °C (2000 °F). By ensuring long-term protection, thermal<br />

spray coatings show real economic advantages during the service lives <strong>of</strong> such items. Coatings are<br />

particularly effective in protecting low-alloy and carbon steels. The table lists thermal spray coatings for<br />

high-temperature applications.<br />

Aluminum has been widely used to protect such steels in a number <strong>of</strong> applications involving high surface<br />

temperatures up to 550°C (1020 °F). Nickel-chromium alloys and some <strong>of</strong> the MCrAlY alloys have also<br />

provided protection in severe environments (up to 1000 °C, or 1830 °F).<br />

375<br />

100<br />

15<br />

4<br />

70<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.4. <strong>Thermal</strong> Barrier Coatings<br />

Effect <strong>of</strong> yttria content on the performance <strong>of</strong> zirconia-base thermal barrier coatings<br />

<strong>Thermal</strong> barrier coatings (TBCs) consist <strong>of</strong> a low thermal conductivity ceramic layer deposited over a<br />

MCrAlY bond coat. Such coatings have been the subject <strong>of</strong> extensive study since the early 1960s, and<br />

the evolution <strong>of</strong> TBCs has been steady since then. The ceramic <strong>of</strong> choice for TBCs is zirconia (ZrO2), but<br />

pure zirconia exhibits a phase change as the temperature approaches 425 °C (800 °F), resulting in<br />

substantial volume change, which can subsequently generate internal stresses and lead to premature<br />

coating failure. Oxide stabilizers (yttria, calcia, magnesia, and ceria, for example) are added to the<br />

zirconia to prevent this mode <strong>of</strong> failure. Yttria (Y2O3) is the most widely used stabilizer for TBCs;<br />

commonly known as yttria-stabilized zirconia (YSZ). Partially stabilized material improves TBS<br />

performance. The amount <strong>of</strong> stabilizer also is important. Initially, fully stabilized zirconia containing 12 to<br />

20% Y2O3 was used, but it has been demonstrated that partially stabilized material improves TBC<br />

performance. It was reasoned that allowing a controlled amount <strong>of</strong> the ceramic to undergo a phase<br />

change created microcracks that relieved high temperature stresses and promoted longer life.<br />

The state-<strong>of</strong>-the-art thermal spray TBC today is a 6 to 8 wt% Y2O3-ZrO2 ceramic deposited on a thickness<br />

<strong>of</strong> 0.25 to 1 mm (0.010 to 0.040 in.), with 10 to 15% porosity. A bond coat <strong>of</strong> NiCrAlY or CoNiCrAlY is<br />

used at a thickness <strong>of</strong> 0.125 mm (0.005 in.).<br />

Source: R.A. Miller, <strong>Thermal</strong> Barrier Coatings for Aircraft Engines-History and Directions, <strong>Thermal</strong> Barrier<br />

Coatings Workshop, March 1995 (Cleveland, OH), NASA Center for Aerospace Information, 1995, p 17-<br />

34, as published in Selected Applications, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.),<br />

ASM International, p 178, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

71<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.5. Functionally Gradient Materials<br />

Functionally gradient materials (FGMs) are a rapidly growing application area with significant promise for<br />

the future production <strong>of</strong> (a) improved materials and devices for applications subject to large thermal<br />

gradients, (b) lower-cost clad materials for combinations <strong>of</strong> corrosion and strength or wear resistance, and<br />

(c) perhaps improved electronic material structures for batteries, fuel cells, and thermoelectric energy<br />

conversion devices.<br />

<strong>Thermal</strong> spray forming <strong>of</strong> such gradient structures has been proposed because <strong>of</strong> the unique ability this<br />

approach has <strong>of</strong> being able to deposit thin, individual layers <strong>of</strong> a wide range <strong>of</strong> materials--metals,<br />

intermetallics, and ceramics--thereby enabling layered or continuously graded structures to be produced.<br />

Schematic <strong>of</strong> a thermally sprayed FGM for burner nozzle applications.<br />

