14.12.2012 Views

surface pretreatment by phosphate conversion coatings – a review

surface pretreatment by phosphate conversion coatings – a review

surface pretreatment by phosphate conversion coatings – a review

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

130 Rev.Adv.Mater.Sci. 9 (2005) 130-177<br />

T.S.N. Sankara Narayanan<br />

SURFACE PRETREATMENT BY PHOSPHATE<br />

CONVERSION COATINGS <strong>–</strong> A REVIEW<br />

Corresponding author: T.S.N. Sankara Narayanan, e-mail: tsnsn@rediffmail.com<br />

© 2005 Advanced Study Center Co. Ltd.<br />

T.S.N. Sankara Narayanan<br />

National Metallurgical Laboratory, Madras Centre CSIR, Complex, Taramani, Chennai-600 113, India<br />

Received: April 22, 2005<br />

Abstract. Phosphating is the most widely used metal <strong>pretreatment</strong> process for the <strong>surface</strong><br />

treatment and finishing of ferrous and non-ferrous metals. Due to its economy, speed of operation<br />

and ability to afford excellent corrosion resistance, wear resistance, adhesion and lubricative<br />

properties, it plays a significant role in the automobile, process and appliance industries. Though<br />

the process was initially developed as a simple method of preventing corrosion, the changing<br />

end uses of <strong>phosphate</strong>d articles have forced the modification of the existing processes and<br />

development of innovative methods to substitute the conventional ones. To keep pace with the<br />

rapid changing need of the finishing systems, numerous modifications have been put forth in<br />

their development - both in the processing sequence as well as in the phosphating formulations.<br />

This <strong>review</strong> addresses the various aspects of phosphating in detail. In spite of the numerous<br />

modifications put forth on the deposition technologies to achieve different types of <strong>coatings</strong> and<br />

desirable properties such as improved corrosion resistance, wear resistance, etc., <strong>phosphate</strong><br />

<strong>conversion</strong> coating still plays a vital part in the automobile, process and appliance industries.<br />

Contents<br />

1. Introduction<br />

2. Chemical <strong>conversion</strong> <strong>coatings</strong><br />

3. Phosphating<br />

3.1. History and development of the<br />

phosphating process<br />

3.2. Chemistry of phosphating<br />

3.3. Acceleration of the phosphating<br />

process<br />

3.3.1. Chemical acceleration<br />

3.3.2. Mechanical acceleration<br />

3.3.3. Electrochemical acceleration<br />

3.4. Kinetics of the phosphating process<br />

3.5. Process details<br />

3.5.1. Cleaning<br />

3.5.2. Rinsing<br />

3.5.3. Phosphating<br />

3.5.4. Rinsing after phosphating<br />

3.5.5. Chromic acid sealing<br />

3.5.6. Drying<br />

3.6. Coating characteristics<br />

3.6.1. Structure and composition<br />

3.6.2. Coating thickness and coating weight<br />

3.6.3. Coating porosity<br />

3.6.4. Stability of the <strong>phosphate</strong> coating<br />

3.7. Influence of various factors on coating<br />

properties<br />

3.7.1. Nature of the substrate<br />

3.7.1.1. Composition of the metal<br />

3.7.1.2. Structure of the metal <strong>surface</strong><br />

3.7.1.3. Surface preparation<br />

3.7.1.4 Surface activation


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

3.7.1.5. Thermal treatments and machining<br />

3.7.2. Phosphating parameters<br />

3.8. Processing problems and remedial<br />

measures<br />

3.9. Defects in <strong>coatings</strong> and remedies<br />

3.10. Characterization of <strong>phosphate</strong><br />

<strong>coatings</strong><br />

3.11. Testing the quality of <strong>phosphate</strong><br />

<strong>coatings</strong><br />

3.11.1. Evaluation of physical<br />

characteristics<br />

3.11.2. Evaluation of corrosion performance<br />

3.12. Applications<br />

3.13. Environmental impact<br />

4. Summary<br />

References<br />

1. INTRODUCTION<br />

Metals have been the backbone of civilization. Efforts<br />

have been spared to find alternatives and replacements<br />

for metals but these still play a major<br />

role in the manufacture and construction and are<br />

likely to do so for many more years. This is due to<br />

the combination of several useful properties like<br />

strength, workability, low-cost and ability to be recycled,<br />

that the metals possess. However, metals<br />

which are extracted from their ores <strong>by</strong> chemical or<br />

electrochemical means show a strong tendency to<br />

revert to their oxide form at the first available opportunity,<br />

i.e., they tend to corrode [1-4] and as a result<br />

they create a tremendous economic loss besides<br />

posing a serious threat to the national resources<br />

of a country.<br />

The methods of corrosion prevention are many<br />

and varied. These methods may be generally classified<br />

[3] as:<br />

• Modification of the metal <strong>by</strong> alloying and/or <strong>surface</strong><br />

modification;<br />

• Modification of the environment <strong>by</strong> the use of inhibitors;<br />

and<br />

• Change of metal/environment potential <strong>by</strong> cathodic<br />

or anodic protection.<br />

The most commonly used method of corrosion<br />

protection involves bulk alloying or <strong>surface</strong> modification.<br />

Surface modification is however, far more<br />

economical than bulk alloying and is more widely<br />

practiced. The methods generally used for <strong>surface</strong><br />

modification involve the formation of a physical barrier<br />

to protect the metal against its corrosive environment<br />

[5]. This can be achieved <strong>by</strong> relatively more<br />

modern methods such as: (i) physical vapour deposition<br />

(PVD); (ii) chemical vapour deposition (CVD);<br />

131<br />

(iii) ion implantation; (iv) laser treatment; (v) deposition<br />

<strong>by</strong> thermal spray, plasma spray and arc methods;<br />

(vi) nitriding; (vii) carbiding; etc., or through more<br />

conventional techniques such as: (i) painting; (ii)<br />

anodizing; and (iii) chemical <strong>conversion</strong> <strong>coatings</strong>.<br />

While the former methods are usually less economic<br />

as they involve the use of sophisticated application<br />

techniques and are meant for specialized applications,<br />

the latter methods are more cost-effective and<br />

have a wider spectrum of end applications.<br />

2. CHEMICAL CONVERSION<br />

COATINGS<br />

Chemical <strong>conversion</strong> <strong>coatings</strong> are adherent, insoluble,<br />

inorganic crystalline or amorphous <strong>surface</strong><br />

films, formed as an integral part of the metal <strong>surface</strong><br />

<strong>by</strong> means of a non-electrolytic chemical reaction<br />

between the metal <strong>surface</strong> and the dipped in<br />

solution [6]. In such <strong>coatings</strong>, a portion of the base<br />

metal is converted into one of the components of<br />

the resultant protective film, which is much less reactive<br />

to subsequent corrosion than the original metal<br />

<strong>surface</strong>. This film imparts an equal potential to the<br />

metal <strong>surface</strong>, neutralizing the potential of the local<br />

anodic and cathodic galvanic corrosion sites [7].<br />

They also serve as absorptive bases for improving<br />

the adhesion to paints and other organic finishes.<br />

Chemical <strong>conversion</strong> <strong>coatings</strong> are preferred because<br />

of their adherent nature and high speed of coating<br />

formation besides being economical. Further these<br />

can be formed using simple equipment and without<br />

the application of any external potential. Chemical<br />

<strong>conversion</strong> coating processes are classified as phosphating,<br />

chromating and oxalating according to their<br />

essential constituents viz., <strong>phosphate</strong>s, chromates,<br />

and oxalates respectively [8]. The present <strong>review</strong><br />

focuses on <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> with a<br />

special emphasize on zinc <strong>phosphate</strong> <strong>coatings</strong> on<br />

mild steel.<br />

3. PHOSPHATING<br />

Phosphating process can be defined as the treatment<br />

of a metal <strong>surface</strong> so as to give a reasonably<br />

hard, electrically non-conducting <strong>surface</strong> coating of<br />

insoluble <strong>phosphate</strong> which is contiguous and highly<br />

adherent to the underlying metal and is considerably<br />

more absorptive than the metal [9]. The coating<br />

is formed as a result of a topochemical reaction,<br />

which causes the <strong>surface</strong> of the base metal to<br />

integrate itself as a part of the corrosion resistant<br />

film.


132 T.S.N. Sankara Narayanan<br />

Table 1. Historical development of the phosphating process.<br />

Sl. Year/ Advancement made/process developed References<br />

No. Period<br />

1. 1906 Phosphating of iron and steel using phosphoric acid and iron filings 11<br />

2. 1908 Treatment of <strong>phosphate</strong> <strong>coatings</strong> with oxidizing agents to reduce 12<br />

process time<br />

3. 1909 Regeneration of the bath and formulation of zinc <strong>phosphate</strong> baths 13,14<br />

requiring high temperature <strong>–</strong> process time of one hour<br />

4. 1911 Formulation of manganese <strong>phosphate</strong> bath requiring high temperature - 15<br />

process time of 2-2.5 hours<br />

5. 1914 Parkerising process with maintenance of total acid to free acid ratio 16,17<br />

6. 1928 Recognition of <strong>phosphate</strong> coating as paint base 18,19<br />

7. 1929 Bonderizing process with the addition of Copper accelerator-coating 20<br />

time: 10 minutes to 1 hour<br />

8. 1933 Use of oxidizing agents like nitrate for acceleration <strong>–</strong> coating time: 21<br />

5 minutes<br />

9. 1934 Use of <strong>phosphate</strong> coating for cold working operations for metals 22<br />

10. 1937 Spray phosphating <strong>–</strong> phospating time: 60-90 seconds 23<br />

11. 1940 Development of non-coating <strong>phosphate</strong> process based on sodium 24<br />

or ammonium <strong>phosphate</strong>s<br />

12. 1940 Development of cold phosphating methods 25<br />

13. 1941 Phosphating of aluminium <strong>surface</strong>s using zinc <strong>phosphate</strong> and fluorides 26<br />

14. 1943 Use of disodium <strong>phosphate</strong> containing titanium as pre-dip before 27<br />

phosphating<br />

15. 1950’s Large scale application of manganese <strong>phosphate</strong> <strong>coatings</strong> as an oil 28<br />

retaining medium <strong>–</strong> for use on bearing or sliding <strong>surface</strong>s etc.<br />

16. 1960’s Use of special additives to control coating weight 29<br />

17. 1960’s Spray process at operating temperature of 25-30 °C 18,19<br />

18. 1970’s Improvement in coating quality, use of spray cleaners based on 18,19<br />

surfactant technology<br />

3.1. History and development of the<br />

phosphating process<br />

The use of <strong>phosphate</strong> <strong>coatings</strong> for protecting steel<br />

<strong>surface</strong>s has been known since the turn of the century<br />

and during this period the greater part of the<br />

World’s production of cars, refrigerators and furniture<br />

were treated this way. The first reliable record<br />

of <strong>phosphate</strong> <strong>coatings</strong> applied to prevent rusting of<br />

iron and steel is a British patent of 1869 granted to<br />

Ross [10]. In the method used <strong>by</strong> him, red hot iron<br />

articles were plunged into the phosphoric acid to<br />

prevent them from rusting. Since then numerous<br />

developments have taken place, of which the major<br />

developments are listed in Table 1.<br />

During the last 30 years, work has been concentrated<br />

mainly on improvements in quality, particularly<br />

to keep in pace with the recent changing<br />

needs of the organic finishing systems. Prominent<br />

among these are: (i) use of low temperature phosphating<br />

baths to overcome the energy crisis [30-<br />

32]; (ii) use of low zinc technology [18,19]; (iii) use<br />

of special additives in the phosphating bath [33-42];<br />

(iv) use of more than one heavy metal ions in existing<br />

composition-particularly tri-cation phosphating<br />

[43]; etc. New types of <strong>phosphate</strong> <strong>coatings</strong> such<br />

as tin, nickel and lead <strong>phosphate</strong> <strong>coatings</strong> have been<br />

introduced [44,45] besides the development of compositions<br />

for simultaneous phosphating of multiple<br />

metal substrates [46,47]. There has been a grow-


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 2. Special additives used in phosphating baths.<br />

Sl. Additive used Purpose Impact References<br />

No.<br />

1. α-hydroxy carboxylic acids To reduce the coating Improve bath life through 49-54<br />

like tartaric and citric acids, weight lesser consumption of<br />

tripoly<strong>phosphate</strong>, sodium, chemicals<br />

potassium tartrate,<br />

nitrobenzene sulphonate<br />

2. Chelants such as EDTA, NTA, To increase the coating Improved corrosion 42, 43<br />

DTPA gluconic acids and weight protection, shorter 55-58<br />

polycarboxy o-amino acids. processing times.<br />

3. Quaternary ammonium Grain refinement Better adhesion of 59-63<br />

compounds, N and P subsequent finishes;<br />

compound containing better corrosion<br />

amido or amino group. protection.<br />

Calcium, Formic acid ester,<br />

chelate of an acidic organic<br />

<strong>phosphate</strong><br />

4. Nickel (II) To improve <strong>surface</strong> Better adhesion of 64, 65<br />

texture subsequent finishes;<br />

better corrosion<br />

protection.<br />

5. Lead compounds, tungstate To accelerate the Reduction in 66-69<br />

ions, gaseous nitrogen phosphating process processing time<br />

peroxide, hydroxylamine<br />

sulphate, hexamine<br />

6. Persulphate and To prevent concentration Better utilization of 70<br />

permonosulphuric acid of ferro-nitroso complex nitrite and reduction<br />

in nitrite accelerated in the evolution of<br />

zinc phosphating bath. toxic vapours<br />

7. Cyclic trimeta<strong>phosphate</strong> To reduce the operating Thinner, smoother and 71<br />

temperature improved corrosion<br />

To increase the resistant coating<br />

tolerance to dissolved<br />

iron<br />

8. Lauric, Palmitic, and To improve lubricant Improved workability 72<br />

Stearic acids with fatty properties of the metal<br />

amines and ethoxylated<br />

amines<br />

9. Carbohydrates, dialkyl- To decrease scaling Improve the service 73-75<br />

triaminepentakis life of heating coils<br />

methylene phosphonic and provide uniform<br />

acid and its salts and heating of the bath<br />

fluoborate or fluosilicate<br />

10. Phosphonic acid ester, To prevent the build up Improved service life 76, 77<br />

Magnesium or zinc nitrate of sludge on tank walls of the equipment<br />

11. Amines, Tin (IV), Arsenic As inhibitors Improves corrosion 78, 79<br />

compounds, zinc salt of resistance<br />

an organic N-compound<br />

12. Zinc carbonate To reduce the acidity Consistent performance 80<br />

of the bath and to of the bath<br />

maintain the equilibrium<br />

133


134 T.S.N. Sankara Narayanan<br />

Table 2. Continued.<br />

13. Thiourea Stabilizer in non- Improved stability of 81<br />

aqueous phosphating the bath<br />

solutions, as inhibitors<br />

14 Sodium lignosulphonate To modify the physical Reduce the tendency 82, 83<br />

form of sludge to form a crust on heat<br />

transfer <strong>surface</strong>s<br />

Improves process<br />

efficiency<br />

15 Methylaminoethoxysilane To prevent rehydration Impart hydrophobic and 84<br />

of <strong>phosphate</strong> dihydrate anti-corrosion properties<br />

16 Hexameta<strong>phosphate</strong> To decrease the <strong>surface</strong> Enables deep drawing 85<br />

roughness and increase of tubes (~50% length)<br />

the extent of absorption with reduction in wall<br />

of sodium sterate thickness<br />

ing use of substitutes to conventional Cr(VI) postrinses<br />

[48] to suit the regulations imposed <strong>by</strong> the<br />

pollution control authorities on the use of Cr(VI) compounds.<br />

The special additives used in phosphating<br />

baths is complied in Table 2 and the alternatives to<br />

Cr(VI) post-rinse treatment is given in Table 3.<br />

3.2. Chemistry of phosphating<br />

All conventional phosphating solutions are dilute<br />

phosphoric acid based solutions of one or more alkali<br />

metal/heavy metal ions, which essentially contain<br />

free phosphoric acid and primary <strong>phosphate</strong>s<br />

of the metal ions contained in the bath<br />

[18,19,24,127,128]. When a steel panel is introduced<br />

into the phosphating solution a topochemical<br />

reaction takes place in which the iron dissolution is<br />

initiated at the microanodes present on the substrate<br />

<strong>by</strong> the free phosphoric acid present in the<br />

bath. Hydrogen evolution occurs at the<br />

microcathodic sites.<br />

Fe + 2H 3 PO 4 → Fe(H 2 PO 4 ) 2 + H 2 ↑ . (1)<br />

The formation of soluble primary ferrous <strong>phosphate</strong><br />

leads to a concurrent local depletion of free acid<br />

concentration in the solution resulting in a rise in<br />

pH at the metal/solution interface. This change in<br />

pH alters the hydrolytic equilibrium which exists<br />

between the soluble primary <strong>phosphate</strong>s and the<br />

insoluble tertiary <strong>phosphate</strong>s of the heavy metal ions<br />

present in the phosphating solution, resulting in the<br />

rapid <strong>conversion</strong> and deposition of insoluble heavy<br />

metal tertiary <strong>phosphate</strong>s [18,19,24,127, 128]. In a<br />

zinc phosphating bath these equilibria may be represented<br />

as:<br />

Zn(H 2 PO 4 ) 2 ⇔ ZnHPO 4 + H 3 PO 4 , (2)<br />

3ZnHPO 4 ⇔ Zn 3 (PO 4 ) 2 + H 3 PO 4 . (3)<br />

A certain amount of free phosphoric acid must<br />

be present to repress the hydrolysis and to keep<br />

the bath stable for effective deposition of <strong>phosphate</strong><br />

at the microcathodic sites. Another factor affecting<br />

the shift in the primary to tertiary <strong>phosphate</strong> equilibria<br />

is the temperature of the bath. Higher temperatures<br />

favour easy precipitation of the tertiary <strong>phosphate</strong>s<br />

in a shorter time. Hence, more amount of<br />

phosphoric acid is needed for the baths operating<br />

at higher temperatures. In contrast, in the case of<br />

phosphating baths operated at room temperature,<br />

the possibility of the increase in acidity during continuous<br />

operation is more likely [129,130] and is<br />

normally neutralised <strong>by</strong> the addition of the carbonate<br />

of the metal which forms the coating (Zn(CO 3 ) 2<br />

in zinc phosphating bath). Hence, depending upon<br />

the working temperatures and the concentrations<br />

of the constituents in the bath, the free phosphoric<br />

acid content must be chosen to maintain the equilibrium<br />

condition. Too much of phosphoric acid not<br />

only delays the formation of the coating, but also<br />

leads to excessive metal loss.<br />

3.3. Acceleration of the phosphating<br />

process<br />

In practice, phosphating reaction tends to be slow<br />

owing to the polarization caused <strong>by</strong> the hydrogen<br />

evolved in the cathodic reaction. In order to achieve<br />

coating formation in a practicable time, some mode<br />

of acceleration must be employed. The importance<br />

of the acceleration of the phosphating process was


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 3. Alternatives to Cr(VI) post treatment.<br />

Sl.No. Type of compounds Compound used Reference<br />

1. Chromium containing a. 0.001% Cr(III) as Cr- chromate complex 86, 87<br />

substitutes b.CrO solution reduced <strong>by</strong> HCHO 3<br />

2- c. Cr O as complex of divalent metal 2 7<br />

d. Chromate and colloidal silicic acid<br />

88<br />

89<br />

90<br />

e. Aqeuous solution of Aluminium chromate polymer 91, 92<br />

f. Trivalent chromium and ferricyanide 93<br />

2. Phosphonic acid a. Alkane phosphonic acids 94, 95<br />

derivatives b. Alkene phosphonic acids 96, 97<br />

c. Polyvinyl phosphonic acids 98<br />

3. Ammonium <strong>phosphate</strong> a. Dilute solution of ammonium primary <strong>phosphate</strong> 99, 100<br />

derivatives and/or trietha nolamine ydrogen <strong>phosphate</strong><br />

4. Carboxylic acids a. Citric, glutaric, maleic, succinic and phthalic acids 101<br />

b. Citrate in combination with nitrite 102, 103<br />

5. Oxidizing agents a. Peroxides and persulphates 104<br />

b. Potassium permanganate 105<br />

6. Polymers a. Modified hydroxystyrene 106<br />

b. Melamine-formaldehyde 107<br />

7. Tannins a. Tannin + thiourea compound 108<br />

b. Tannin + melamine <strong>–</strong> formaldehyde 109<br />

8. Tin salts Aqueous stannous salt solution containing Mn, Pb,<br />

Cd, Co and Ni<br />

110-113<br />

9. Zirconium and Titanium a. Ti (III) in acid solution 114, 115<br />

compounds b. Zr or Ti with an inositol <strong>phosphate</strong> ester 116<br />

c. Zr + myo-inositol <strong>phosphate</strong> and/or its salt 117<br />

10. Molybdenum type<br />

rinses<br />

a. Dilute aqueous solution of MoO in HNO 2 3 118<br />

11. Hypophosphorus acid a. Hypophosphorus acid + Hydrofluorosilicic acid 119<br />

and hypophosphite b. Sodium hypophosphite + Hydrofluorosilicic acid<br />

12. Mixture of organo- a. Carboxyethylene phosphonic acid esters of butyl 120<br />

phophates or<br />

phosphonates and<br />

fluoride<br />

diglycidyl ether and bisphenol + fluorosilicic acid<br />

13 Mixture of Zr, V, F a. Hydrofluorozirconic acid + fluoroboric acid + 121<br />

and <strong>phosphate</strong> ions Phosphoric acid + amm. metavanadate<br />

14 Mixture of Li, Cu and<br />

Ag ions<br />

a. Nitrate (or) carboxylate (or) acetates of Li, Cu and Ag 122, 123<br />

15 Mixture of Ti, V, Mo, a. As nitrates, sulphates (or) chloride 124<br />

Ag, Sn, Sb + cathodic<br />

treatment at 0.1 to<br />

10 A/dm<br />

b. Salt of hydroxycarboxylic acid<br />

2<br />

16 Tetravalent Ti (or) a. Titanyl acetylacetonate 125<br />

Divalent Mn, Co, Ni b. Divalent manganese/cobalt/nickel/copper/ethanote<br />

or Cu c. Divalent<br />

17 Ni + Co + Sn + Pb a. As acetates + Pb sheets + stannous chloride 126<br />

135


136 T.S.N. Sankara Narayanan<br />

felt way back in the 19th century and its development<br />

has gained a rapid momentum with the advent<br />

of the Bonderite process in 1929. Recently,<br />

Sankara Narayanan et al. [131] have made an overview<br />

on the acceleration of the phosphating process<br />

and justified its role on the hunting demand to reduce<br />

process time. The different means of accelerating<br />

the formation of <strong>phosphate</strong> <strong>coatings</strong> can be<br />

broadly classified as: (i) Chemical acceleration; (ii)<br />

Mechanical acceleration; and (iii) Electrochemical<br />

acceleration.<br />

3.3.1. Chemical acceleration<br />

Oxidizing substances [132,133] and metals more<br />

noble than iron such as, Cu, Ni, [19] etc., constitute<br />

the most important class of chemical accelerators.<br />

They accelerate the deposition process<br />

through different mechanisms. Oxidizing agents<br />

depolarize the cathode half-cell reaction <strong>by</strong> preventing<br />

the accumulation of hydrogen at the cathodic<br />

sites, whereas noble metal ions promote metal dissolution<br />

<strong>by</strong> providing low over-potential cathode sites<br />

<strong>by</strong> their deposition [134]. Since acceleration through<br />

depolarization is preferred to mere promotion of<br />

metal dissolution, oxidizing agents have found widespread<br />

use than metals. Moreover, they prevent the<br />

excessive build up of iron in the bath, which can be<br />

detrimental to good coating formation [25]. The most<br />

commonly employed oxidizing accelerators are nitrites,<br />

chlorates, nitrates, peroxides and organic nitro<br />

compounds either alone or in various combinations.<br />

Common combinations are nitrite-nitrate, nitrite-chlorate-nitrate<br />

and chlorate-nitrobenzene sulphonic<br />

acid. The characteristics of some of the commonly<br />

used oxidizing accelerators are given in Table 4.<br />

Some reducing agents such as alkali metal sulphites<br />

[135], hypophosphites [136], phosphites [137], formaldehyde,<br />

benzaldehyde, hydroxylamine [138], acetaldehyde<br />

oxime [139], Pyridine N-oxime [140],<br />

morpholine N-oxime [140], quinones [141], etc., have<br />

also been tried as accelerators but have not been<br />

as successful as oxidizing accelerators from the<br />

industrial point of view.<br />

3.3.2. Mechanical acceleration<br />

When a phosphating solution is sprayed forcibly on<br />

to a metal <strong>surface</strong>, <strong>coatings</strong> are formed more readily<br />

than would be <strong>by</strong> immersion in the same solution,<br />

since the former process eliminates the delay due<br />

to the diffusion of the constituents in the solution to<br />

the metal <strong>surface</strong>. The comparative kinetics of spray<br />

and dip phosphating was determined <strong>by</strong> Laukonis<br />

[142]. The resultant <strong>coatings</strong> are thin, fine-crystalline<br />

and perfectly suitable as a paint base. Other<br />

means of physical acceleration are the action of<br />

brushes and rollers [143] on the <strong>surface</strong> during processing.<br />

3.3.3. Electrochemical acceleration<br />

Several electrochemical methods of acceleration,<br />

both anodic, cathodic and pulse method have been<br />

described in literature [144-161]. Coslett recognized<br />

acceleration of the phosphating process <strong>by</strong> cathodic<br />

treatment as early as 1909 <strong>by</strong> Coslett [144]. Subsequent<br />

studies reveal that anodic methods are<br />

more appropriate and advantageous than the cathodic<br />

methods, as they promote metal dissolution<br />

as well as passivity. Zantout and Gabe [148] claim<br />

to have achieved higher coating weights of low porosity<br />

in a shorter treatment time <strong>by</strong> the application<br />

of a small current. Ravichandran et al., [157-159]<br />

established the utility of galvanic coupling of steel<br />

substrate with metals which are more noble than<br />

steel for accelerating phosphating processes. This<br />

methodology employs the galvanic corrosion principle<br />

for accelerating the metal dissolution reaction,<br />

which enables a quicker consumption of free phosphoric<br />

acid and an earlier attainment of the point of<br />

incipient precipitation, resulting in higher amount of<br />

<strong>phosphate</strong> coating formation. The application of a<br />

cathodic current to form <strong>phosphate</strong> coating on stainless<br />

steel using a calcium modified zinc phosphating<br />

bath was patented <strong>by</strong> Bjerrum et al. [160]. Sinha<br />

and Feser [161] have also studied the formation of<br />

<strong>phosphate</strong> coating on steel and stainless steel substrates<br />

<strong>by</strong> this method.<br />

These three methods of acceleration described<br />

above are widely practiced in industries; but<br />

each of them has its own merits and demerits.<br />

Though chemical accelerators (Primarily those of<br />

the oxidizing type) accelerate the phosphating process<br />

<strong>by</strong> their mere addition, their concentration in<br />

the bath is very critical to yield the desired results.<br />

Acceleration <strong>by</strong> mechanical means is limited to<br />

spray processes, which are capable of providing<br />

fresh bath solution constantly. Electrochemical<br />

methods of acceleration though capable of yielding<br />

higher deposition rate, the practical difficulty in adding<br />

‘electrics’ to the processing stage makes it less<br />

popular.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 4. Different types of accelerators used in phosphating.<br />

Type of Example Effective Advantages Disadvantages<br />

accelerator concentration<br />

- NO NaNO 1-3% Lower sludge Reduction of FePO increases<br />

3<br />

3 4<br />

Zn(NO ) the iron content of the coating<br />

3 2<br />

Ni(NO ) 3 2<br />

NO 2<br />

ClO 3<br />

- NaNO 2 0.1-0.2 g/l Affords rapid processing Corrosive fumes.<br />

137<br />

even at low temperatures Highly unstable at high bath<br />

temperatures.<br />

Frequent addition is required.<br />

- Zn(ClO 3 ) 2 0.5 - 1% Stable in liquid Corrosive nature of chlorate and<br />

concentrates. Can be its reduction products.<br />

used for bath make-up High concentrations poison<br />

and replenishment. the bath. Removal of gelati<br />

Overcomes the white nous precipitate from the<br />

staining problem. resultant <strong>phosphate</strong> <strong>coatings</strong> is<br />

difficult.<br />

H 2 O 2 H 2 O 2 0.05 g/l Low coating weight. Bath control tends to be<br />

(liquid) No harmful products. critical. Heavy sludge<br />

Free from staining formation. Limited stability.<br />

Continuous addition is required.<br />

Perborate Sodium <strong>–</strong> No separate neutraliser Continuous addition is<br />

perborate is required. required. Voluminous sludge.<br />

Good corrosion<br />

resistance<br />

Nitroguanidine Nitroguanidine <strong>–</strong> Neither the accelerator Slightly soluble.<br />

nor its reduction No control for build-up<br />

products are corrosive. of ferrous iron in the bath.<br />

Highly expensive.<br />

Oxime Acetaldehyde 1-5 g/l Stable in acidic <strong>–</strong><br />

oxime environment<br />

Environmentally<br />

acceptable<br />

Organic Pyridine 0.3-2 g/l <strong>–</strong> Need to use highly<br />

N-oxide N-oxide concentrated activating<br />

Morpholine solution before<br />

N-oxide phosphating<br />

Hydroxylamine Hydroxylamine 0.5-5 g/l Spherical and/or Decomposes in presence of Cu<br />

Hydroxylamine columnar <strong>phosphate</strong> or H 2 O 2 in the bath<br />

sulphate crystal<br />

Does not decompose<br />

on its own<br />

Quinones Benzoquinone <strong>–</strong> Solubility of Quinone High concentration causes a<br />

Chloranil wavy pattern. Lower concentra<br />

tion not effective in acceleration<br />

Amido Benzoicacid 0.1-6 g/l Immediate operation Effectiveness is impaired in<br />

sulphonic acid sulphamide of the bath presence of Ca ions in the<br />

and their Benzenesulphoanilide Fine grained coating bath<br />

N-substituted N-cyclohexyl Improved lacquer<br />

derivates sulphamic acid adhesion and corrosion<br />

resistance. Sludge<br />

formation is suppressed<br />

Alkali metal Sodium bromate 0.8-1.1 + Fine grained coating <strong>–</strong><br />

bromate + +nitrocompound 0.25-0.5 Improved adhesion and<br />

Aromatic m-nitro benzene- g/l corrosion resistance<br />

sulphonate


138 T.S.N. Sankara Narayanan<br />

3.4. Kinetics of the phosphating<br />

process<br />

Kinetics of the phosphating process reveals the<br />

steps involved in the course of phosphating and their<br />

rates. Three different methods have been used so<br />

far for investigating the kinetics of formation of <strong>phosphate</strong><br />

<strong>coatings</strong> namely: (i) the gravimetric method,<br />

<strong>by</strong> the quantitative determination of the quantity of<br />

<strong>phosphate</strong> deposited per unit time; (ii) the electrochemical<br />

method based on the determination of free,<br />

reactive uncoated areas through electrochemical<br />

passivation; and (iii) the radiographic method based<br />

on the determination of the intensity of the characteristic<br />

X-ray of the resulting compound.<br />

All the three methods gave a similar picture of<br />

the phosphating process that showed the coating<br />

formation did not take place in a linear fashion, rather<br />

it was initially very fast, after which the rate slowly<br />

decreased with time.<br />

Studies on the kinetics of phosphating indicate<br />

that there are four distinct stages in coating formation<br />

(Fig. 1), namely, the induction stage (α), the<br />

commencement of film growth (β), the main exponential<br />

growth stage (γ) and the stage of linear increase<br />

in film growth (δ). During the induction period,<br />

the oxide film remaining on the <strong>surface</strong> even<br />

after cleaning is removed. When film growth commences,<br />

the first nuclei are formed and the rate of<br />

nucleation increases rapidly with time. This, however,<br />

depends considerably on the conditions of the<br />

<strong>surface</strong>, the <strong>pretreatment</strong> procedures adopted and<br />

the oxidizing agents present in the phosphating bath.<br />

Growth occurs in the main exponential growth stage.<br />

Addition of accelerators reduces the induction period<br />

and extends the stage of linear growth.<br />

In the opinion of Gebhardt [162], the rate of phosphating<br />

depends on the rate of diffusion of Fe 2+ ions<br />

from the structural lattice to the coating/solution<br />

interface through the coating formed. Machu [25]<br />

has found that the rate of the phosphating reaction<br />

is a function of microanodes on the <strong>surface</strong>.<br />

-dF A /dt = K . F A dt<br />

where dt <strong>–</strong> change in time; F A <strong>–</strong> <strong>surface</strong> area of anodes<br />

in microcells; and K <strong>–</strong> reaction rate constant.<br />

The rate of formation of the <strong>phosphate</strong> coating<br />

depends primarily upon the metal viz., the ratio of<br />

anode <strong>surface</strong> which was initially present, F Ao , to<br />

the anode <strong>surface</strong> at any given moment, F A . The<br />

influence of various other factors controlling the rate<br />

of reaction, e.g., temperature, <strong>surface</strong> condition etc.,<br />

is reflected in the value of the rate constant K, which<br />

is different for different processes.<br />

Fig. 1. The various stages in the growth of <strong>phosphate</strong><br />

coating (in presence of accelerator).<br />

The overall coating growth process can be followed<br />

<strong>by</strong> potential-time curves (Fig. 2), which indicate<br />

the different stages of coating besides indicating<br />

when the effective phosphating has ceased.<br />

Correlation of potential-time relationships with film<br />

properties leads to the conclusion that coating formation<br />

proceeds through the following stages:<br />

(a) electrochemical attack of steel;<br />

(b) amorphous precipitation;<br />

(c) dissolution of the base metal;<br />

(d) crystallization and growth; and<br />

(e) crystal reorganization.<br />

But, in actual practice, it is difficult to identify the ‘b’<br />

and ‘c’ stages, and the curve mainly exhibits the<br />

process of metal dissolution, coating formation and<br />

coating completion. The use of potential-time measurements<br />

for monitoring the phosphating process<br />

was first described <strong>by</strong> Machu [25]. It has also been<br />

used <strong>by</strong> several workers to reveal the nature of the<br />

Fig. 2. Potential-time curve showing the various<br />

steps of phosphating.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

process taking place during phosphating [145] and<br />

the properties of the coating formed [163]. Sankara<br />

Narayanan et al. [164,165] have discussed the usefulness<br />

of the potential-time curves in predicting the<br />

kinetics of the phosphating process. The utility of<br />

potential-time measurement for effective on-line<br />

monitoring of the phosphating process is also established<br />

[166,167]. The correlation between coating<br />

weight and potential time measurements is established<br />

<strong>by</strong> Sankara Narayanan [168-170], which<br />

enables calculation of crystallization kinetics from<br />

potential time measurement.<br />

3.5. Process details<br />

In general, phosphating sequence comprises of seven<br />

operations, as indicated in the flow chart. However,<br />

depending upon the <strong>surface</strong> conditions of the base<br />

metal, some of these operations may be omitted or<br />

additional operations may be incorporated into the<br />

system. A typical seven-stage sequence of <strong>phosphate</strong><br />

<strong>pretreatment</strong> process is given in Fig. 3.<br />

3.5.1. Cleaning<br />

Perhaps, the most important requisite for proper<br />

coating formation is a clean substrate, free from<br />

contaminants such as oils, greases, waxes, corrosion<br />

products and other soils. Many coating failures<br />

can be attributed to the poor metal <strong>surface</strong><br />

preparation [171]. An ideal cleaning agent is the<br />

one, which is capable of removing all the contaminants<br />

from the metal <strong>surface</strong>, and prevents their redeposition<br />

or the formation of other detrimental reaction<br />

products [172]. A variety of methods such<br />

as sand blasting, solvent degreasing, vapour<br />

degreasing, alkaline cleaning and pickling have been<br />

used to achieve this end.<br />

Sand blasting is an effective method of mechanical<br />

cleaning. However, it is highly expensive and its<br />

use is justified as a field procedure where chemical<br />

treatments cannot be used and it is necessary to<br />

remove the loosened mill scale as well as paint [173].<br />

Organic solvents are widely used to remove organic<br />

contaminants from the metal substrates. But<br />

they are of toxic and flammable nature and need to<br />

be used in large quantities, which is uneconomical.<br />

This has led to its replacement <strong>by</strong> the vapour<br />

degreasing technique. The unique advantage of the<br />

latter technique over solvent degreasing is that, continuous<br />

cleaning with small quantities of solvent is<br />

possible [174].<br />

Alkaline cleaning provides an economical and<br />

effective alternative to the use of organic solvents to<br />

Degreasing<br />

pickling<br />

Rinsing<br />

Phosphating<br />

Rinsing<br />

Chromic acid sealing<br />

Drying<br />

139<br />

Fig. 3. Flow chart depicting the operating sequence<br />

involved in phosphating process.<br />

remove greases, oils and waxes. They are also used<br />

in conjunction with <strong>surface</strong> active (wetting) agents<br />

and emulsified hydrocarbon solvents [174]. Alkaline<br />

cleaners are particularly efficient when used hot<br />

(approx. 79 °C). While alkaline cleaning is free from<br />

the fire and toxicity hazards associated with organic<br />

solvent cleaning (unless emulsified solvents have<br />

been incorporated), the corrosive effects of alkaline<br />

materials on the skin and on ordinary clothing must<br />

be guarded against. Caustic soda in particular can<br />

cause serious burns to the skin and eyes and is<br />

extremely irritating to the nasal and bronchial membranes<br />

if inhaled.<br />

Acid cleaning or pickling using acids such as<br />

HCl, H 2 SO 4 , and H 3 PO 4 is a very effective method


140 T.S.N. Sankara Narayanan<br />

for the removal of rust and mill scale [175]. Dilute<br />

solutions (5-10% <strong>by</strong> weight) of H 2 SO 4 and HCl are<br />

used in presence of inhibitors to remove the inorganic<br />

contaminants <strong>by</strong> converting them into their<br />

ferrous salts. Pickling in H 2 SO 4 is usually performed<br />

at high temperatures (about 60 °C). H 3 PO 4 is an<br />

excellent time-tested cleaning agent which not only<br />

removes organic and inorganic solids present on<br />

the metal but also causes chemical etching of the<br />

<strong>surface</strong> <strong>by</strong> reacting with it to produce a mechanically<br />

and chemically receptive <strong>surface</strong> for subsequent<br />

coating formation [176].<br />

Electrolytic pickling is an alternative to chemical<br />

pickling, which provides better and rapid cleaning<br />

through an increased hydrogen evolution, resulting<br />

in greater agitation and blasting action [174].<br />

3.5.2. Rinsing<br />

The rinsing step followed <strong>by</strong> cleaning plays a vital<br />

role in the phosphating sequence [172]. Rinsing<br />

prevents the dragout of chemicals used in the earlier<br />

cleaning that may contaminate the subsequent<br />

stages.<br />

3.5.3. Phosphating<br />

Suitably cleaned <strong>surface</strong>s are next subjected to<br />

phosphating, which causes the formation of an insoluble,<br />

corrosion resistant <strong>phosphate</strong> layer on the<br />

substrate <strong>surface</strong>. A wide variety of phosphating<br />

compositions are available. However, the right choice<br />

of the components and the operating conditions of<br />

the phosphating bath are made based on the nature<br />

of the material to be treated and its end use.<br />

All the phosphating compositions are essentially<br />

dilute phosphoric acid based solutions containing<br />

alkali metal/heavy metal ions in them besides suitable<br />

accelerators [18,19,24,127,128]. Based on the<br />

nature of the metal ion constituting the major component<br />

of the phosphating solution, these compositions<br />

are classified as zinc, manganese and iron<br />

phosphating baths. The characteristics of the <strong>coatings</strong><br />

obtained using these baths are presented in<br />

Table 5.<br />

Phosphating can be effectively performed on both<br />

ferrous and non-ferrous metals. Among the ferrous<br />

metals, mild steels are most frequently used although<br />

maraging steels, galvanized steels and stainless<br />

steels can also be coated [177-180]. Non-ferrous<br />

metals that can be <strong>phosphate</strong>d include zinc,<br />

aluminium, magnesium and cadmium [181-183].<br />

Physical properties like hardness, tensile strength<br />

and workability of the original metal are retained af-<br />

ter phosphating [21]. The dimensional change<br />

caused <strong>by</strong> <strong>phosphate</strong> <strong>coatings</strong> on the metal <strong>surface</strong><br />

is of the order of 10 -3 mm.<br />

Phosphate deposition can be achieved through<br />

the use of both spray and immersion processes and<br />

the choice of the appropriate method depends upon<br />

the size and shape of the substrate to be coated<br />

and based on the end use for which the coating is<br />

made. Spray process is preferred where shorter processing<br />

times are required. This method, however,<br />

requires more factory floor space and special equipment<br />

for their application. Immersion process though<br />

slower, produce uniform <strong>coatings</strong> and they require<br />

less factory floor space as the process tanks can<br />

be arranged in a compact manner. The benefits of<br />

phosphating <strong>by</strong> total immersion were considered <strong>by</strong><br />