The most immediate application for FGMs is in thermally protective claddings, where large thermal<br />

stresses could be minimized and component lifetimes improved by "tailoring" the coefficients <strong>of</strong> thermal<br />

expansion, thermal conductivity, and oxidation resistance. These FGMs could, and indeed are, finding<br />

use in turbine components, rocket nozzles, chemical reactor tubes, incinerator burner nozzles, or other<br />

critical furnace components. The diagram illustrates an example <strong>of</strong> a thermally sprayed FGM proposed for<br />

the protection <strong>of</strong> copper using a layered FGM ceramic structure. Successful fabrication <strong>of</strong> this would have<br />

application as improved burner nozzles, molds, and furnace walls.<br />

Source: R. Knight and R.W. Smith, ASM Handbook, Vol 7, Powder Metal Technologies and Applications,<br />

ASM International, 1998, p 415.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

72<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.6. Wear and Abrasion Resistance <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> Coatings<br />

Abrasive wear data for selected thermal spray coatings<br />

Material Type Wear rate, mm 3 /1000 rev<br />

Carballoy 883 Sintered 1.2<br />

WC-Co Detonation gun 0.8<br />

WC-Co Plasma spray 16.0<br />

WC-Co Super D-Gun 0.7<br />

WC-Co High-velocity oxyfuel 0.9<br />

ASTM G65 dry sand/rubber wheel test, 50/70 mesh Ottawa silica, 200 rpm, 30<br />

lb load, 3000-revolution test duration<br />

Erosive wear data for selected thermal spray coatings<br />

Material Type Wear rate, µm/g<br />

Carballoy 883 Sintered 0.04<br />

WC-Co Detonation gun 1.3<br />

WC-Co Plasma spray 4.6<br />

AISI 1018 steel Wrought 21<br />

Silica-based erosion test; particle size, 15 µm; particle velocity, 129 m/s; particle<br />

flow, 5.5 g/min, ASTM Recommended Practice G 75<br />

One <strong>of</strong> the most important uses <strong>of</strong> thermal spray coatings is for wear resistance. They are used to resist<br />

virtually all forms <strong>of</strong> wear, including abrasive, erosive, and adhesive, in virtually every type <strong>of</strong> industry.<br />

The materials used range from s<strong>of</strong>t metals to hard metal alloys to carbide-based cermets to oxides.<br />

Generally, the wear resistance <strong>of</strong> the coatings increases with their density and cohesive strength, so the<br />

higher-velocity coatings such as HVOF and particularly detonation gun coatings provide the greatest wear<br />

resistance for a given composition.<br />

A variety <strong>of</strong> laboratory tests have been developed to rank thermal spray coatings and compare them with<br />

other materials. Examples <strong>of</strong> abrasive and erosive wear data are shown in Tables 10 and 11. It should be<br />

kept in mind that laboratory tests can seldom duplicate service conditions. Therefore these tests should<br />

only be used to help select candidate coatings for evaluation in service. Only rarely, with good baseline<br />

data, can any precise prediction <strong>of</strong> wear life in service be made from laboratory data.<br />

Source: R.C. Tucker, Jr., <strong>Thermal</strong> <strong>Spray</strong> Coatings, ASM Handbook, Vol 5, Surface Engineering, ASM<br />

International, 1994, p 508<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

73<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.7. Metallography <strong>of</strong> NiCrAl/Bentonite Abradable Coatings<br />

Typical microstructure for Ni-4Cr-4Al/Bentonite<br />

Abradable coatings (such as Ni-4Cr-4Al/Bentonite) entail a family <strong>of</strong> coatings used throughout jet<br />

engines, primarily as sacrificial coatings into which moving components wear. The coatings generally<br />

consist <strong>of</strong> a metallic phase and a nonmetallic phase, and contain relatively high porosity levels (to 40%).<br />

Typical locations for application <strong>of</strong> an abradable coating include the fan, and low- and high-pressure<br />

compressor sections.<br />

The composite nature <strong>of</strong> abradable coatings presents unique challenges from a metallography<br />

standpoint. Due to thickness and porosity considerations, vacuum impregnation with a low-viscosity cold<br />

mount epoxy is the recommended mounting method. To facilitate impregnation with fast-cure epoxies<br />