Wyvill [184]. But, immersion processes are more<br />

susceptible to contamination during continuous<br />

operation than are spray processes. Smaller parts<br />

can be effectively and economically <strong>phosphate</strong>d <strong>by</strong><br />

immersion process whereas spray process is more<br />

suitable for larger work pieces. Nowadays, a combination<br />

of both spray and immersion process has<br />

been successfully used particularly in automobile<br />

industries [185].<br />

Phosphating may be carried out at temperatures<br />

ranging from 30-99 °C and processing time can be<br />

varied from a few seconds to several minutes. Suitable<br />

choice of these parameters is determined <strong>by</strong><br />

factors such as nature of the metal to be coated,<br />

thickness and weight of the coating required and<br />

bath composition. The process of phosphating involves<br />

a consistent depletion of bath constituents<br />

and in order to obtain a satisfactory <strong>phosphate</strong> coating,<br />

the bath parameters such as: (i) the free acid<br />

value (FA) which refers to the free H + ions present in<br />

the phosphating solution; (ii) total acid value (TA)<br />

which represents the total <strong>phosphate</strong> content of the<br />

phosphating solution; (iii) the ratio of FA to TA, expressed<br />

as the acid coefficient; (iv) accelerator content;<br />

(v) iron content; and (vi) other metallic and nonmetallic<br />

constituents present, have to be strictly<br />

controlled within the optimum limits.<br />

3.5.4. Rinsing after phosphating<br />

The <strong>surface</strong> that has been subjected to phosphating<br />

should be thoroughly rinsed with deionized water<br />

to remove any acid residue, soluble salts and<br />

non-adherent particles present on it which would<br />

otherwise promote blistering of paint films used for<br />

finishing. Generally overflow rinsing and spray rinsing<br />

are preferred [186].


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 5. Characteristics of <strong>phosphate</strong> <strong>coatings</strong>.<br />

Characteristic Type of coating<br />

Iron Phosphate Zinc Phosphate Heavy Phosphate<br />

Coating weight 0.16-0.80 g/m 2 1.4 - 4.0 g/m 2 7.5-30 g/m 2<br />

Types Cleaner/coater Standard Manganese <strong>phosphate</strong><br />

Standard Nickel-modified Zinc <strong>phosphate</strong><br />

Organic <strong>phosphate</strong> Low-zinc Ferrous <strong>phosphate</strong><br />

Calcium-modified<br />

Manganese-modified<br />

Common accelerators Nitrite/nitrate Nitrite/nitrate None<br />

Chlorate Chlorate Chlorate<br />

Molybdate Nitrobenzene Nitrate<br />

Sulphonic acid Nitroguanidine<br />

Operating temp. Room <strong>–</strong> 70 o C Room-70 °C 60-100 o C<br />

Free acid (points) -2.0 to 2.0 0.5 - 3.0 3.6 - 9.0<br />

Total acid (points) 5-10 10-25 20-40+<br />

Pre<strong>phosphate</strong> None Titanium <strong>phosphate</strong> Manganese Phosphate<br />

conditioners None Titanium <strong>phosphate</strong><br />

None<br />

Primary use Paint base for low- Paint base for high- Unpainted applications<br />

corrosion environments corrosion environments<br />

Limitations Low painted corrosion Poor unpainted Expensive, long processing<br />

resistance; low corrosion resistance times<br />

unpainted corrosion<br />

resistance<br />

Materials needed Low-carbon steel Low-carbon steel, Stainless steel or low-carbon<br />

for tanks stainless steel or steel<br />

plastic-lined steel<br />

Application method Spray and immersion Spray and immersion Immersion only<br />

3.5.5. Chromic acid sealing<br />

The <strong>phosphate</strong> <strong>coatings</strong> are usually porous in nature.<br />

The porous nature will have a detrimental influence<br />

on the corrosion resistance of the <strong>phosphate</strong><br />

coating unless they are sealed. A hot (70-80 °C)<br />

dilute chromic acid rinse is usually used for this<br />

purpose. This treatment reduces the porosity <strong>by</strong><br />

about 50%. It improves the corrosion resistance <strong>by</strong><br />

the deposition of insoluble chromates on the bare<br />

areas of the coating [25]. In addition to the benefits<br />

derived from precipitating insoluble salts and passivating<br />

any metal that might be exposed, the dilute<br />

chromic acid treatment is advantageous in that it<br />

helps to dissolve protruding crystals of the <strong>phosphate</strong><br />

coating [187]. Besides, chromic acid posttreatment<br />

leaves a residue that is slightly acidic,<br />

which most paints can tolerate. Usually a concen-<br />

141<br />

tration range of 0.0125-0.050% is used. A gross<br />

excess of chromic acid causes blistering, poor adhesion<br />

and irregular yellowing of paint [188]. A mixture<br />

of chromic acid and phosphoric acid has also<br />

been used [188]. Acidic solutions of chromic acid<br />

of pH 2-5 are preferred, the free acid being controlled<br />

between 0.2 and 0.8 points and the total acid below<br />

5 points.<br />

The role of chromates in improving the passivity<br />

of <strong>phosphate</strong>d steel and in improving the paint film<br />

adhesion has been the subject of many papers.<br />

According to Cheever [189], chromium atoms are<br />

distributed over the <strong>surface</strong> and not just between<br />

the <strong>phosphate</strong> crystals. Wenz and Claus [190] have<br />

shown that chromic acid final rinsing completely<br />

removes any adsorbed calcium from tap water rinsing<br />

after phosphating, deposits some chromium and


142 T.S.N. Sankara Narayanan<br />

alters the Zn/Fe ratio. Maeda and Yamamoto [187]<br />

have studied the nature of chromate-treated <strong>phosphate</strong>d<br />

<strong>surface</strong> using XPS. The Cr 2P spectra<br />

3/2<br />

suggest that the coating mainly consisted of trivalent<br />

chromium oxide with a small amount of a<br />

hexavalent chromium compound (Fig. 4).<br />

Deconvolution of the trivalent chromium peak results<br />

in two more peaks which can be attributed to Cr O 2 3<br />

(major part) and Cr(OH) . 1.5H2O (minor part), sug-<br />

3<br />

gesting an intermediate compound (polymerized<br />

state) with -ol (Cr<strong>–</strong>OH) and -oxo (Cr<strong>–</strong>O<strong>–</strong>) bonds.<br />

From the calculation based on the ratio of the two<br />

components, the chemical composition of the chromium<br />

oxide(III) was formulated to be CrOOH . 0.2H O. 2<br />

The CrOOH . 0.2H O contributes to increased paint<br />

2<br />

adhesion due to hydrogen bonding with resin components.<br />

Though the chromic acid sealing improves the<br />

corrosion resistance, the need for regular disposal<br />

of Cr(VI) effluents is a matter of concern because<br />

Cr(VI) causes serious occupational and health hazards.<br />

Several alternatives for Cr(VI) treatment were<br />

proposed (Table 3) [191]. In spite of the development<br />

of various alternative treatments, there still<br />

exists a strong belief that the extent of corrosion<br />

protection provided <strong>by</strong> them is not as good as that<br />

from Cr(VI) treatment and from the point of view of<br />

use on an industrial scale, none of these alternative<br />

post-treatments have been proved to be a completely<br />

acceptable replacement to Cr(VI).<br />

3.5.6. Drying<br />

After chromic acid rinsing the parts must be dried<br />

before finishing, the conventional methods used<br />

being simple evaporation, forced drying <strong>by</strong> blowing<br />

air or <strong>by</strong> heating [88]. Where evaporation conditions<br />

are good, warm air circulating fans and compressed<br />

air blow offs are the most economical methods. After<br />

drying the <strong>phosphate</strong>d panels are ready for application<br />

of further finishes such as paints, oils, varnishes,<br />

etc.<br />

3.6. Coating characteristics<br />

3.6.1. Structure and composition<br />

Phosphate <strong>coatings</strong> produced on steel, zinc, zinccoated<br />

steel, aluminium and other similar metals<br />

show a crystalline structure with crystals ranging<br />

from a few to about 100 micrometer in size. Various<br />

workers [24,192,193] have reported a large number<br />

of different constituents of <strong>phosphate</strong> <strong>coatings</strong>.<br />

Machu [25] listed 30 such <strong>phosphate</strong> compounds<br />

identified in a <strong>phosphate</strong> coating. Neuhaus and<br />

Fig. 4. Cr2p 3/2 XPS spectra of <strong>phosphate</strong> coating<br />

post-treated with chrome rinse (Parlen 60) (Reprinted<br />

from Progress in Organic Coatings, Vol. 33,<br />

S. Maeda and M. Yamamoto, The role of chromate<br />

treatment after phosphating in paint adhesion, pp.<br />

83-89 (1989) with permission from Elsevier Science).<br />

Gebhardt [192] have tabulated the main phases in<br />

<strong>phosphate</strong> <strong>coatings</strong> formed on metals from baths of<br />

various <strong>phosphate</strong>s (Table 6). The composition of<br />

<strong>phosphate</strong> <strong>coatings</strong> is influenced <strong>by</strong> a number of<br />

factors such as the method of application (spray or<br />

dip), the degree of agitation of the bath, bath chemistry,<br />

the type and quantity of accelerator and the<br />

presence of other metal ions. Chamberlain and Eisler<br />

[194] have found using radioactive tracers that the<br />

base layer was formed initially from the metal being<br />

attacked during the first few seconds of contact with<br />

the phosphating bath producing a very thin film. The<br />

film contains oxides and <strong>phosphate</strong>s of the metal<br />

being treated. Ferrous <strong>phosphate</strong> is most likely to<br />

be present in the case of steel. The growth of <strong>phosphate</strong><br />

coating is initiated <strong>by</strong> the formation of a subcrystalline<br />

layer on which crystalline layer of <strong>phosphate</strong>s<br />

build up rapidly. The number of crystals on<br />

which growth has occurred is essentially constant<br />

with time because nucleation and growth takes<br />

place only at a limited constant number of areas


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 6. Phase constituents of <strong>phosphate</strong> <strong>coatings</strong> on Fe, Zn and Al.<br />

Metal in the bath Substrate<br />

Fe Zn Al<br />

Alkali<br />

Fe<br />

Mn<br />

Fe (PO ) . 8H2O Zn (PO ) . 4H2O 3 4 2<br />

3 4 2<br />

Fe H (PO ) . 4H2O Zn (PO ) . 4H2O 3 2 4 4<br />

3 4 2<br />

(Fe hureaulite) Zn Fe(PO ) . 4H2O 2 4 2<br />

FePO . 2H2O (strengite) Fe H (PO ) . 4H2O 4<br />

5 2 4 4<br />

(Mn.Fe) H (PO ) . 4H2O Zn (PO ) . 4H2O 5 2 4 4<br />

3 4 2<br />

Mn H (PO ) . 4H2O 5 2 4 4<br />

(Mn hureaulite)<br />

AlPO4 Mn H (PO ) . 4H2O 5 2 4 4<br />

Zn Zn Fe(PO ) . 4H2O 2 4 2<br />

(phosphophyllite)<br />

Zn (PO ) . 4H2O 3 4 2<br />

Zn (PO ) . 4H2O 3 4 2<br />

(hopeite)<br />

ZnCa Zn Fe(PO ) . 4H2O 2 4 2<br />

Zn Ca(PO ) . 2H2O 2 4 2<br />

(scholzite)<br />

Zn (PO ) . 4H2O 3 4 2<br />

Zn (PO ) . 4H2O 3 4 2<br />

Zn Ca(PO ) . 2H2O 2 4 2<br />

Zn Ca(PO ) . 2H2O 2 2 2<br />

Zn (PO ) . 4H2O 3 4 2<br />

[195]. It is believed that the formation of <strong>phosphate</strong><br />

coating follows an active site mechanism i.e., only<br />

a small percentage of the <strong>surface</strong>, participates in<br />

the nucleation of the growth sites. Machu has postulated<br />

that these active sites were the growth areas<br />

located predominantly at the grain boundaries<br />

of the steel [196].<br />

Light and electron microscopic studies made on<br />

zinc <strong>phosphate</strong>d steel, have shown that the formation<br />

and growth process occur in three stages.<br />

1st stage: Depending on the zinc phosphating process,<br />

10 4 -10 6 /cm 2 platelet crystals are formed<br />

on the <strong>surface</strong>. These are randomly oriented to<br />

the steel substrate; some are parallel, some<br />

vertical and others are inclined at an angle. Those<br />

platelets, which are not, attached parallel to the<br />

<strong>surface</strong> are needle-like in appearance. These first<br />

stage crystals grow primarily laterally over the<br />

substrate.<br />

2nd stage: It consists of nucleation and growth of<br />

several crystals on the upper <strong>surface</strong> of the original<br />

crystals that are attached parallel to the <strong>surface</strong>.<br />

This growth process is usually vertical to<br />

the site and gives the appearance of additional<br />

needle-like crystals.<br />

3rd stage: Finally, a thin layer of zinc <strong>phosphate</strong><br />

spreads from the base of the original crystals.<br />

For paint-base zinc <strong>phosphate</strong> <strong>coatings</strong>, the<br />

143<br />

growth is primarily lateral. For <strong>coatings</strong> used for<br />

oil retention and as bases for cold forming lubricants,<br />

considerable vertical growth also occurs.<br />

The crystal habit and size depends on many factors<br />

like the bath composition, temperature, method<br />

of <strong>surface</strong> preparation etc. Crystals may take the<br />

form of plates, needles and grains having dimension<br />

from a few to tens of micrometers.<br />

Many instrumental methods are available for the<br />

examination of the constituents of <strong>phosphate</strong> <strong>coatings</strong>.<br />

These include energy dispersive spectroscopy<br />

(EDS), electron spectroscopy for chemical analysis<br />

(ESCA) and X-ray diffraction (XRD). Neuhaus et<br />

al. [193] have reported from XRD analysis that<br />

phosphophyllite, Zn Fe(PO ) . 4H2O and hopeite,<br />

2 4 2<br />

Zn (PO ) . 4H2O are the essential constituents of<br />

3 4 2<br />

zinc <strong>phosphate</strong> <strong>coatings</strong> on ferrous substrate, substantiated<br />

<strong>by</strong> others [197, 198] also.<br />

Phosphophyllite is formed essentially at the <strong>surface</strong><br />

of contact with the basis metal. The quantitative<br />

ratio of these phases is variable and depends<br />

on the total iron content of the solution. Iron<br />

Hureaulite, Fe H (PO ) . 4H2O is formed when the<br />

5 2 4 4<br />

Fe(II) content in the solution is high. This is not<br />

stable under atmospheric conditions and has a detrimental<br />

influence proportional to its content in the<br />

coating.


144 T.S.N. Sankara Narayanan<br />

Miyawaki et al. [199] have introduced the concept<br />

of ‘P ratio’ to express quantitatively the proportion<br />

of these phases in <strong>phosphate</strong> <strong>coatings</strong>. It is<br />

defined as:<br />

Phosphophyllite<br />

‘P ratio’ = Phosphophyllite + Hopeite<br />

Several authors [199,200] have correlated the ‘P<br />

ratio’ and corrosion resistance of the <strong>phosphate</strong><br />

coating. But according to Richardson et al. [200]<br />

the ratio itself is not sufficient to predict the corrosion<br />

resistance.<br />

X-ray and electron diffraction studies have shown<br />

that hopeite and phosphophyllite are oriented perpendicular<br />

to the plane of the support besides<br />

exhibiting that there is an excellent orientation between<br />

the substrate and the zinc <strong>phosphate</strong> coating<br />

that follows an epitaxial growth relationship.<br />

(010) - hopeite and (100) - phosphophyllite || (100)α-Fe<br />

(100) - hopeite and (001) - phosphophyllite || A o -α-<br />

Fe and<br />

(001) - hopeite and (010) - phosphophyllite || A o -α-<br />

Fe<br />

The formation of primary valency bonds between<br />

the coating and the polarized elements of the α-iron<br />

lattices with the production of a two dimensional<br />

contact layer of the wustite-type accounts for the<br />

excellent adhesion of the coating to the substrate.<br />

3.6.2. Coating thickness and coating<br />

weight<br />

One of the principal factors involved in the choice of<br />

a phosphating bath is in fact the thickness of the<br />

deposit that it will provide. Neglecting intercrystalline<br />

voids and <strong>surface</strong> irregularities and considering the<br />

<strong>phosphate</strong> coating as completely homogeneous, the<br />

thickness can be measured [201]. Phosphate <strong>coatings</strong><br />

range in thickness from 1 to 50 microns but for<br />

practical purposes the thickness is usually quantified<br />

in terms of weight per unit area (usually as g/m 2<br />

or mg/ft 2 ) and commonly referred to as coating<br />

weight. The reason for the adoption of coating weight<br />

rather than coating thickness as the usual measure<br />

of <strong>coatings</strong> is the difficulty in measuring the<br />

latter, compounded <strong>by</strong> the uneven nature of the substrate<br />

and of the coating.<br />

According to Lorin [127] the ratio between coating<br />

weight (g/m 2 ) and coating thickness (μm) varies<br />

between 1.5 and 3.5 for the majority of industrial<br />

<strong>phosphate</strong> <strong>coatings</strong>. For light and medium weight<br />

<strong>coatings</strong> 1 μm can be regarded as equivalent to 1.5-<br />

2 g/m 2 .<br />

The determination of coating weight is a destructive<br />

test, which involves weighing a standard test<br />

panel before and after stripping the coating in a<br />

medium, which dissolves the coating and not the<br />

substrate. Usually methods such as stripping the<br />

coating in concentrated hydrochloric acid containing<br />

antimony trioxide (20 g/l) as an inhibitor or high<br />

concentration of chromic acid solution (5%) or sodium<br />

hydroxide (15%) is used. A non-destructive<br />

method based on specular reflectance infrared absorption<br />

(SRIRA) for the determination of zinc <strong>phosphate</strong><br />

coating weight has been developed <strong>by</strong> Cheever<br />

[202, 203]. Tony Mansour [204] has shown a good<br />

agreement of the results obtained <strong>by</strong> this method<br />

with X-ray fluorescence (XRF) data as well as the<br />

gravimetric measurements. Yap et al. [205] suggest<br />

that XRF is a nondestructive and accurate technique<br />

for measuring the thickness of thin <strong>phosphate</strong> <strong>coatings</strong>.<br />