(which are typically more viscous than slow-cure epoxies), the resin can be heated to approximately<br />

150°F (65°C) prior to mixing with the hardener. Holding the epoxy at elevated temperature for 15 to 20<br />

minutes should result in a significant improvement in the viscosity <strong>of</strong> the epoxy.<br />

Source: Doug Puerta, KHA Metallurgical, Portland, Oregon<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

74<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.8. Abradable Seals<br />

Suggested temperature range for turbine compressor abradables. Those shown to<br />

be compatible with titanium blading also will function against steel and superalloy<br />

blades. However, the reverse is not true<br />

<strong>Thermal</strong>ly sprayed abradable seal coatings are employed in turbomachinery and can be machined in situ<br />

by the rotating components, such as blades, resulting in very close tolerances, reduced blade wear and<br />

overall improved engine efficiency. The abradable materials are deposited to a sufficient hardness to<br />

withstand gas and particle erosion, yet maintain abradability and functionality for the designed life span. A<br />

variety <strong>of</strong> coating materials are used; most notable for compressor cold sections are polymer filled<br />

aluminum alloys (325°C max. use temperature) or graphite or hexagonal boron nitride s<strong>of</strong>t phases,<br />

incorporated within an aluminum alloy, for use up to 450°C. These are generally applied using APS or<br />

combustion spray systems. Hot compressor section abradables include combustion sprayed NiCrAlbentonite<br />

(650°C max. temperature) or plasma-sprayed composites consisting <strong>of</strong> MCrAlYs, polyester, and<br />

hexagonal boron nitride (650-750°C max.). Figure shows temperature ranges for different compressor<br />

abradables.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

75<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

For maximum turbine service temperatures up to 1200°C, seals can be manufactured from porous yttriazirconia<br />

based ceramics. The use <strong>of</strong> alternative zirconia stabilizers such as Dy2O3 can lead to drastic<br />

improvements in thermal shock resistance as is seen in the case <strong>of</strong> the Dy2O3 stabilized coatings.<br />

Depending on the specific application conditions (temperature, incursion speed/blade rotation/material),<br />

tipped blades may or may not be required. A typical high temperature tip material is cubic boron nitride; in<br />

some cases, silicon carbide tips are also used. Benefits <strong>of</strong> DySZ ceramic abradables are the ability to cut<br />

under certain conditions with untipped blades and excellent thermal shock properties.<br />

Source: R.K. Schmid, F. Ghasripoor, M. Dorfman, and X. Wei, An Overview <strong>of</strong> Compressor Abradables,<br />

ITSC 2000, First International <strong>Thermal</strong> <strong>Spray</strong> Conference, May 2000, Montreal, Quebec, Canada, ASM<br />

International, p 1087-1093; as published in Selected Applications, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong><br />

<strong>Technology</strong>, J.R. Davis (Ed.), ASM International, p 178, 2005.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

76<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.9. Processing Electronic Devices: Beryllia<br />

A BeO meanderline substrate produced in near-net shape by inert chamber plasma spraying. Crosssectional<br />

dimensions are 1.016 x 1.016 mm (0.040 x 0.040 in.). the device was formed by milling the<br />

meanderline pattern in high-quality graphite. Inert chamber plasma spraying with a helium atmosphere<br />

was used to fill the milled pattern, the ceramic flashing being ground away. The net shape was completed<br />

by firing the structure at >800 °C (>1470 °F), which removed the graphite, leaving the freestanding<br />

ceramic.<br />

<strong>Spray</strong>ing dense, pure beryllia (BeO) requires higher substrate temperatures and very hot plasma.<br />

Occupational Safety and Health Administration (OSHA) requirements for handling this toxic, diseasecausing<br />

material are stringent and usually economically prohibitive. Excellent thermal and dielectric<br />

properties have been developed in plasma sprayed products, and structures have been produced for<br />

microwave and nuclear fusion applications.<br />

Source: From Selected Applications, as published in Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis<br />