The phosphating industry generally uses coating<br />

weight as a method of quality control; but it is<br />

widely agreed that except at extreme values, coating<br />

weight does not directly relate to corrosion performance.<br />

Hence coating weight alone is of little<br />

value in assessing the quality of <strong>coatings</strong> and must<br />

be considered in relation to other characteristics of<br />

the <strong>coatings</strong>, viz., thickness, structure homogeneity,<br />

etc.<br />

3.6.3. Coating porosity<br />

The layer of <strong>phosphate</strong> coating consists of numerous<br />

crystals of very different sizes, which have<br />

spread from centers of nucleation to join and finally<br />

cover the <strong>surface</strong>. A constitution of this type inherently<br />

implies the existence of fissures and channels<br />

through to the basis metal at inter-crystalline<br />

zones. Porosity is generally fairly low, of the order<br />

of 0.5-1.5% of the <strong>phosphate</strong>d <strong>surface</strong> [25]. It is<br />

generally believed that the porosity decreases with<br />

increasing thickness of the <strong>phosphate</strong> coating. Porosity<br />

depends upon the type of <strong>phosphate</strong> solution,<br />

the treatment time, the iron content of the bath<br />

and, the chemical composition of the coating [206].<br />

In recent years, much attention has been focused<br />

on the porosity of <strong>phosphate</strong> <strong>coatings</strong> due to the<br />

presence of tightly bound carbonaceous residues<br />

formed on steel during steel making [207]. Since<br />

these cannot be removed <strong>by</strong> alkaline cleaning process,<br />

they interfere with the effective deposition of<br />

<strong>phosphate</strong> <strong>coatings</strong> resulting in the formation of<br />

porous <strong>coatings</strong> with inferior performance.<br />

Several chemical and electrochemical methods<br />

have been developed to determine the porosity of


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Fig. 5. FTIR absorption-reflection spectra of zinc<br />

<strong>phosphate</strong> coating formed on steel<br />

(a) as-prepared; (b) partially dehydrated at 150 °C<br />

for 1 hour; (c) partially dehydrated and rehydrated<br />

at 95% RH for 70 hours. (Reprinted from Journal of<br />

Molecular Structure, Vol. 511-512, A. Stoch, Cz.<br />

Paluszkiewicz and E. Dlugon, An effect of<br />

methylaminoethoxysilane on zinc <strong>phosphate</strong> rehydration,<br />

pp. 295-299 (1999) with permission from<br />

Elsevier Science).<br />

<strong>phosphate</strong> <strong>coatings</strong>. The time taken for the deposition<br />

of metallic copper from a copper sulphate-sodium<br />

chloride solution and the electron microprobe<br />

determination of the amount of copper deposited in<br />

the bare areas of the coating are the measures of<br />

coating porosity [163]. The number of Prussian blue<br />

spots formed per unit area of a filter paper soaked<br />

in potassium ferricyanide-sodium chloride-gelatin<br />

mixture placed on a <strong>phosphate</strong>d sample is yet another<br />

method of determination of the porosity of<br />

these <strong>coatings</strong> [208].<br />

Machu’s [25] method of determination of porosity<br />

of coating <strong>by</strong> electrochemical means is based<br />

on the anodic passivation of uncoated areas in sodium<br />

sulphate solution. A more recent electrochemical<br />

polarization method is based on the magnitude<br />

of current generated <strong>by</strong> the reduction of oxygen at<br />

the cathodically active sites on the porous coating<br />

[209,210].<br />

Although porosity of the coating has a detrimental<br />

effect on its corrosion performance, it has some<br />

advantages too. The pores in the <strong>phosphate</strong> <strong>coatings</strong><br />

act as large reservoirs into which organic ma-<br />

145<br />

terials can collect [195]. This attribute is made use<br />

of in the protection of coated articles <strong>by</strong> impregnation<br />

of oil in the pores. Thus, a homogeneous finecrystalline<br />

coating is desirable with respect to adhesion<br />

of a paint film while a coarse-crystalline coating<br />

is preferred for protection <strong>by</strong> oils, varnishes etc.<br />

3.6.4. Stability of the <strong>phosphate</strong><br />

coating<br />

The stability of the <strong>phosphate</strong> coating is an important<br />

characteristic property. Since <strong>phosphate</strong> <strong>coatings</strong><br />

serve as an effective <strong>pretreatment</strong> for subsequent<br />

paint finishes, it is imperative that they must<br />

be compatible with the applied paint systems. In<br />

recent years, numerous developments were made<br />

in the paint finishes. Among them, the cathodic<br />

electrophoretic deposition has received a widespread<br />

acceptance. In order to be compatible with<br />

this deposition technology, the <strong>phosphate</strong> coating<br />

must process an excellent thermal and alkaline<br />

stability. The importance of stability of <strong>phosphate</strong><br />

coating to suit the modern day finishing systems<br />

was discussed <strong>by</strong> several authors<br />

[18,19,24,127,211-218].<br />

During cathodic electrophoretic deposition, the<br />

decomposition of water produces hydroxyl ions, the<br />

creation of which is considered to be critical as they<br />

can cause dissolution of the <strong>phosphate</strong> coating [140,<br />

141]. Several researchers have confirmed the dissolution<br />

of <strong>phosphate</strong> coating in high-pH environments<br />

[142-147]. It was found that approximately<br />

30-40% of the <strong>phosphate</strong> coating gets dissolved<br />

during cathodic electrophoretic deposition. This has<br />

lead to a greater porosity of the <strong>phosphate</strong> coating.<br />

Moreover, the occlusion of the dissolved ions, which<br />

are subsequently, concentrated during paint baking<br />

affects the corrosion resistance. After cathodic<br />

electrocoating the coated panels are usually cured<br />

at a temperature of 180 °C for 20 minutes. According<br />

to Kojima et al. [148] and Sugaya and Kondo<br />

[149], under such curing conditions the <strong>phosphate</strong><br />

coating will undergo a definite weight loss associated<br />

with a structural change in the constituent crystals.<br />

It is generally advised that the loss in weight<br />

should be restricted to less than 15%, which would<br />

otherwise cause detoriation of the <strong>phosphate</strong> coating<br />

and a loss in corrosion resistance. The other<br />

important factor is the ability of the dehydrated <strong>phosphate</strong><br />

crystals to revert back to its original hydrated<br />

form when subjected to humid service conditions.<br />

The dehydration-rehydration phenomenon is further<br />

confirmed <strong>by</strong> X-ray diffraction. During heating at 150<br />

°C for 1 hour the lines at 9.65 and at 19° 2θ dimin-


146 T.S.N. Sankara Narayanan<br />

Fig. 6. X-ray diffraction pattern of zinc <strong>phosphate</strong> coating formed on steel (a) as-prepared; (b) partially<br />

dehydrated at 150 °C for 1 hour; (c) partially dehydrated and rehydrated at 95% RH for 70 hours. (Reprinted<br />

from Journal of Molecular Structure, Vol. 511-512, A. Stoch, Cz. Paluszkiewicz and E. Dlugon, An effect of<br />

methylaminoethoxysilane on zinc <strong>phosphate</strong> rehydration, pp. 295-299 (1999) with permission from Elsevier<br />

Science).<br />

ished or disappeared and new lines were observed<br />

at 10.90, 11.20 and 22° 2θ. During rehydration the<br />

line at 11.20° 2θ diminished while that at 9.65° 2θ<br />

grew (Fig. 6). It has been observed that the rehydration<br />

of the <strong>phosphate</strong> crystals induces residual<br />

stresses and reduce the <strong>phosphate</strong>-paint film adhesiveness<br />

[150].<br />

Numerous research reports are available to date,<br />

which elaborate the behavior of <strong>phosphate</strong> <strong>coatings</strong><br />

during chemical and thermal aggressions. The dissolution<br />

of zinc <strong>phosphate</strong> coating in high-pH environments<br />

was first reported <strong>by</strong> Wiggle et al. [142].<br />

Later, using sodium hydroxide as a simulated medium,<br />

Roberts et al. [143] demonstrated the selective<br />

leaching of phosphorous from the coating. Van<br />

Ooij and de Vries [144] interpreted X-ray photoelectron<br />

spectroscopic studies to show that hydroxyl<br />

ions exchange <strong>phosphate</strong> ions in the first few layers<br />

of the <strong>phosphate</strong> coating. Servais et al. [140]<br />

determined the chemical stability of zinc <strong>phosphate</strong><br />

<strong>coatings</strong> <strong>by</strong> subjecting them to immersion treatment<br />

in 0.8 g/l sodium hydroxide solution (pH 12.3) at 40<br />

°C for 3 minutes. Similarly, the effects of high-pH<br />

environments on zinc <strong>phosphate</strong> <strong>coatings</strong> were determined<br />

<strong>by</strong> Sommer and Leidheiser Jr. [141], using<br />

0.01, 0.1 and 1.0 M sodium hydroxide solutions.<br />

Kwiatkowski et al. [151] recommended a borate<br />

buffer solution containing 0.01 M ethylenediaminetetraacetic<br />

acid (EDTA) as the medium to test the<br />

alkaline stability of the coating and suggested its<br />

usefulness in predicting the same at a shorter time<br />

interval. All these studies uniformly agree that <strong>coatings</strong><br />

richer in phosphophyllite possess greater alkaline<br />

solubility and can serve as effective bases<br />

for cathodic electrophoretic painting.<br />

Phosphate <strong>coatings</strong> undergo marked changes<br />

when subjected to variation of temperature.<br />

Thermogravimetric technique is usually used to<br />

monitor these changes [152]. The changes occurring<br />

in zinc <strong>phosphate</strong> coating on steel upon heating<br />

was studied <strong>by</strong> Kojima et al. [148] and Sugaya<br />

and Kondo [149]. When heated above room temperature,<br />

a gradual loss in weight occurs. This was<br />

attributed to a nonstructural weight loss. However,<br />

with subsequent increase in temperature dehydration<br />

of the constituent phases of the <strong>phosphate</strong> crystals,<br />

occurs. The dehydration of hopeite<br />

[Zn (PO ) . 4H2O] and phosphophyllite<br />

3 4 2<br />

[Zn Fe(PO ) . 4H2O] commences respectively at 80<br />

2 4 2<br />

and 110 °C; however, a more pronounced weight<br />

loss occurs at approximately 150 °C, where hopeite


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

loses two molecules of water of hydration. When<br />

heated above 150°C, both phosphophyllite and<br />

hopeite are transformed into their dehydrated forms.<br />

The hopeite phase is completely dehydrated at a<br />

temperature of approximately 240 °C. The weight<br />

loss considerably increased between 250 and 600<br />

°C. Above 600 °C sublimation of zinc and phosphorous<br />

occurs, which results in complete breakdown<br />

of the coating. The change in appearance, colour<br />

and morphology of <strong>phosphate</strong> <strong>coatings</strong> remain practically<br />

unaltered up to 200 °C. However, above this<br />

temperature, the grey crystalline <strong>phosphate</strong> coating<br />

changes to a silver grey form and appears dusty.<br />

Above 500 °C the colour changes to brown and above<br />

600 °C complete breakdown of coating occurs [148].<br />

Bonara and his co-workers [152,153] have accounted<br />

for the observed difference in temperature<br />

at which both the hopeite and phosphophyllite crystals<br />

lose their water of crystallization. According to<br />

them, hopeite crystals seem to possess only one<br />

type of crystallization water species, which may be<br />

readily released during a progressive heating. In<br />

contrast, the phosphophyllite have three different<br />

crystallization water types, each with different bond<br />

types, thus causing a differential release during progressive<br />

heating.<br />

Based on the above facts on thermal stability of<br />

<strong>phosphate</strong> <strong>coatings</strong> and the normal baking conditions<br />

used (180 °C for 20 minutes) for curing cathodic<br />

electrocoating, it is quite obvious that both<br />

hopeite and phosphophyllite will exist as bihydrated<br />

crystals. The rehydration of these bihydrated hopeite<br />

and phosphophyllite was a subject of considerable<br />

importance as it determines the wet adhesion property.<br />

It has now been established that when placed<br />

under a rehydration condition, like immersion in an<br />

aqueous solution or exposure to high humid atmospheres,<br />

the hopeite bihydrated crystals undergo<br />

rehydration to the tetrahydrated phase along with<br />

the formation of zinc oxide. The formation of these<br />

products induces stresses and bond relaxation at<br />

the paint-<strong>phosphate</strong> interface and ultimately affects<br />

the adhesiveness. In contrast to this behavior, the<br />

phosphophyllite bihydrated crystals are found to<br />

resist the rehydration phenomenon. Liebau [233]<br />

has accounted for the difference in rehydration<br />

behaviour of hopeite and phosphophyllite bihydrate<br />

crystals. According to him, in hopeite it is the flexibility<br />

of the Zn 2+ ion co-ordination state, which can<br />

exist in either octahedral or tetrahedral coordination,<br />

thus allowing the dehydration process to occur.<br />

However, in phosphophyllite, the Fe 2+ ions occupy<br />

the octahedral co-ordination sites, which are<br />

the unsaturated co-ordination states. Hence an ir-<br />

147<br />

reversible structural modification has resulted during<br />

the dehydration process.<br />

Hence it is evident that a <strong>phosphate</strong> coating richer<br />

in phosphophyllite phase is best suited to withstand<br />

the environments that occur during cathodic electrophoretic<br />

painting, paint curing and under service<br />

conditions. This has necessitated modification of<br />

the phosphating formulations to produce <strong>phosphate</strong><br />

<strong>coatings</strong> richer in phosphophyllite phase. The influence<br />

of the phase constituents on the stability of<br />

<strong>phosphate</strong> <strong>coatings</strong> studied <strong>by</strong> Sankara Narayanan<br />

[213] also confirms this generalization.<br />

The introduction of pre-coated steels such as<br />

zinc, zinc alloy coated steels, etc., for automotive<br />

body panels present new challenges regarding the<br />

stability of <strong>phosphate</strong> <strong>coatings</strong> [234]. When coated<br />

using a zinc phosphating bath, the zinc and zinc<br />

alloy coated steels will produce <strong>phosphate</strong> <strong>coatings</strong>,<br />

which comprise only the hopeite phase. As<br />

stated earlier, <strong>phosphate</strong> <strong>coatings</strong> richer in hopeite<br />

is not desirable for subsequent cathodic<br />

electrocoating finishes. However, these pre-coated<br />

steels show good performance with regard to formability<br />

and weldability. The extent of perforation and<br />

cosmetic corrosion is also less with these materials<br />

and they also posses a good <strong>surface</strong> appearance.<br />

Hence to effectively make use of the zinc and<br />

zinc alloy coated steels, the phosphating formulation<br />

has to be modified to produce <strong>coatings</strong> that are<br />

not richer in hopeite. It is now realized that although<br />

low hopeite content is the principle requirement for<br />

cathodic electrocoating, it is not essential that the<br />

coating should be richer in phosphophyllite as long<br />

as the coating has substantial amorphous material.<br />

The introduction of nickel and manganese modified<br />

low-zinc phosphating formulations are the most<br />

significant modification proposed to cope up with<br />

the cathodic electocoating, particularly in the automobile<br />

industry. The presence of Ni 2+ and/or Mn 2+<br />

ions in these phosphating baths has yielded several<br />

advantages. Although there is a decrease in<br />

coating weight, crystal formation is observed at an<br />

early stage and the crystal size is fine for zinc <strong>phosphate</strong><br />

crystals obtained from solutions containing<br />

Ni 2+ and Mn 2+ ions. It is confirmed that these heavy<br />

metal ions participate in forming the crystal and<br />

brings about the crystal refinement and such an effect<br />

is found to increase with an increase in their<br />

concentration [235,236]. It is also reported that Mn 2+<br />

ions are more effective in causing the nucleation of<br />

the crystal when compared to the effect of Ni 2+ ions<br />

[236]. The usefulness of the addition of Mn 2+ ion<br />

has been advocated based on several other factors.


148 T.S.N. Sankara Narayanan<br />

Mn 2+ ions demonstratably improve the corrosion<br />

resistance of <strong>coatings</strong> obtained from a low zinc phosphating<br />

process. The presence of Mn 2+ ions in low<br />

zinc phosphating bath increases the rate of formation<br />

of <strong>phosphate</strong> <strong>coatings</strong>. Hence, it is possible to<br />

decrease the bath temperature, which in turn allows<br />

a considerable amount of savings in the heating<br />

costs. Moreover, the addition of Mn 2+ ions in the<br />

phosphating bath helps to increase the working width<br />

of the phosphating bath. Also, the presence of manganese<br />

in the zinc <strong>phosphate</strong> coating resists the<br />

formation of white spots on galvanized steel. However,<br />

it is also cautioned that if the manganese content<br />

exceeds a certain level, a decrease in corrosion<br />

resistance may occur. Advanced analytical<br />

techniques have shown that the manganese tends<br />

to be distributed throughout the coating whereas<br />

nickel is concentrated at the interface and the presence<br />

of both contributes to maximum performance.<br />

Based on the results of the electron spin resonance<br />

(ESR) technique, Sato et al. [237] confirmed that<br />

both nickel and manganese are present in the zinc<br />

<strong>phosphate</strong> films respectively, as Ni(II) and Mn(II).<br />

According to them, the nickel and manganese doped<br />

hopeite are formed <strong>by</strong> the following mechanisms:<br />

Zn(H PO ) → ZnHPO + H PO , (4)<br />

2 4 4 3 4<br />

3Zn(H PO ) → Zn Me (PO ) . 4H2O + H PO . (5)<br />

2 4 3-x x 4 2<br />

3 4<br />

The chemical structure of the modified hopeite<br />

crystal in general was considered as<br />

Zn Me (PO ) . 4H2O, where Me = Ni or Mn.<br />

3-x x 4 2<br />

Accordingly, the chemical structure of the doped<br />

hopeite when nickel and manganese coexist is<br />

suggested as Zn (Ni Mn )(PO ) . 4H2O. The<br />

3-x-z x Z 4 2<br />

possibility of existence of such a structure was<br />

confirmed <strong>by</strong> various analytical techniques such as<br />

electron spin resonance (ESR), extended X-ray<br />

absorption fine structure (EXAFS), laser Raman<br />

spectroscopy, etc. [237-243].<br />

Besides the modification of the phosphating bath<br />

<strong>by</strong> Ni2+ and/or Mn2+ ions, developments were also<br />

made in modifying the <strong>surface</strong> of the steel sheets.<br />

The formation of a Fe-P coating of about 2 g/m2 containing a phosphorous content of 0.5% or below,<br />

on Zn-Ni or Zn-Fe steel sheets is another notable<br />

development [234, 244]. These materials are<br />

commonly known as double-layered Fe-P/Zn-Ni and<br />

Fe-P/Zn-Fe alloy plated steel sheets. The<br />

phosphatability of these steel sheets are highly<br />

comparable to that of cold-rolled steel sheets. It is<br />

the presence of micro quantity of phosphorous, uniformly<br />

dispersed in the iron <strong>coatings</strong>, which activates<br />

the <strong>surface</strong> and enhances the reaction with<br />

the phosphating bath, resulting in the production of<br />

closely packed phosphophyllite rich coating, having<br />

a crystal size of 3-5 mm. The formation of such<br />

a coating improves the wet adhesion property. Hence<br />

the modifications made in the phosphating formulations<br />

to produce <strong>coatings</strong> richer in phosphophyllite,<br />

to serve as an effective base for electrocoat finishing<br />

systems [225], are rightly justified.<br />

3.7. Influence of various factors on<br />

coating properties<br />

3.7.1. Nature of the substrate<br />

3.7.1.1. Composition of the metal. The presence<br />

of alloying elements and their chemical nature cause<br />

distinct difference in the extent of phosphatability of<br />

the substrate. It is generally believed that steels<br />

with small amounts of more noble metals such as<br />

chromium, nickel, molybdenum and vanadium can<br />

be <strong>phosphate</strong>d without much difficulty. However, if<br />

the concentration of these elements exceeds certain<br />

limit then the steel will encounter with a decrease<br />

in the intensity of acid attack resulting in a<br />

poor coating formation and among the noble metals<br />

in causing such an effect, chromium is considered<br />

to be the most detrimental one [245]. The<br />

phosphatability of cold rolled steel containing varying<br />

amounts of chromium, nickel and copper is given<br />

in Table 7. It is evident from Table 7 that the<br />

phosphatability is not affected when the total concentration<br />

of chromium, nickel and copper is in the<br />

range of 200-300 ppm and gets drastically affected<br />

when the total concentration of these metals exceeds<br />

800 ppm. The amount of carbon, phosphorus,<br />

sulphur, manganese and silicon can also influence<br />

the phosphatability of the steel to a great extent.<br />

Low carbon steels undergo phosphating easily<br />

and produce superior quality <strong>coatings</strong>. With increasing<br />

carbon content, the rate of phosphating<br />

becomes slower and the resultant crystals are<br />

larger. In fact, when dealing with the carbon concentration<br />

in steel, it is the distribution and the form<br />

in which it is present is rather an important criterion<br />

in predicting its probable effect on sensitization,<br />

nucleation and crystallization. The presence of ferrite<br />

crystals improve metal attack while on the other<br />

hand increasing concentrations of pearlite leads to<br />

coarsening of the <strong>phosphate</strong> crystals as the crystal<br />

nuclei develop only to a small extent on the islands<br />

of pearlite [127,246]. The deleterious effect of <strong>surface</strong><br />

carbon content has been the subject of many<br />

investigations and based on the amount of <strong>surface</strong><br />

carbon, steels have been classified as ‘bad’ or ‘good’<br />

with reference to their suitability to phosphating


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 7. Phosphatability of cold rolled steel containing varying amounts of chromium, nickel and copper<br />

(after Kim [245]).<br />

Substrate Cu + Ni + Cr Surface roughness Phosphat- Coating weight (g/m 2 )<br />

designation (ppm) (R a ) (mm) ability* In the centre In the edge<br />

CRS-A 300 1.2 1 4.04 5.39<br />

CRS-C 800 1.2 4 3.54 3.59<br />

* Based on visual appearance. Rating: 1 <strong>–</strong> Good; 5 <strong>–</strong> Poor.<br />

[247,248]. Grossman [248] has found up to 30 times<br />

as much carbon on “bad” steels as on the “good”<br />

steel samples. A direct correlation between <strong>surface</strong><br />

carbon content and the porosity of the <strong>phosphate</strong><br />

<strong>coatings</strong> on the one hand and their salt spray resistance<br />

when painted on the other hand was found <strong>by</strong><br />

Hospadaruk et al. [207,249]. Wojkowiak and Bender<br />

[250] using a multiple regression analysis have confirmed<br />

the existence of a positive correlation between<br />

the extent of underfilm attack and the <strong>surface</strong> carbon<br />

content of the steel.<br />

Takao et al. [251,252] have investigated the effect<br />

of phosphorus on the phosphatability of ultra<br />

low-carbon steel sheets. According to them, the<br />

addition of phosphorus was beneficial in refining the<br />

grain size and increasing the <strong>surface</strong> coverage.<br />

Moreover, they have proved that the phosphorus<br />

addition helps to improve the perforation corrosion<br />

resistance after painting. In contrast, Kargol and<br />

Jordan [253] have claimed that phosphorus alloy<br />

addition inhibits the hydrogen recombination during<br />

the initial stages of phosphating resulting in increased<br />

porosity and inferior corrosion performance. The rate<br />

of attack of the metal in phosphating solution is also<br />

affected <strong>by</strong> the concentration of sulphur. The<br />

phosphatability of the dual-phase (Si-Mn) steel is<br />

primarily controlled <strong>by</strong> the balance of silicon and<br />

manganese [254].<br />

Although it is clearly evident that every common<br />

constituent element of steel has its own influence<br />

in determining its phosphatability either to a greater<br />

or smaller extent, the most unfavorable from the<br />

phosphating point of view are those which are alloyed<br />

with carbide forming elements such as chromium<br />

and tungsten, and the <strong>surface</strong> carbon content.<br />

As a result, attempts have been made to eliminate<br />

such troublesome factors. Accordingly, the<br />

<strong>surface</strong> carbon content of the steel chosen for phos-<br />

149<br />

phating has to be within the specified limits. Fujino<br />

et al. [255] have concluded that a contamination of<br />

greater than 8 mg/m 2 deteriorates the <strong>phosphate</strong><br />

coating and resistance to corrosion after painting.<br />

Blumel et al. [256] and Balboni [257] have predicted<br />

an upper limit of 7 mg/m 2 . But Coduti [258] has concluded<br />

that this amount has to be a maximum of<br />

4.3 g/m 2 for good performance since only an average<br />

performance has resulted when the concentration<br />

lies between 4.3 and 6.4 mg/m 2 .<br />

3.7.1.2. Structure of the metal <strong>surface</strong>. The structure<br />

of the metal <strong>surface</strong> has also its own influence<br />

and considered to be equally as important as that<br />

of the composition. Ghali [259] has made a thorough<br />

study of the <strong>surface</strong> structure in relation to<br />

<strong>phosphate</strong> treatment. It is generally believed that<br />

greater the <strong>surface</strong> roughness, the higher the weight<br />

of the coating deposited per unit of the apparent<br />

<strong>surface</strong> area and the shorter the time of treatment<br />

required. Moreover, increasing <strong>surface</strong> roughness<br />

gives a corresponding increase in the fineness of<br />

the coating structure whereas polished <strong>surface</strong>s<br />

respond poorly to phosphating. Beauvais and Bary<br />

[260] have reported that if the metal <strong>surface</strong> possesses<br />

a greater number of <strong>surface</strong> concavities and<br />

fissures, there will be an increased acid attack during<br />

phosphating resulting in good anchorage of the<br />

coating. According to Kim [245], phosphatability can<br />

be improved <strong>by</strong> increasing the <strong>surface</strong> roughness.<br />

Balboni [257] has considered that a <strong>surface</strong> roughness<br />

of 0.76-1.77 μm is acceptable for most of the<br />

cases.<br />

The growth of the <strong>phosphate</strong> coating is considered<br />

to be an epitaxial growth phenomena and literature<br />

reports have convincingly established the<br />

following epitaxial relationships in the case of a zinc<br />

<strong>phosphate</strong> coating on mild steel constituting the<br />

phosphophyllite (FeZn (PO ) . 4H2O) and the hopeite<br />

2 4 2


150 T.S.N. Sankara Narayanan<br />

(Zn 3 (PO 4 ) 2 .4H 2 O), <strong>by</strong> means of X-ray diffraction. According<br />

to Ursini [261, 262] the orientation relationship<br />

(412) phosphophyllite /(111) Fe α role is very<br />

important for epitaxial growth on cold rolled steel<br />

and for an increasing adhesive bond. Laukonis [142]<br />

has studied the role of the oxide films on phosphating<br />

of steel and concluded that it is easier to deposit<br />

zinc <strong>phosphate</strong> <strong>coatings</strong> on ferric oxides than<br />

on ferrous oxides. Hence it is evident that it is not<br />

only the <strong>surface</strong> roughness but also the orientation<br />

and the presence of the oxide film will influence the<br />

phosphating process to a considerable extent.<br />

3.7.1.3. Surface preparation. The cleaning methods<br />

adopted can also influence the phosphating of<br />

steel sheets. It is essential to remove any greasy<br />

contaminants and corrosion products from the <strong>surface</strong><br />

to obtain a good <strong>phosphate</strong> finish. Degreasing<br />

in organic solvents usually promotes the formation<br />

of fine-grained <strong>coatings</strong> while strong alkaline solutions<br />

and pickling in mineral acids yield coarse <strong>coatings</strong>.<br />

Though, the prevention of the excessive metal<br />

attack during pickling is effected <strong>by</strong> potent inhibitors,<br />

it is generally observed that <strong>surface</strong>s that have<br />

been pickled with an effective pickling inhibitor are<br />

difficult to <strong>phosphate</strong> unless a strong alkaline cleaning<br />

operation is employed to remove the inhibitor<br />

residues. Moreover, it is believed that the smut<br />

formed during pickling operation can also influence<br />

the amount and crystal size of the <strong>phosphate</strong> coating<br />

[188]. However, there are reports available, which<br />

are successfully employing potent inhibitors in the<br />

pickling bath, which eliminates the usual detrimental<br />

effects on phosphating. It is claimed that the<br />

inhibitors adsorbed onto the metal <strong>surface</strong> reduces<br />

the amount of hydrogen in the <strong>surface</strong> layer of the<br />

<strong>phosphate</strong>d metal and the <strong>phosphate</strong> coating produced<br />

in such a way are satisfactory for lowering<br />

friction, facilitating cold working and as undercoats<br />

for paint [263].<br />

3.7.1.4. Surface activation. The activating effect<br />

of colloidal titanium <strong>phosphate</strong> was discovered <strong>by</strong><br />

Jernstedt [264] and latter explored <strong>by</strong> several others<br />

[265]. The mechanism of the activation has been<br />

established <strong>by</strong> Tegehall [266]. The colloids in the<br />

aqueous dispersion are disc shaped particles which<br />

have the composition of Na TiO(PO ) . 0-7H2O. These<br />

4 4 2<br />

particles are physically adsorbed on the metal <strong>surface</strong><br />

during the application of colloidal dispersion.<br />

When the activated substrate comes in contact to<br />

a zinc phosphating bath, an ion exchange between<br />

the sodium ions on the <strong>surface</strong> of the titanium <strong>phosphate</strong><br />

particles and the zinc ions of the phosphating<br />

solution takes place [267]. The ion-exchanged<br />

particles act as nucleation agents for the zinc <strong>phosphate</strong><br />

crystals because they have nearly the same<br />

stoichiometry and offer a crystallographic plane for<br />

an epitaxial growth.<br />

The rate of nucleation of the <strong>phosphate</strong> coating<br />

may be markedly increased <strong>by</strong> imposing an additional<br />

activation of the cleaned metal <strong>surface</strong>. A process<br />

for activating steel <strong>surface</strong> prior to phosphating<br />

was patented <strong>by</strong> Yamamato et al. [268] and<br />

Donofrio [269]. Hamilton [270] has proposed a combination<br />

of acid cleaning and a phosphating compound<br />

for <strong>surface</strong> activation, while Hamilton and<br />

Schneider [271] and Morrison and Deilter [272] have<br />

proposed a highly alkaline titanated cleaner prior to<br />

phosphating. Use of 1-2% disodium <strong>phosphate</strong> solution<br />

containing 0.01% of titanium compounds have<br />

been extensively used in industries [273]. Other<br />

compounds like dilute solutions of cupric or nickel<br />

sulphates, oxalic acid and polyphosphonates help<br />

in increasing the number of initial nuclei formed during<br />

phosphating and their subsequent growth, to<br />

yield thin and compact <strong>coatings</strong> of fine-grained nature.<br />

3.7.1.5. Thermal treatments and machining.<br />

Thermal or thermo-chemical treatments cause the<br />

formation of heterogeneous phases, which usually<br />

alter the grain size. Since the initial metal attack<br />

during phosphating occurs mainly at the grain boundaries,<br />

the size of the grains becomes an important<br />

factor in influencing phosphating and dictates the<br />

effect of thermal or thermo-chemical treatments.<br />

Moreover, upon thermal treatment, the distribution<br />

of the constituent elements of heterogeneous alloys<br />

may vary and depending upon which the cathodic<br />

and anodic sites on the <strong>surface</strong> will result<br />

and decides the phosphatability. When studying the<br />

effects of annealing of a 0.3Mn-0.2Ti modified steel<br />

at 280 °C in H 2 -N 2 atmosphere, Usuki et al. [274]<br />

has concluded that such a treatment leads to the<br />

formation of manganese oxide and titanium oxide.<br />

Since the <strong>surface</strong> concentration of manganese is<br />

3-4 times as large as that of titanium, the treatment<br />

enhances the phosphatability of the modified steel<br />

<strong>by</strong> promoting the formation of a predominant proportion<br />

of manganese oxide on the <strong>surface</strong>. Like<br />

wise, Hada et al. [275] have studied the effect of<br />

thermal treatment on the phosphatability of the uncoated<br />

side of one-side painted steel. When heated<br />

for 2 min. at 280 °C, the crystal size and the coating<br />

weight were increased and the ‘P’ ratio was declined<br />

causing an inferior paint adhesion. On the<br />

other hand when heated for a longer period there<br />

results the diffusion of manganese from the bulk to<br />

the <strong>surface</strong> thus improving the phosphatability.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

The influence of cold-rolling and the conditions<br />

of annealing on the formation of <strong>phosphate</strong> <strong>coatings</strong><br />

were studied <strong>by</strong> several authors [276-278] According<br />

to Shimada and Maeda [278], increased deformation<br />

during cold-rolling and a high-treatment<br />

temperature favour the <strong>phosphate</strong> coating formation<br />

with a good <strong>surface</strong> coverage whereas a rapid rise<br />

in temperature in the heat treatment yields an<br />

unfavourable condition. Cavanagh and Ruble [279]<br />

have showed that tempering treatments following<br />

<strong>surface</strong> hardening processes has a significant effect<br />

on the response of steel to phosphating. Machined<br />

<strong>surface</strong>s accept <strong>phosphate</strong> coating very easily<br />

and often experience a very high rate of deposition.<br />

3.7.2. Phosphating parameters<br />

The composition of the phosphating solution and<br />

the concentration of its constituents determine the<br />

nature of the <strong>coatings</strong> formed. Higher concentrations<br />

of heavy metal ions in accelerated phosphating<br />

solutions yield <strong>coatings</strong> of better protective value<br />

[280]. Concerning the acidity of the bath, the free<br />

acid value (FA), total acid value (TA) and their ratio<br />

(FA:TA) should be maintained at the required optimum<br />

level to obtain <strong>coatings</strong> of improved quality.<br />

Increase in total acid pointage generally produce<br />

<strong>coatings</strong> of higher coating weight and the free acid<br />

value is considered to be critical since the initial<br />

etching attack of the free phosphoric acid present<br />

in the bath forms the basis of <strong>phosphate</strong> coating<br />

formation. As the <strong>conversion</strong> of soluble primary <strong>phosphate</strong><br />

into insoluble heavy metal tertiary <strong>phosphate</strong><br />

takes place with the regeneration of phosphoric acid,<br />

it is believed that a certain amount of free phosphoric<br />

acid must be present to repress the hydrolysis<br />

and to keep the bath stable for effective deposition<br />

of <strong>phosphate</strong> coating. Higher temperatures favour<br />

easy precipitation of tertiary <strong>phosphate</strong>s in a shorter<br />

time. Hence, more amount of phosphoric acid is<br />

needed for the baths operating at higher temperatures.<br />

In contrast, in the case of phosphating baths<br />

operated at room temperature, the possibility of the<br />

increase in acidity during continuous operation is<br />

more likely [129,130] and is normally neutralized<br />

<strong>by</strong> the addition of the carbonate of the metal, which<br />

forms the coating (Zinc carbonate in zinc phosphating<br />

bath). Hence, depending upon the working temperatures<br />

and the concentration of the constituents<br />

in the bath, the free phosphoric acid content must<br />

be chosen to maintain the equilibrium conditions.<br />

Too much of phosphoric acid not only delays the<br />

151<br />

formation of the coating, but also leads to excessive<br />

metal loss.<br />

Literature reports have revealed the effects of<br />

insufficient metal dissolution as well as the high<br />

acidity of the phosphating bath [281,282]. According<br />

to Kanamaru et al. [281] when the dissolution of<br />

iron from the steel during phosphating is insufficient,<br />

the hopeite epitaxial growth plane conforms to the<br />

phosphophyllite epitaxial growth plane (100) and<br />

mixed crystals with a high Zn/P ratio grow along<br />

the steel <strong>surface</strong> whereas when the dissolution is<br />

sufficient, the iron concentration in the solution increases,<br />

the intrinsically free precipitation of<br />

phosphophyllite without the restraint of epitaxy becomes<br />

predominant and <strong>phosphate</strong> crystals with a<br />

lower Zn/P ratio grow. However, when the acid concentration<br />

is very high, the <strong>phosphate</strong>d material was<br />

found to be susceptible to hydrogen embrittlement.<br />

The effect is compounded if the exposure time is<br />

also increased and it is generally believed that even<br />

the relief treatments that are normally practiced are<br />

not effective when the acid concentration is abnormally<br />

high [282]. Hence, it is clearly evident that<br />

the acid coefficient must be maintained properly to<br />

produce good quality <strong>coatings</strong>. Usually, baths of<br />

higher acid coefficients form more fine-grained <strong>coatings</strong>.<br />

The <strong>conversion</strong> ratio, defined as the ratio of<br />

amount of iron dissolved during phosphating and the<br />

corresponding coating weight [31] is considered to<br />

be important as they can predict the efficiency of<br />

the phosphating process. However, the inability to<br />

generalize its definition for many phosphating baths<br />

limits its acceptance as a control parameter in phosphating<br />

operations [283].<br />

El-Mallah et al. [284] have considered the pH of<br />

the bath as initial pH and the pH at which the consumption<br />

of all the free phosphoric acid completes<br />

as the final pH and correlates the difference between<br />

them with the extent of phosphatability. According<br />

to them, reducing the difference between the initial<br />

and final pH leads to the acceleration of the phosphating<br />

process and in causing such an effect the<br />

heavy metals and reducing agents such as hydrazine,<br />

potassium borohydride were evaluated. Several<br />

additives have been attempted to maintain the<br />

pH of the phosphating baths. In this respect, organic<br />

acids and their salts proved their ability to<br />

buffer the pH of an iron phosphating bath made up<br />

with hard water [285]. Similarly, the decrease of the<br />

total and free acidity <strong>by</strong> means of some buffer additives<br />

to prevent any significant dissolution of zinc<br />

coating during the phosphatization of zinc coated<br />

steels.