(Ed.), ASM International, 2005, p 199.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

77<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.10. Aluminum Coatings and Zinc Coatings<br />

Corrosion rate <strong>of</strong> zinc and aluminium in NaCl aqueous solution as a function <strong>of</strong> pH<br />

The TSA and TSZ coatings are usually sprayed onto large structures via flame and electric arc-based<br />

processes, and with the coating material supplied in wire form to the spray gun. There are standards for<br />

the composition and diameter <strong>of</strong> these wires: ISO 14919 and AWS C2.25/C2.25M. Among these, four<br />

classes <strong>of</strong> materials—zinc, aluminum, Zn-15Al, and Al-5Mg—are specified for the general purpose <strong>of</strong><br />

corrosion protection in ISO 2063.<br />

Alloys with other compositions, as well as composite coatings, can be used for corrosion protection, but<br />

applications are limited. One way to fabricate composite coatings is to simultaneously feed wires <strong>of</strong><br />

different materials, such as zinc and aluminum, to the arc spraying gun. The molten droplets <strong>of</strong> the two<br />

metals mix on the substrate and form a composite coating called a pseudoalloy.<br />

Source: S. Kuroda and A. Sturgeon, <strong>Thermal</strong> <strong>Spray</strong> Coatings for Corrosion Protection in Atmospheric<br />

and Aqueous Environments, ASM Handbook, Vol 12B, Corrosion: Materials, ASM International, 2005, p<br />

423.<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

78<br />

Back to TOC<br />

Back to Overview


Compiled by Jo Ann Gan, Edited and advised by Christopher C. Berndt<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> <strong>Thermal</strong> <strong>Spray</strong> Group (SwinTS)<br />

Please contact Pr<strong>of</strong>. Christopher Berndt at cberndt @swin.edu.au for further enquiries<br />

9.11. <strong>Thermal</strong> <strong>Spray</strong> Polymer Coatings<br />

Selection <strong>of</strong> polymers applied as coatings via thermal spray processes<br />

Polymer Maximum temperature resistance<br />

Information and data acquired from ASM International <strong>Thermal</strong> <strong>Spray</strong> Society<br />

website at http://asmcommunity.asminternational.org/portal/site/tss/<strong>Spray</strong><strong>Tips</strong>/<br />

°C °F<br />

Ethylene methacrylic acid copolymer (EMAA) 40 – 60 105 – 140<br />

Polyethylene (PE) 40 – 80 105 – 175<br />

Polypropylene (PP) 70 160<br />

Nylon 6, 6 (PA) 65 150<br />

Polyphenylene sulphide (PPS) 110 230<br />

Polyethylene-tetrafluoroethylene copolymer (PE-TFE) 160 320<br />

Polyetheretherketone (PEEK) 125 255<br />

Liquid-crystal polymers (LCPs) 250 480<br />

Phenolic epoxy 130 265<br />

Polyimide (PI) 300 570<br />

Polymer spraying is a one-coat process that serves as both the primer and the sealer, with no additional<br />

cure times. Polymer powders are specified by their chemistry, morphology, molecular weight distribution<br />

or melt-flow index, and particle size distribution. <strong>Spray</strong> parameters must be selected to accommodate<br />

each particular polymer formulation. The heat source <strong>of</strong> the thermal spray torch can be a plasma,<br />

combustion flame, or combustion exhaust, as in the HVOF process. The technique is chosen on the basis<br />

<strong>of</strong> the polymer-melting characteristics. A simple combustion torch is well-suited for use with low meltingpoint<br />

polymers having a large processing window, such as polyethylene. High melting-point polymers<br />

may require plasma deposition for maximum quality to melt the polymer without causing oxidation.<br />

Therefore, processing parameters must be selected for each polymer chemistry and deposition<br />

technique.<br />

Source: Selected Applications, Handbook <strong>of</strong> <strong>Thermal</strong> <strong>Spray</strong> <strong>Technology</strong>, J.R. Davis (Ed.), ASM<br />

International, 2005, p 209.<br />

79<br />

Back to TOC<br />

Back to Overview

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!