152 T.S.N. Sankara Narayanan<br />

The concentration of the accelerators is also very<br />

important. Though increase in accelerator concentration<br />

favours better coating formation, too high concentration<br />

may cause passivation of metal <strong>surface</strong><br />

and inhibit the growth. When phosphating bath is<br />

modified <strong>by</strong> the addition of <strong>surface</strong> active agents,<br />

the concentration of accelerator should be optimized<br />

in accordance with the influence of these additives<br />

[286]. The iron content of the bath is also very critical.<br />

Although, a small quantity of iron salts favour<br />

<strong>phosphate</strong> precipitation (Break-in of the bath), the<br />

corrosion performance is largely affected as the fraction<br />

of the ferrous salts in the coating is increased<br />

[287]. The unfavourable effect of Fe(II) is usually<br />

decreased <strong>by</strong> the addition of complexing agents like<br />

Trilon B [288]. A method for controlling the iron content<br />

of zinc phosphating bath was proposed <strong>by</strong> Hill<br />

[289]. In the case of an iron phosphating bath, the<br />

ratio of ferrous to ferric iron in solution has been<br />

related to the tendency for flash rusting. The frequency<br />

and intensity of flash rust increases as the<br />

ratio of ferrous to ferric iron in solution decreased<br />

[290].<br />

Addition of specific compounds to the phosphating<br />

baths has also their own influence on phosphating.<br />

The incorporation of calcium ions in a zinc phosphating<br />

bath has resulted in a considerable change<br />

in the crystal structure, grain-size and corrosion<br />

resistance of the <strong>phosphate</strong> <strong>coatings</strong>. In fact, the<br />

structure of the zinc <strong>phosphate</strong> coating changes<br />

from the phosphophyllite-hopeite to schlozite-hopeite<br />

[291]. The reduction in grain-size (25 μm to 4 μm)<br />

and the improvement in the compactness of the<br />

coating and corrosion resistance, makes this kind<br />

of modification of zinc phosphating baths as an important<br />

type of phosphating and it has been classified<br />

as calcium-modified zinc phosphating [172].<br />

Similarly, the inclusion of manganese and nickel<br />

ions in the zinc phosphating bath proves to be useful<br />

in refining the crystal size and improving the corrosion<br />

resistance of the resultant <strong>phosphate</strong> <strong>coatings</strong>.<br />

It has been established that manganese and<br />

nickel modified zinc <strong>phosphate</strong> deposits on steel<br />

have an ordered structure and possess a high corrosion<br />

resistance [292]. The immense use of manganese<br />

and nickel ions in modifying the hopeite<br />

deposits obtained on galvanized steel in such a way<br />

that the modified hopeite deposit becomes equivalent<br />

to phosphophyllite to withstand the thermal and<br />

chemical aggressions, was very well established<br />

[293] and as a result of such a pronounced influence<br />

of the nickel and manganese ions, they have<br />

also been classified as nickel and manganese modi-<br />

fied zinc phosphating [172]. Moreover, such modifications<br />

have lead to the development of tri-cation<br />

phosphating baths consisting of zinc, manganese<br />

and nickel ions which explored its potential utility<br />

to <strong>phosphate</strong> aluminum, steel and galvanized steel<br />

using the same formulation [294-296]. Besides<br />

these major types of modifications, phosphating<br />

baths have experienced a variety of additives incorporated<br />

in the bath intended for a specific purpose.<br />

Each additive has its own influence on the<br />

phosphatability of steel depending upon the operating<br />

conditions [297-301]. Sankara Narayanan et al.<br />

[41] have classified the type of special additives used<br />

in phosphating and the purpose of their use. The<br />

favourable condition and the precautions in using<br />

these additives were recommended in the respective<br />

documents.<br />

Every phosphating bath reported in literature has<br />

a specific operating temperature and the baths were<br />

formulated in such a way that an equilibrium condition<br />

of the <strong>conversion</strong> of soluble primary <strong>phosphate</strong><br />

to insoluble tertiary <strong>phosphate</strong> exists at that temperature.<br />

Insufficient reach of temperature does not<br />

favour the precipitation of tertiary <strong>phosphate</strong> resulted<br />

from the <strong>conversion</strong>. However, when the baths were<br />

overheated above the recommended operating temperature,<br />

it causes an early <strong>conversion</strong> of the primary<br />

<strong>phosphate</strong> to tertiary <strong>phosphate</strong> before the<br />

metal has been treated and as a result increases<br />

the free acidity of the bath, which consequently<br />

delays the precipitation of the <strong>phosphate</strong> coating.<br />

Hence, in an operating line, which has operating<br />

time, this effect leads to the production of <strong>phosphate</strong><br />

<strong>coatings</strong> with a poor coating weight and inferior<br />

corrosion performance. Sankara Narayanan and<br />

Subbaiyan [302] have discussed the decisive role<br />

of overheating the phosphating baths and outlines<br />

the possible way of eliminating the difficulties encountered<br />

due to this problem.<br />

Likewise, every formulation has been assigned<br />

a fixed operating time based on the kinetics of the<br />

phosphating process using the particular bath followed<br />

<strong>by</strong> coating weight measurements or potential<br />

measurements with respect to the treatment<br />

time. It has been established <strong>by</strong> many workers that<br />

increasing the treatment time beyond the saturation<br />

point do not have any influence on the performance<br />

of the coating. But it has been warned that<br />

any attempt of reducing the treatment time to decrease<br />

the amount of <strong>phosphate</strong> deposition will be<br />

disasters.<br />

The method of deposition of <strong>phosphate</strong> coating<br />

or the phosphating methodology can influence the


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

coating formation to a great extent. The most commonly<br />

used methods of deposition namely the dip<br />

and spray process, although have received considerable<br />

attention, it is believed that the dip process<br />

is most likely to produce an equiaxed structure<br />

which is consider to be beneficial from the corrosion<br />

protection point of view. The influence of permanent<br />

magnetic field during phosphating is studied<br />

<strong>by</strong> Bikulcius et al. [303]. According to them,<br />

when exposed to perpendicular magnetic field, the<br />

zinc <strong>phosphate</strong> coating results in larger crystallites<br />

with lower corrosion resistance whereas the effect<br />

of parallel magnetic field is insignificant. However,<br />

the application of a magnetic field either in perpendicular<br />

or parallel mode results in a fine-grained,<br />

uniform and more compact zinc calcium <strong>phosphate</strong><br />

coating.<br />

Electrochemical method of phosphating always<br />

yields a heavy coating weight [148,304]. In recent<br />

years, studies have been attempted in electrochemical<br />

phosphating to produce <strong>phosphate</strong> <strong>coatings</strong><br />

richer in a particular phase <strong>by</strong> selecting an appropriate<br />

applied potential [147]. The success of this<br />

kind of phosphating methodology has been restricted<br />

<strong>by</strong> the difficulties encounter in adding<br />

‘electrics’ to the existing process plants. Mechanical<br />

vibration of steel during phosphating is considered<br />

as a method for obtaining a fine grained <strong>phosphate</strong><br />

coating. However, the amount of coating<br />

formed is found to be inversely proportional to the<br />

frequency of vibration [305]. Ultrasonically induced<br />

cavitation produces a great number of active centres,<br />

which results in a high rate of nucleation. This<br />

leads to a uniform fine-grained <strong>phosphate</strong> coating<br />

with low porosity [307,308].<br />

3.8. Processing problems and<br />

remedial measures<br />

Maintenance of the bath parameters and operating<br />

conditions at an optimum level is a major and complicated<br />

process, especially in continuous and largescale<br />

phosphating. In actual practice, the finishers<br />

have come across several situations that the bath<br />

is not working properly. These problems viz., overheating<br />

of the bath, creation of an excess acidity at<br />

the metal/solution interface especially in the case<br />

of cold phosphating, change in acidity of the bath<br />

due to carryover of the alkaline solution used for<br />

degreasing, the local increase in acidity due to the<br />

scaling of the heating coils and so on, were discussed<br />

elsewhere [129, 302]. Moreover, baths that<br />

have been used for a long time are rendered ineffective<br />

due to the formation of sludge of ferric phos-<br />

153<br />

phate. Sludging occurs to a greater extent in<br />

unaccelerated and highly acidic baths and has to<br />

be removed periodically to assure proper bath operation<br />

[18,19]. In order to maintain bath parameters<br />

and to enhance bath life, make-up solutions and<br />

solids are usually required. These additives are so<br />

formulated that they can be handled easily and are<br />

inexpensive.<br />

3.9. Defects in <strong>coatings</strong> and remedies<br />

The most commonly encountered defects in <strong>phosphate</strong>d<br />

parts are low corrosion resistance, stained<br />

coating with variable corrosion resistance and <strong>coatings</strong><br />

covered with a loose white powdery deposit.<br />

The possible cause for these defects are insufficient<br />

care in degreasing and cleaning, incorrect acid coefficient,<br />

incorrect solution composition, use of incorrect<br />

operating conditions, incorrect maintenance<br />

of chemicals, excessive sludge formation and faults<br />

in after-treatment. These causes, either singly or in<br />

combination, lead to these defects.<br />

Low corrosion resistance of <strong>coatings</strong> may be a<br />

result of incorrect acid coefficient, incorrect bath<br />

parameters and operating conditions, presence of<br />

certain metallic impurities like aluminium, antimony,<br />

tin and lead compounds and the presence of chloride<br />

ions in the bath. Improper sealing and abnormally<br />

thin <strong>coatings</strong> may also lead to inferior corrosion<br />

resistance. Stained <strong>coatings</strong> may be formed<br />

due to improper cleaning and degreasing. Incorrect<br />

distribution of articles in the phosphating bath and<br />

wrong ratio of work <strong>surface</strong> to solution volume may<br />

be other causes. Over heating of the phosphating<br />

bath, make-up during processing, heavy sludging<br />

and sludge suspended in the bath leads to the formation<br />

of <strong>coatings</strong> with a loose powdery deposit.<br />

Proper cleaning and degreasing, correct maintenance<br />

of work <strong>surface</strong> to solution volume ratio,<br />

control of bath parameters and operating conditions<br />

within the strict limits, avoidance of overheating and<br />

excessive sludge formation will yield <strong>coatings</strong> of consistent<br />

good quality and corrosion resistance.<br />

3.10. Characterization of <strong>phosphate</strong><br />

<strong>coatings</strong><br />

A variety of instrumental methods, which include,<br />

scanning electron microscopy (SEM), electron probe<br />

microanalysis (EPMA), X-ray diffraction (XRD), Auger<br />

electron spectroscopy (AES), X-ray photoelectron<br />

spectroscopy (XPS), electron spin resonance<br />

(ESR), X-ray fluorescence (XRF), extended X-ray<br />

absorption fine structure (EXFAS), Fourier transform


154 T.S.N. Sankara Narayanan<br />

infra red spectroscopy (FTIR), Raman spectroscopy,<br />

differential thermal analysis (DTA), differential scanning<br />

calorimetry (DSC), secondary ion mass spectrometry<br />

(SIMS), atomic force microscopy (AFM),<br />

glow discharge optical emission spectrometry<br />

(GDOES), quartz crystal impedance system (QCIS),<br />

<strong>conversion</strong> electron Mossbauer spectrometry<br />

(CEMS), acoustic emission (AE) testing, etc. were<br />

used to characterize <strong>phosphate</strong> <strong>coatings</strong> [308-340].<br />

SEM is the most commonly and widely used<br />

technique for characterizing <strong>phosphate</strong> <strong>coatings</strong>. It<br />

is used to determine the morphology and crystal<br />

size of <strong>phosphate</strong> <strong>coatings</strong>. SEM serves as an effective<br />

tool to study the nucleation and growth of<br />

the <strong>phosphate</strong> <strong>coatings</strong> and makes evident of the<br />

fact that the initial growth of the <strong>phosphate</strong> coating<br />

is kinetically controlled and at a latter stage it tends<br />

towards mass transport control [322]. SEM also<br />

substantiates the fact that preconditioning the substrate<br />

before phosphating enables the formation of<br />

a fine-grained and adherent <strong>phosphate</strong> coating.<br />

EPMA is used in the determination of the porosity<br />

of <strong>phosphate</strong> <strong>coatings</strong> in which the number of<br />

copper spots deposited in the pores following immersion<br />

in copper sulphate solution (pH 5.0) [323].<br />

XRD is primarily used to detect the phase constituents<br />

present in <strong>phosphate</strong> <strong>coatings</strong> <strong>–</strong> the<br />

phosphophyllite and hopeite phases in zinc <strong>phosphate</strong><br />

coating. Though the manganese and nickel<br />

modification of zinc <strong>phosphate</strong> coating reveals only<br />

the presence of hopeite phase, there observed to<br />

be significant variations in the orientations of the<br />

hopeite crystals as evidenced <strong>by</strong> the relative intensities<br />

of the H(311), H(241) and H(220), H (040)<br />

peaks. XRD is also used to characterize <strong>phosphate</strong><br />

<strong>coatings</strong> in terms their ‘P’ ratio.<br />

Van Ooij et al. [324] applied high resolution Auger<br />

electron spectroscopy (AES) in combination with<br />

energy dispersive X-ray spectrometry (EDX) to study<br />

the nature of chromium post passivation treatment.<br />

According to them, Cr(III) was not detected at the<br />

metal/<strong>phosphate</strong> boundry but on the <strong>surface</strong> of <strong>phosphate</strong><br />

crystals and that the Fe/Zn ratio was increased<br />

in the <strong>surface</strong> and sub<strong>surface</strong> layers.<br />

XPS is used to detect the nature of various species<br />

in <strong>phosphate</strong> coating. The presence of fatty<br />

acid-like contaminants on the metal substrates can<br />

be identified from the C 1s spectra. Similarly, the<br />

Zn 2p 3/2 and P 2p spectra enable identification of<br />

Zn 3 (PO 4 ) 2 . Besides, the formation of ZnO or Zn(OH) 2<br />

and NaHPO 4 could be identified from the Zn 2p 3/2<br />

and P 2p spectra, respectively. The Fe 2p 3/2 will give<br />

an idea about the presence of Fe 2 O 3 and FePO 4 in<br />

the coating [187]. Cu 2p spectra obtained from the<br />

sample <strong>phosphate</strong>d using a solution containing 1<br />

ppm of Cu 2+ ions show two components at binding<br />

energies 933.6 eV and 932.2 eV [325]. The former<br />

signifies Cu 2+ , although the latter could arise from<br />

either Cu + or Cu metal. All evidence supports the<br />

Cu deposition occurring during the phosphating process.<br />

ESR is used to confirm the modified structure of<br />

hopeite films formed on the <strong>surface</strong> of pre-treated<br />

steel sheets. The zinc <strong>phosphate</strong> film formed from<br />

the bath that does not contain manganese or nickel<br />

ions exhibits no ESR signal. This is because ESR<br />

could only detect paramagnetic transition metal ions<br />

with an unpaired electron whereas in zinc <strong>phosphate</strong><br />

coating the ten electron spins of zinc (II) metal ions<br />

(3d 10 ) in the ‘d’ orbitals are paired with one another.<br />

However, ESR detects the manganese and nickel<br />

components in the modified zinc <strong>phosphate</strong> coating<br />

and proves that manganese and nickel exist as<br />

Mn(II) and Ni(II) in these <strong>coatings</strong> [316].<br />

XRF is used to determine the nature of nickel in<br />

nickel modified zinc <strong>phosphate</strong> coating. The XRF<br />

spectrum obtained for metallic nickel is compared<br />

with that of the modified zinc <strong>phosphate</strong> coating.<br />

The Ni L α peak of metallic nickel occurs at 34.18°<br />

whereas the Ni L α peak of the nickel component of<br />

modified <strong>phosphate</strong> <strong>coatings</strong> occurs at 34.06°, indicating<br />

that the nickel component in the films is<br />

not in the metallic state [316].<br />

EXFAS was used to assess the crystal structure<br />

of manganese modified zinc <strong>phosphate</strong> coating<br />

[318]. The radial distributions of the first<br />

neighbouring atoms appeared at a distance of 0.146<br />

nm for unmodified hopeite and at 0.144 nm for manganese<br />

modified hopeite. This decrease is due to<br />

the disorderness of modified zinc <strong>phosphate</strong> coating<br />

following the introduction of manganese. EXFAS<br />

study also confirms that the manganese component<br />

is substituted for zinc component in the octahedral<br />

structure.<br />

SIMS was used to detect the presence of titanium,<br />

which are adsorbed on to the metal <strong>surface</strong><br />

and act as sites for crystal nucleation [313].<br />

The crystal structure of hopeite and phosphophyllite<br />

are hydrated. So once heated, dehydration<br />

reactions are expected to occur. Differential thermal<br />

analysis (DTA) performed on <strong>phosphate</strong> <strong>coatings</strong><br />

shows that there is a clear disparity in the dehydration<br />

process of these two phases as evidenced<br />

<strong>by</strong> a 50 °C difference in temperature in the endothermic<br />

peaks corresponding to hopeite and<br />

phosphophyllite phases. In the case of nickel and


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Fig. 7. Thermal behaviour of zinc <strong>phosphate</strong> crystals<br />

as hopeite and phosphophyllite studied <strong>by</strong> differential<br />

thermal analyzer: Curve A: Unmodifed<br />

hopeite; Curve B: Hopeite with 1.8 wt.% Ni + Mn;<br />

Curve C: Hopeite with 5.6 wt.% Ni + Mn; Curve D:<br />

Phosphophyllite) (Reprinted from Surface & Coatings<br />

Technology, Vol. 30, N. Sato, Effects of heavy<br />

metal additions and crystal modification on the zinc<br />

phosphating of electrogalvanized steel sheet, pp.<br />

171-181 (1987) with permission from Elsevier Science).<br />

manganese modified zinc <strong>phosphate</strong> <strong>coatings</strong> the<br />

temperature of the endothermic peaks shifts to higher<br />

temperatures and becomes equivalent to the endothermic<br />

peak of phosphophyllite (Fig. 7) [326].<br />

DSC is used to assess the thermal behavior of<br />

<strong>phosphate</strong> <strong>coatings</strong>. Manganese and nickel modified<br />

zinc <strong>phosphate</strong> <strong>coatings</strong> formed on steel and<br />

electrozinc coated steel, in the as deposited condition,<br />

show strong endothermic peaks at temperatures<br />

below 175 °C associated with the loss of water<br />

molecules from the <strong>phosphate</strong>. On the other hand<br />

the DSC traces obtained for these <strong>phosphate</strong> <strong>coatings</strong><br />

after stripping the cured paint film exhibit no<br />

strong endothermic peak over the temperature range<br />

up to 200 °C, confirming that under paint stoving<br />

conditions the more labile water molecules have<br />

been stripped from the <strong>phosphate</strong> coating (Fig. 8)<br />

[313].<br />

155<br />

Fig. 8. DSC curves of nickel and manganese modified<br />

zinc <strong>phosphate</strong> coating formed on (a) steel and<br />

(b) electrozinc coated steel. Solid line represents<br />

before stoving and broken line represents after<br />

stoving and removal of paint. (Source: R.A.<br />

Choudhery and C.J. Vance in Advances in Corrosion<br />

Protection <strong>by</strong> Organic Coatings, D. Scantelbury<br />

and M. Kendig (Eds.), The Electrochemical Society,<br />

NJ, 1989, p.64. Reproduced <strong>by</strong> permission of<br />

The Electrochemical Society, Inc.).<br />

DSC also proves the ability of nickel and manganese<br />

modified zinc <strong>phosphate</strong> coating in resisting<br />

the rehydration compared to unmodified zinc<br />

<strong>phosphate</strong> coating. Unmodified zinc <strong>phosphate</strong> <strong>coatings</strong><br />

obtained on zinc coated steel exhibit two endothermic<br />

peaks at 103 and 135 °C with a total heat<br />

flow of 100 J/g. After exposure to 100% RH for 63<br />

hours, these endothermic peaks appear again, with<br />

a slight shift in peak temperature towards higher<br />

temperatures, indicating that the unmodified zinc<br />

<strong>phosphate</strong> coating is completely rehydrated (Fig.<br />

9). In contrast, manganese and nickel modified zinc<br />

<strong>phosphate</strong> coating exhibit a much stronger resistance<br />

to the redhydration process when subjected


156 T.S.N. Sankara Narayanan<br />

Fig. 9. DSC curves of unmodified zinc <strong>phosphate</strong><br />

coating (a) before stoving; and (b) after stoving and<br />

rehydrated at 100%RH for 63 hours. (Source: R.A.<br />

Choudhery and C.J. Vance in Advances in Corrosion<br />

Protection <strong>by</strong> Organic Coatings, D. Scantelbury<br />

and M. Kendig (Eds.), The Electrochemical Society,<br />

NJ, 1989, p.64. Reproduced <strong>by</strong> permission of<br />

The Electrochemical Society, Inc.).<br />

to similar conditions (Fig. 10) [313]. The main limitation<br />

of DSC is that the <strong>phosphate</strong> coating has to<br />

be physically removed from the substrate.<br />

Handke has demonstrated the use of FTIR to<br />

study the mechanism of <strong>phosphate</strong> coating formation<br />

[327]. FTIR is sufficiently sensitive to distinguish<br />

between different types of <strong>coatings</strong> in terms<br />

of composition and crystallinity [313]. FTIR is also<br />

sensitive to the degree of hydration of <strong>phosphate</strong><br />

<strong>coatings</strong>. The FTIR spectra of nickel and manganese<br />

modified zinc <strong>phosphate</strong> coating on cold rolled<br />

steel (CRS) and zinc coated steel (ZCS) under conditions<br />

namely, before heating, after stoving at 180<br />

°C for 20 minutes and after exposure to 100% RH<br />

for 14 days was analyzed (Figs. 11 and 12). The<br />

assignments of the FTIR spectra for CRS and ZCS<br />

Fig. 10. DSC curves for nickel and manganese<br />

modified zinc <strong>phosphate</strong> coating. Solid line represents<br />

before stoving and broken line represents after<br />

stoving and subsequent rehydration. (Source:<br />

R.A. Choudhery and C.J. Vance in Advances in<br />

Corrosion Protection <strong>by</strong> Organic Coatings, D.<br />

Scantelbury and M. Kendig (Eds.), The Electrochemical<br />

Society, NJ, 1989, p.64. Reproduced <strong>by</strong><br />

permission of The Electrochemical Society, Inc.).<br />

are given in Tables 8 and 9, respectively. The FTIR<br />

spectra of <strong>phosphate</strong> <strong>coatings</strong> on CRS and ZCS<br />

are similar before heating. On heating there are significant<br />

changes in the FTIR spectra of ZCS while<br />

the changes for steel are less noticeable. On exposing<br />

to 100% RH it is evident that <strong>phosphate</strong> <strong>coatings</strong><br />

formed on ZCS shows greater tendency to rehydrate,<br />

indicated <strong>by</strong> the increase in intensity of all<br />

peaks associated with water vibrations. On the other<br />

hand <strong>phosphate</strong> coating formed on CRS exhibit very<br />

little tendency to rehydrate; the spectrum essentially<br />

remains constant.<br />

The infrared and Raman spectra of α-hopeite and<br />

phosphophyllite are shown in Figs. 13-15 [341]. They<br />

resemble the room temperature spectra reported <strong>by</strong><br />

Hill and Jones [342] and Hill [343]. The infrared and<br />

Raman spectra of both compounds display a large<br />

number of bands. In addition, the infrared and Raman<br />

spectra of these compounds resemble each other<br />

very strongly. Hence distinction of the respective<br />

peaks in <strong>phosphate</strong> <strong>coatings</strong> is not easy. However,<br />

there are some peaks, which are different for these<br />

compounds, are useful in identifying these compounds.<br />

For example the Raman bands at 1150,<br />

1055 and 310 cm -1 are typical for α-hopeite whereas<br />

the Raman bands at 1135, 1070 and 118 cm -1 are<br />

useful for identifying phosphophyllite in <strong>phosphate</strong><br />

coating.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Table 8. The assignments of the FTIR spectra of nickel and manganese modified zinc <strong>phosphate</strong> coating<br />

on cold rolled steel (after Choudhery and Vance [313]).<br />

Wavenumber Shape/Intensity Assignment<br />

(cm -1 ) Before heating After stoving After exposure<br />

for 20 minutes at 180 °C to 100%<br />

RH for 14 days<br />

3536 Sharp/Strong Free O-H Lost All peaks<br />

3214 V. broad/Strong OH/H 2 O stretching Broadens similar to stoved<br />

1643 Sharp/Medium-St H 2 O bending Shifts to 1629 sample with<br />

1145 W. Sharp/Strong P-O stretching Shoulder 1170 only a slight<br />

appears increase in peak<br />

and shoulder at height<br />

1082 now a peak<br />

1121 Sharp/Strong P-O stretching Shoulder 1170<br />

appears and<br />

shoulder at 1082<br />

now a peak<br />

1080 Sh./Strong P-O stretching Shoulder 1170<br />

appears and<br />

shoulder at 1082<br />

now a peak<br />

1020 Sharp/Strong P-O stretching Shoulder 1170<br />

appears and<br />

shoulder at 1082<br />

now a peak<br />

997 Sharp/Strong H 2 O rocking Lost<br />

933 Sharp/Strong P-O stretching <strong>–</strong><br />

643 Sharp/Medium O-P-O bending/ Broadens<br />

H 2 O waging<br />

The influence of metal components in hopeite<br />

films was investigated using Infrared and Raman<br />

spectra [243]. The IR spectra of hopeite films exhibit<br />

peaks at 3500 to 3000, 1640, 1200 to 900 and<br />

640 cm-1 . The peaks at 3500 to 3000 cm-1 correspond<br />

to the stretching vibration of O-H in H O, the<br />

2<br />

peak at 1640 cm-1 to the deformation of H-O-H in<br />

H O and the three peaks at 1200 to 900 and that at<br />

2<br />

640 cm-1 3- to PO in the hopeite films. There are four<br />

4<br />

3- basic vibrations for PO , namely, ν1 , ν , ν and ν in<br />

4<br />

2 3 4<br />

which ν and ν are inactive in IR absorption whereas<br />

1 2<br />

ν and ν are active. Hence the three peaks at 1200<br />

3 4<br />

to 900 cm-1 correspond to the ν mode and that at<br />

3<br />

640 cm-1 to the ν mode. A comparison of the IR<br />

4<br />

spectra of nickel and manganese modified hopeite<br />

with unmodified hopeite indicate that the incorporation<br />

of nickel and manganese into the hopeite films<br />

157<br />

3- affects the coordination state of PO and cause<br />

4<br />

splitting or shifting of peaks. The three peaks at<br />

1200 to 900 cm-1 corresponding to the ν mode of<br />

3<br />

hopeite film shows splitting and the peak at higher<br />

wavenumber is shifted <strong>by</strong> 10 cm-1 towards the lower<br />

wavenumber. The ν peak of nickel and manganese<br />

4<br />

modified hopeite generates a new peak at 580<br />

cm-1 .<br />

Laser Raman spectra of hopeite films exhibit four<br />

peaks in the region between 1150 and 930 cm-1 ,<br />

with the main peak appearing at 996 cm-1 . These<br />

3- four peaks are due to PO , which has four basic<br />

4<br />

vibration modes ν , ν , ν and ν , all of them are<br />

1 2 3 4<br />

3- Raman active. If the PO in the hopeite film exists<br />

4<br />

as a perfect regular tetrahedron, then only peaks of<br />

ν and ν should appear. But in practice, there are<br />

1 3<br />

3- four peaks. PO in hopeite films affects the sym-<br />

4


158 T.S.N. Sankara Narayanan<br />

Table 9. The assignments of the FTIR spectra of nickel and manganese modified zinc <strong>phosphate</strong> coating<br />

on zinc coated steel (after choudhery and vance [313]).<br />

Wavenumber Shape/Intensity Assignment<br />

(cm -1 ) Before heating After stoving After exposure<br />

for 20 minutes at 180 °C to 100%<br />

RH for 14 days<br />

3545 Sharp/Strong Free O-H Lost Water stretching<br />

3290 V. broad/Strong OH/H 2 O stretching Broadens peaks intensify.<br />

1635 Sharp/Strong H 2 O bending Shifts to 1649 Spectrum begins<br />

1135 Sh./Strong P-O stretching Shoulder 1184 to resemble the<br />

appears and one obtained for<br />

shoulder at 1026 sample before<br />

now a peak heating<br />

1098 Sharp/Strong P-O stretching Shoulder 1184<br />

appears and<br />

shoulder at 1026<br />

now a peak<br />

1069 Sharp/Strong P-O stretching Shoulder 1184<br />

appears and<br />

shoulder at 1026<br />

now a peak<br />

1026 Sh./Strong P-O stretching Shoulder 1184<br />

appears and<br />

shoulder at 1026<br />

now a peak<br />

1003 Sharp/Strong H 2 O rocking Lost<br />

925 Sharp/Strong P-O stretching <strong>–</strong><br />

638 Sharp/Medium O-P-O bending/ Broadens<br />

H 2 O waging<br />

metry of the regular tetrahedron structure <strong>by</strong> interaction<br />

with the surrounding crystalline structure so<br />

that the symmetry will become distorted. As a result,<br />

the degeneracy of the vibration modes will be<br />

united and split and hence results in four peaks.<br />

Three peaks, except for the main peak, correspond<br />

to the stretching vibration mode of ν , which<br />

3<br />

split into three peaks; the main peak around 900<br />

cm-1 corresponds to the symmetrical stretching vi-<br />

3- bration of ν . The level of interaction of PO with its<br />

1 4<br />

surroundings affects the number of basic vibrations,<br />

the Raman band or the intensity of the Raman spectrum.<br />

With the incorporation of manganese or nickel<br />

in the hopeite films, the Raman band of the main<br />

peak as well as the three peaks of ν are shifted to<br />

3<br />

a lower wavelength. Increasing the metal content in<br />

hopeite films has an increasing effect on the data of<br />

Raman band. Raman spectra showed a sensitivity<br />

to the degree of modification of hopeite films <strong>by</strong> metal<br />

components to a greater extent than did IR spectra.<br />

AFM is used to study the effect of <strong>surface</strong> preconditioning<br />

on phosphatability of zinc coated steel.<br />

AFM is useful in getting a better understanding of<br />

the activation process [328].<br />

Quartz crystal impedance system (QCIS) was<br />

used to study the formation of <strong>phosphate</strong> coating<br />

from a manganese modified low-zinc phosphating<br />

bath [329]. It involves rapid and simultaneous measurement<br />

of admittance spectra of the zinc coated<br />

piezoelectric quartz crystal resonator (PQC). By<br />

measuring the equivalent circuit parameters and frequency<br />

shift of the zinc coated PQC resonator, the<br />

growth kinetics of the phosphating process can be


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Fig. 11. FTIR Spectra of nickel and manganese modified zinc <strong>phosphate</strong> coating formed on steel. (a) Full<br />

range; and (b) Selected range. The top, middle and bottom spectra represent before stoving, after stoving<br />

and after stoving with subsequent rehydration, respectively. (Source: R.A. Choudhery and C.J. Vance in<br />

Advances in Corrosion Protection <strong>by</strong> Organic Coatings, D. Scantelbury and M. Kendig (Eds.), The Electrochemical<br />

Society, NJ, 1989, p.64. Reproduced <strong>by</strong> permission of The Electrochemical Society, Inc.).<br />

Fig. 12. FTIR Spectra of nickel and manganese modified zinc <strong>phosphate</strong> coating formed on zinc. (a) Full<br />

range; and (b) Selected range. The top, middle and bottom spectra represent before stoving, after stoving<br />

and after stoving with subsequent rehydration, respectively. (Source: R.A. Choudhery and C.J. Vance in<br />

Advances in Corrosion Protection <strong>by</strong> Organic Coatings, D. Scantelbury and M. Kendig (Eds.), The Electrochemical<br />

Society, NJ, 1989, p.64. Reproduced <strong>by</strong> permission of The Electrochemical Society, Inc.).<br />

159


160 T.S.N. Sankara Narayanan<br />

Fig. 13. Infrared spectra of α-hopeite and phosphophyllite<br />

in the region 4000<strong>–</strong>1400 cm -1 (Reprinted<br />

from Spectrochimica Acta Part A, Vol. 57, O.<br />

Pawlig, V. Schellenschlager, H.D. Lutz and R.<br />

Trettin, Vibrational analysis of iron and zinc <strong>phosphate</strong><br />

<strong>conversion</strong> coating constituents, pp. 581-590<br />

(2001) with permission from Elsevier Science).<br />

monitored. QCIS is also used to measure the change<br />

in viscoelasticity of zinc <strong>phosphate</strong> <strong>coatings</strong>, which<br />

decreases with increase in phosphating time and<br />

the concentration of sodium nitrite [330].<br />

The utility of GDOES in the analysis of <strong>phosphate</strong><br />

<strong>coatings</strong> has been attempted <strong>by</strong> several authors<br />

[331-334] and <strong>phosphate</strong> <strong>coatings</strong> deposited<br />

on cold-rolled, hot-dipped galvanized, galvanneled,<br />

electrogalvanized and Zn-Ni or Zn-Fe electroalloycoated<br />

sheets were characterized. It has been reported<br />

[331] that the profiles of the elemental distribution<br />

throughout the film thickness can be obtained<br />

with high sensitivity. GDOES provides more detailed<br />

information on the constituent elements, depth, di-<br />

Fig. 14. Infrared spectra of α-hopeite and phosphophyllite<br />

in the region 1200<strong>–</strong>100 cm -1 (Reprinted<br />

from Spectrochimica Acta Part A, Vol. 57, O.<br />

Pawlig, V. Schellenschlager, H.D. Lutz and R.<br />

Trettin, Vibrational analysis of iron and zinc <strong>phosphate</strong><br />

<strong>conversion</strong> coating constituents, pp. 581-590<br />

(2001) with permission from Elsevier Science).<br />

rection and the crystal growth process than conventional<br />

chemical and X-ray methods. Quantification<br />

of the results has also been attempted in the<br />

case of GDOES based on the integrated intensity<br />

of the constituent elements. Maeda et al. [331] have<br />

suggested the possibility of utilizing this method for<br />

quantitative determination, after establishing a linear<br />

relationship between integrated intensity and<br />

coating thickness of phosphorus and nickel.<br />

GDOES is considered to be the most appropriate<br />

method for depth profiling thick <strong>phosphate</strong> <strong>coatings</strong><br />

[335]. GDOES due to its higher detection sensitivity<br />

compared to XPS, confirms the adsorption of<br />

titanium <strong>phosphate</strong> on zinc coated steel [328].


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Fig. 15. Raman spectra of α-hopeite and phosphophyllite<br />

in the region 1300<strong>–</strong>100 cm -1 (Reprinted<br />

from Spectrochimica Acta Part A, Vo. 57, O. Pawlig,<br />

V. Schellenschlager, H.D. Lutz and R. Trettin, Vibrational<br />

analysis of iron and zinc <strong>phosphate</strong> <strong>conversion</strong><br />

coating constituents, pp. 581-590 (2001) with<br />

permission from Elsevier Science).<br />

AE is also used as a tool to characterize <strong>phosphate</strong><br />

<strong>coatings</strong> [336-340]. Phosphate coating exhibit<br />

only a single peak during AE in a ring-down<br />

count-rate against strain curves. The amplitude distribution<br />

verses events curve recorded for steel/<strong>phosphate</strong><br />

systems at a strain rate of 10 percent showed<br />

emissions with amplitudes only below 40 dB. Rooum<br />

and Rawlings [338, 339] have shown that it is possible<br />

to assign the source of failure mechanism to<br />

the peaks of amplitude distribution. Each failure<br />

mechanism was characterized <strong>by</strong> events centred<br />

on specific amplitudes. In their study they attribute<br />

the peak at 22 dB to the cracking of the hopeite<br />

needles and the peak at 26 dB to the adhesion fail-<br />

161<br />

ure between the phosphophyllite and hopeite. Hence<br />

it is evident that with the help of AE testing it is<br />

possible to detect the onset of a process such as<br />

cracking of the coating and/or adhesion failure and<br />

the extent to which these failures shall occur.<br />

3.11. Testing the quality of <strong>phosphate</strong><br />

<strong>coatings</strong><br />

Methods to evaluate the quality of <strong>phosphate</strong> <strong>coatings</strong><br />

involve the determination of its physical characteristics<br />

as well as the performance in corrosive<br />

environments. Kwiatkowski et al. [344] have <strong>review</strong>ed<br />

the testing methods of <strong>phosphate</strong> <strong>coatings</strong>.<br />

3.11.1. Evaluation of physical characteristics<br />

(a) Examination of physical appearance. Phosphate<br />

films on steel may range in colour from light gray<br />

to dark gray, depending on the type of bath and<br />

the grade of steel substrate used. After sufficiently<br />

strong scratching of <strong>phosphate</strong> coating<br />

with a fingernail, a white scratch should appear<br />

on the <strong>surface</strong> without causing an injury to the<br />

coating visible to the naked eye.<br />

(b) Determination of coating thickness and coating<br />

weight. The local thickness of <strong>phosphate</strong> <strong>coatings</strong><br />

on steel is usually determined <strong>by</strong> magnetic,<br />

electromagnetic and microscopic methods<br />

[18,19,127,128]. The average thickness is usually<br />

expressed in g/m 2 or mg/ft 2 . Destructive<br />

methods of determination of coating weight are<br />

widely adopted. These methods involve the determination<br />

of change in weight of a coated specimen<br />

after dissolution of the coating in a suitable<br />

medium. Concentrated hydrochloric acid containing<br />

20 g/l of antimony trioxide is usually used at<br />

room temperature for this purpose. Other solutions<br />

commonly used are 5% solution of chromic<br />

acid at room temperature and 20% sodium<br />

hydroxide at 90 °C. The difference in weight after<br />

coating removal is divided <strong>by</strong> the <strong>surface</strong> area<br />

of the work in m 2 to obtain unit coating weight in<br />

g/m 2 .<br />

(c) Determination of acid resistance. It is calculated<br />

as the difference in weight per unit area of the<br />

panel before phosphating and after stripping off<br />

the coating; and is expressed in g/m 2 .<br />

(d) Testing of physical properties. The absorption<br />

value and hygroscopicity are the important parameters<br />

relating to the physical properties of<br />

the <strong>phosphate</strong> coating. Evaluation of the absorption<br />

value of the <strong>phosphate</strong> coating involves the<br />

measure of the gain in weight per unit area when<br />

subjected to immersion in diacetone alcohol for


162 T.S.N. Sankara Narayanan<br />

two minutes and allowed to drain the excess for<br />

three minutes [345]. The gain in weight of the<br />

<strong>phosphate</strong>d panels, when subjected to a humid<br />

atmosphere in a closed container at room temperature<br />

for six hours, gave the hygroscopicity<br />

of the coating [345].<br />

(e) Estimation of porosity. Estimation of the porosity<br />

of the <strong>phosphate</strong> coating involves both chemical<br />

and electrochemical methods. The chemical<br />

(Ferroxyl indicator) method was based on<br />

formation of blue spots (Prussian blue) on a filter<br />

paper dipped in potassium ferricyanide - sodium<br />

chloride - gelatin mixture when applied over<br />

the <strong>phosphate</strong>d <strong>surface</strong> for one minute. The number<br />

of blue spots per sq. cm gives a measure of<br />

the porosity of the coating. The electrochemical<br />

method of determination was based on the measurement<br />

of the oxygen reduction current density<br />

when immersed in air-saturated sodium hydroxide<br />

solution (pH 12). The current density<br />

values measured at -550 mV, where oxygen reduction<br />

is the dominant reaction at the uncoated<br />

areas reveals the porosity of the coating<br />

[209,210]. Although, both the Ferroxyl test and<br />

the electrochemical test are used in an industrial<br />

scale to assess the porosity of <strong>phosphate</strong>d<br />

steel, the former method is found to be qualitative<br />

and not very effective in distinguishing the<br />

porosities of panels <strong>phosphate</strong>d using different<br />

baths. In contrast, the electrochemical method<br />

is far reliable besides its simplicity in operation<br />

and quicker measurements of the porosity. The<br />

advantage of the electrochemical method has<br />

been discussed elsewhere [209, 346].<br />

(f) Determination of thermal and chemical stabilities.<br />

The thermal stability of the coating was usually<br />

determined <strong>by</strong> calculating the percentage<br />

loss in weight when the <strong>phosphate</strong>d panels were<br />

subjected to drying at 120 and 180 °C.<br />

The chemical stability of the coating, in particular,<br />

the alkaline stability is very important as it<br />

determines the effectiveness of the <strong>phosphate</strong><br />

coating as a base for cathodic electrophoretic<br />

finishes. This is determined <strong>by</strong> calculating the<br />

percentage residual coating when immersed in<br />

alkaline media. Immersion in sodium hydroxide<br />

solution is recommended to test the alkaline<br />

stability. Recently, Kwiatkowski et al. [347] have<br />

recommended a borate buffer solution containing<br />

0.01M EDTA as the medium to test the alkaline<br />

stability of the coating and proved its usefulness<br />

in predicting the same. A similar calculation<br />

of the percentage residual coating when<br />

subjected to immersion treatments in buffered<br />

solutions of varying pH from 2-14 can give an<br />

insight about ability of the <strong>phosphate</strong> <strong>coatings</strong><br />

to withstand different chemical aggressions and<br />

to prove their effectiveness in preventing the cosmetic<br />

corrosion [219].<br />

(g) Evaluation of <strong>surface</strong> morphology. The <strong>surface</strong><br />

morphology of the <strong>phosphate</strong> coating is usually<br />

assessed <strong>by</strong> scanning electron microscope<br />

(SEM). This technique reveals the distinct features<br />

of the crystal structure, grain size of the<br />

crystallites, the coating coverage and uniformity;<br />

the parameters that determine the performance<br />

of the coating.<br />

(h) Determination of ‘P ratio’. The ‘P ratio’ is usually<br />

determined <strong>by</strong> measuring the intensity of the<br />

characteristic planes of the constituent crystallites<br />

of the <strong>phosphate</strong> coating. In a zinc <strong>phosphate</strong><br />

coating, which consists essentially of<br />

hopeite (Zn (PO ) . 4H2O) and phosphophyllite<br />

3 4 2<br />

(Zn Fe(PO ) . 4H2O), the ‘P ratio’ is defined as<br />

2 4 2<br />

[199]:<br />

‘P ratio’ = P<br />

P + H<br />

Where P <strong>–</strong> Intensity of the characteristic planes<br />

of phosphophyllite and H <strong>–</strong> Intensity of the<br />

characteristic planes of hopeite.<br />

(i) Adhesion measurements. The standard laboratory<br />

method of estimation of adhesion is the peel<br />

off test, which involves the determination of the<br />

extent of adhesion at scribed areas using a pressure<br />

sensitive adhesive tape. Usually adhesion<br />

in the dry state will be good since it mainly depends<br />

on the cohesive failure of the paint film.<br />

Wet adhesion is of prime importance and is usually<br />

determined after subjecting the painted panels<br />

to immersion treatment in de-ionized water<br />

at 45 °C for 240 hours. Depending upon the extent<br />

of peeling, rating will be made between 0<br />

and 5B as per ASTM D 3359-87 [348].<br />

3.11.2. Evaluation of corrosion performance<br />

(a) Rapid preliminary quality control tests. Preliminary<br />

investigation of coating quality for on-line<br />

monitoring mainly involves two empirical tests<br />

of which one involves inspection for rust spots<br />

after immersion in 3% sodium chloride solution<br />

for 5-30 minutes and the other is concerned with<br />

the time taken for metallic copper deposition from<br />

a copper sulphate-sodium chloride-hydrochloric<br />

acid mixture.<br />

(b) Laboratory corrosion resistance tests. The most<br />

frequently used laboratory tests for evaluating


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

the corrosion resistance of <strong>phosphate</strong> <strong>coatings</strong><br />

include (i) immersion test, (ii) salt spray (fog)<br />

test, (iii) humidity test and (iv) A.R.E. salt droplet<br />

test.<br />

(i) Immersion test. This test consists of determining<br />

the time required for the first appearance of<br />

corrosion on the basis metal when immersed in<br />

3% sodium chloride solution. The change in<br />

weight expressed as g/m 2 for every 24-hour period<br />

of immersion is also determined and correlated<br />

to the corrosion resistance of the coating.<br />

(ii) Salt spray test. Evaluation of corrosion resistance<br />

of the <strong>phosphate</strong>d and finished panels is<br />

performed <strong>by</strong> subjecting them to a salt mist of<br />

5% sodium chloride solution in a salt spray<br />

chamber for a specified length of time. The extent<br />

of spread of corrosion from a scribe made<br />

on the panel, rated after ASTM B 117-85 specifications,<br />

is a measure of the corrosion resistance<br />

of the <strong>phosphate</strong> coating [349].<br />

(iii) Humidity test. This test is used for the evaluation<br />

of <strong>phosphate</strong>d panels with and without further<br />

finishes. By subjecting the <strong>phosphate</strong>d panels<br />

to highly humid conditions (90-95% relative<br />

humidity) at slightly elevated temperatures (42-<br />

48 o C), assessment of corrosion likely to be induced<br />

due to the porosity of the coating can be<br />

made. The extent of blistering of paints due to<br />

the presence of soluble salts on <strong>phosphate</strong>d and<br />

finished panels when subjected for 1000 hours<br />

in the humidity chamber is also related to its<br />

corrosion performance.<br />

Fig. 16. Cyclic voltammogram obtained from zinc <strong>phosphate</strong> coated steel in 5 wt.% sodium chloride solution<br />

(pH 6.5; 25 °C) (After Kiss and Coll-Palagos [355]).<br />

163<br />

(iv) A.R.E. salt droplet test. This test consists of<br />

the evaluation of the corrosion resistance of the<br />

<strong>phosphate</strong>d panels <strong>by</strong> determining the loss in<br />

weight after five days of exposure in humid condition<br />

inside a closed cabinet at room temperature<br />

with a single spray of synthetic seawater<br />

on each day [350].<br />

(c) Electrochemical methods of testing. The electrochemical<br />

methods of testing the <strong>phosphate</strong><br />

coating mainly involves the anodic polarization<br />

studies in 0.6M ammonium nitrate [346,351] and<br />

AC impedance measurements in 3% sodium<br />

chloride solution [352-363].<br />

Zurilla and Hospadaruk [209] proposed a method<br />

to assess the porosity of <strong>phosphate</strong>d steel based<br />

on oxygen reduction current density in air saturated<br />

0.1 N sodium hydroxide solution. Since cathodic<br />

reduction of oxygen occurs at the uncoated areas,<br />

the measure of current density values at a fixed<br />

potential (<strong>–</strong>550 mV vs. SCE) reveal the porosity of<br />

the coating. Kiss and Coll-Palagos [355] utilized<br />

cyclic voltammetry to evaluate the porosity of <strong>phosphate</strong><br />

<strong>coatings</strong>. Fig. 16 shows the typical cyclic<br />

voltammogram obtained for zinc <strong>phosphate</strong> coated<br />

steel in 5% sodium chloride solution (pH 6.5; 25<br />

°C). The voltammogram can be classified into three<br />

major regions based on the potential values. The<br />

anodic peak observed between <strong>–</strong> 550 mV and <strong>–</strong> 950<br />

mV is believed to be due to the oxidation of Fe, Fe 2+<br />

and/or the formation of a complex compound in the<br />

pores of the <strong>phosphate</strong>d steel and its intensity is


164 T.S.N. Sankara Narayanan<br />

Fig. 17. Anodic polarization curve obtained for zinc <strong>phosphate</strong> coated steel in 0.6 M ammonium nitrate.<br />

related to the porosity of the <strong>phosphate</strong> coating.<br />

Kiss and Coll-Palagos [355] termed this peak as<br />

‘porosity peak’. The peak height at <strong>–</strong>800 mV or the<br />

integrated area under this anodic peak was used to<br />

predict the porosity.<br />

Anodic polarization study in 0.6 M ammonium<br />

nitrate solution is one of the recommended methods<br />

of electrochemical evaluation of corrosion resistance<br />

of <strong>phosphate</strong> <strong>coatings</strong> (Fig. 17) [346]. During<br />

anodic polarization in 0.6 M ammonium nitrate<br />

solution, at potentials more negative than <strong>–</strong>0.33 V,<br />

<strong>phosphate</strong>d steel undergoes active dissolution.<br />

Above <strong>–</strong>0.33 V, the first passivation region occurs<br />

due to the adsorption of hydroxide ions at the electrode<br />

<strong>surface</strong>. The occurrence of the second active<br />

region is due to the replacement of hydroxide ions<br />

<strong>by</strong> <strong>phosphate</strong> ions available at the electrode/solution<br />

interface. Replacement of the adsorbed <strong>phosphate</strong><br />

ions <strong>by</strong> nitrate causes the occurrence of the<br />

second passive region. Hence it is clear that these<br />

active and passive regions are the result of the competitive<br />

and potential dependent adsorption of anions<br />

at the electrode <strong>surface</strong>. It should be noted<br />

here that the appearance of this second current<br />

density maximum is specific to <strong>phosphate</strong>d steel<br />

and it is not observed for uncoated steel when tested<br />

under similar conditions. Hence the value of the<br />

second current density maximum can be used to<br />

evaluate the corrosion resistance of different <strong>phosphate</strong><br />

<strong>coatings</strong> [346].<br />

The other important method of evaluation of the<br />

corrosion resistance of <strong>phosphate</strong> <strong>coatings</strong> is the<br />

electrochemical impedance spectroscopy (EIS). It<br />

provides a rapid, nondestructive means of evaluating<br />

corrosion rate and mechanism of corrosion of<br />

<strong>phosphate</strong> <strong>coatings</strong>. Literature reports on the evaluation<br />

<strong>phosphate</strong> <strong>coatings</strong> [357-363] suggest that the<br />

corrosion behaviour of <strong>phosphate</strong> <strong>coatings</strong> in contact<br />

with the corrosive medium (3.5% NaCl) can be<br />

explained on the basis of a porous film model since<br />

the electrolyte/coating-metal interface approximates<br />

such a model. Accordingly, the <strong>phosphate</strong>d substrates<br />

are considered as partially blocked electrodes<br />

when comes in contact with 3.5% NaCl solution.<br />

This implies that the metal substrate is corroding<br />

in the same way when unprotected in a much<br />

smaller area where coverage is lacking. Since the<br />

capacitive and resistive contributions vary directly<br />

and indirectly, respectively, with respect to the area,<br />

based on these measured parameters, predictions<br />

on the corrosion rate of different <strong>phosphate</strong> <strong>coatings</strong><br />

can be easily made. Fig. 18 shows the Nyquist<br />

plot obtained for zinc <strong>phosphate</strong> coated steel in 3.5%<br />

NaCl solution exhibiting a semicircle in the high fre-


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

Fig. 18. Nyquist plot obtained for zinc <strong>phosphate</strong> coated steel in 3.5 wt.% sodium chloride solution.<br />

quency region followed <strong>by</strong> a linear portion in the low<br />

frequency region. An equivalent electrical circuit<br />

model, which involves charge transfer resistance<br />

(R ct ), the double layer capacitance (C dl ) and the<br />

Warburg impedance (Z w ) can be proposed to simulate<br />

the behaviour of <strong>phosphate</strong> coating in 3.5% NaCl<br />

solution. The change in the values of these parameters<br />

with time gives an account of how the coating<br />

degradation has occurred and helps to establish a<br />

mechanistic pathway. High values of charge transfer<br />

resistance and low values of double layer capacitance<br />

signify a coating of better performance.<br />

The appearance of Warburg impedance suggests<br />

that corrosion of <strong>phosphate</strong>d steel is a diffusion<br />

controlled. [352-354,362-365].<br />

3.12. Applications<br />

Phosphate <strong>coatings</strong> have been put to a wide variety<br />

of applications; salient among them is the corrosion<br />

protection, as bases for paint, to provide wear<br />

resistance and an aid in cold forming of steel<br />

[18,19,24,127,128,366,367].<br />

Phosphate <strong>coatings</strong> provide an effective physical<br />

barrier to protect corrosion-prone metals against<br />

their environment. Due to their insulating nature, the<br />

<strong>phosphate</strong> <strong>coatings</strong> prevent the onset and spreading<br />

of corrosion. These <strong>coatings</strong> provide effective<br />

corrosion protection to ferrous and non-ferrous met-<br />

165<br />

als alike. The extent to which these <strong>coatings</strong> provide<br />

corrosion resistance is dependent on the thickness<br />

and weight of the coating used. However, better<br />

corrosion protection can be achieved <strong>by</strong> finishing<br />

these <strong>coatings</strong> with paints, oils, etc. Phosphate<br />

<strong>coatings</strong> provide an effective base for the application<br />

of paints and this constitutes their most widespread<br />

application [368]. They can be used as an<br />

excellent base for more recent methods of paint<br />

applications such as electrophoretic painting and<br />

powder coating [369,370] and shown to improve the<br />

corrosion resistance of steel coated subsequently<br />

with cadmium, zinc, nickel, etc., both in industrial<br />

as well as marine atmospheres [371]. The application<br />

of <strong>phosphate</strong> coating also improves the adhesive<br />

bonding of plain carbon steels [372].<br />

The mechanism of pyrite oxidation in acidic media<br />

involves preferential release of iron into the medium.<br />

The <strong>phosphate</strong> treatment results in the formation of<br />

iron <strong>phosphate</strong> coating and significantly reduces the<br />

chemical oxidation of pyrite [373]. The formation of<br />

a <strong>phosphate</strong> layer on the <strong>surface</strong> of iron powders<br />

induces a significant improvement of the oxidation<br />

resistance in the temperature range of 300-700°C<br />

[374]. The oxidation resistance of ultrafine copper<br />

powder is increases <strong>by</strong> phosphating [375]. Rebeyrat<br />

et al. [376] based on thermogravimetric experiments<br />

suggest that phosphating of bulk α-iron significantly<br />

decreases the gain in weight due to oxidation.


166 T.S.N. Sankara Narayanan<br />

The Nd<strong>–</strong>Fe<strong>–</strong>B type magnets are highly susceptible<br />

to attack from both climatic and corrosive environments,<br />

which result in corrosion of the alloy and<br />

deterioration in both its physical and magnetic properties.<br />

The poor corrosion resistance of Nd<strong>–</strong>Fe<strong>–</strong>B<br />

type magnets in many aggressive environments is<br />

associated with the presence of ~ 35 wt.% of neodymium<br />

in their composition. Neodymium, along with<br />

other rare earth (RE) elements belongs among the<br />

most electrochemically active metals. The application<br />

of a zinc <strong>phosphate</strong> coating proved to be a most<br />

useful method to improve the corrosion resistance<br />

of NdFeB magnet [377,378]. The improvement in<br />

corrosion resistance of NdFeB magnet following the<br />

application of a zinc <strong>phosphate</strong> coating is shown in<br />

Fig. 19.<br />

Phosphating is a widely used method of reducing<br />

wear on machine elements and moving parts<br />

[379,380]. Phosphate <strong>coatings</strong> function as lubricants,<br />

in addition their ability to retain oils and soaps<br />

further enhances this action. Heavy manganese<br />

<strong>phosphate</strong> <strong>coatings</strong>, supplemented with proper lubricants<br />

are most commonly used for wear resistance<br />

applications [381]. The manganese <strong>phosphate</strong>s<br />

widely used in automotive industry are the<br />

best to improve the ease of sliding and the reduction<br />

of associated wear of two steel <strong>surface</strong>s sliding<br />

one against the other. The <strong>phosphate</strong> <strong>coatings</strong> pos-<br />

sess no intrinsic lubricating properties but can absorb<br />

or hold a considerable quantity of lubricant <strong>by</strong><br />

virtue of their porosity [382,383]. This combination<br />

favours an easier running-in at higher <strong>surface</strong> pressures<br />

<strong>by</strong> forming a non-metallic barrier that separates<br />

the two metal <strong>surface</strong>s and reduces the danger<br />

of seizure and associated pitting. There is also<br />

less noise produced at such <strong>surface</strong>s and they have<br />

an in-built capacity, in emergencies, to run dry for a<br />

limited period [384]. Phosphating increases the sliding<br />

distance to scuffing as well as the scuffing load,<br />

whilst marginally reducing the coefficient of friction.<br />

No advantage was found in phosphating dry sliding<br />

<strong>surface</strong>s. Phosphating reduces the likelihood of<br />

adhesive wear in marginal or poorly lubricated sliding<br />

couples. The choice of <strong>phosphate</strong> coating is<br />

primarily dependent on the <strong>surface</strong> finish of the sliding<br />

counterface; thin <strong>coatings</strong> are suitable for smooth<br />

<strong>surface</strong>s whereas against rougher <strong>surface</strong>s thicker<br />

<strong>coatings</strong> are preferred [385].<br />

The power used in deep drawing operations, sets<br />

up a great amount of friction between the steel <strong>surface</strong><br />

and the die. This will decrease the speed of<br />

drawing operation and the service life of tools and<br />

dies [386,387]. Application of light to medium weight<br />

non-metallic zinc <strong>phosphate</strong> coating to steel <strong>surface</strong>s,<br />

which permit the distribution and retention of<br />

a uniform film of lubricant over the entire <strong>surface</strong>,<br />

Fig. 19. Potentiodynamic polarization curves of NdFeB magnet in 3.5% sodium chloride solution (a) Uncoated<br />

NdFeB magnet; and (b) Zinc <strong>phosphate</strong> coated NdFeB magnet.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

prevents metal to metal contact and makes possible<br />

the cold forming and extrusion of more difficult<br />

shapes, than is possible without the coating<br />

[85,388-390]. A combination of zinc <strong>phosphate</strong> and<br />

lubricant film prevents welding and scratching of steel<br />

in drawing operations and greatly decreases the<br />

rejections. Bustamante et al. [391] suggest the<br />

usefulness of pre-<strong>phosphate</strong> coating of zinc and zinc<br />

alloy coated steel sheet in preventing the damage<br />

during the forming process.<br />

Phosphate coating is used as an absorbent coating<br />

in laser <strong>surface</strong> hardening of steel [392-394].<br />

Although, all the three major type of <strong>phosphate</strong> <strong>coatings</strong>,<br />

namely, the manganese, zinc and iron <strong>phosphate</strong><br />

<strong>coatings</strong> were used as absorbent <strong>coatings</strong>,<br />

the former was found to be frequently employed in<br />

majority of the cases. Besides these, <strong>phosphate</strong><br />

<strong>coatings</strong> have also been found to be useful as thermal<br />

control <strong>coatings</strong> in satellite components [395].<br />

3.13. Environmental impact<br />

The environmental impact of metal finishing operations<br />

is a matter of serious concern and <strong>phosphate</strong><br />

<strong>pretreatment</strong> operation is no exception to this. The<br />

wastes generated from <strong>phosphate</strong> <strong>pretreatment</strong> line<br />

can be broadly classified into liquid wastes arising<br />

from acid pickling, alkaline cleaning, rinsing stages<br />

and chromic acid sealing stage, etc. and solid<br />

wastes arising from the sludges formed during phosphating.<br />

It is estimated that the total outlet of wastewater<br />

is about 20,800 g/m 2 and the solid residue<br />

generated in the whole phosphating process is<br />

about 49.0 g/m 2 . The phosphating process discharges<br />

400 g/m 2 CO 2- equivalent mass to the environment,<br />

which demonstrates that conventional<br />

phosphating, has some green house effect. Fortunately,<br />

the influence on human health is not so heavy<br />

since the average local toxic level (LTL) is estimated<br />

to be 1.6 ppm/ppm [396]. Unlike the wastewater<br />

discharge, the environment cannot absorb the solid<br />

residue. The phosphating sludges are generally considered<br />

as hazardous waste materials and are,<br />

therefore, subject to strict regulations as to disposal.<br />

It is estimated that fifty million pounds of sludge<br />

results from phosphating operations each year.<br />

Hence it is not only desirable but also mandatory to<br />

identify a means for recovering various sludge components.<br />

Phosphating sludge, in general, has 20 wt.% iron,<br />

10 wt.% zinc, 1-3 wt.% manganese,


168 T.S.N. Sankara Narayanan<br />

ing the particle size of the dried sludge to less than<br />

about 20 mesh. The dried and ground <strong>phosphate</strong><br />

sludge has been found to be an excellent lubricant<br />

additive, which is suitable for use in lubricant formulations<br />

designed for the metal treatment, metal forming<br />

and industrial lubrication.<br />

United States Patent 4986977 [402] suggests a<br />

method for treating the phosphating sludge with an<br />

aqueous base to achieve a pH greater than 10 that<br />

results in precipitation of iron hydroxide. The iron<br />

hydroxide is recovered and the aqueous phase is<br />

acidified to a pH of 7-10 to cause precipitation of<br />

zinc hydroxide.<br />

Baldy [403] suggests that the phosphating<br />

sludge generated from automotive <strong>pretreatment</strong> process<br />

may contain up to 25 wt.% of oil and removal<br />

of oil is essential for the full recovery of all useful<br />

components. The removal of oil is accomplished <strong>by</strong><br />

the addition of H 2 O 2 to an acidic dispersion of the<br />

sludge that displaces the oil from the remaining<br />

mixture. According to him, the recovery of zinc <strong>phosphate</strong><br />

sludge involves the following four stages:<br />

Phase I is the separation of oil from the sludge.<br />

Phase II is the extraction of zinc, manganese and<br />

nickel from the sludge following digestion with phosphoric<br />

acid. Phase III is the <strong>conversion</strong> of the iron<br />

<strong>phosphate</strong> residue from phase II to pigment grade<br />

iron oxide and sodium <strong>phosphate</strong> that may be acidified<br />

to an iron <strong>phosphate</strong> concentrate.<br />

The use of phosphating sludge in the process of<br />

clinker production is suggested as one of the possible<br />

mode of reclamation of such waste <strong>by</strong> Caponero<br />

and Tenorio [404]. Their study proves that an addition<br />

of up to 7.0% of phosphating sludge to the raw<br />

cement meal of Portland cement did not cause any<br />

damage to the clinkerization process. X-ray diffraction<br />

analysis shows that there is no significant modification<br />

in the yielded clinker proportional to the<br />

sludge additions, neither is there atypical phases<br />

formed with additions up to 5.0% of phosphating<br />

sludge. Differential thermal analysis of the mixture<br />

with up to 7.0% dry sludge additions does not show<br />

any significant difference from the analysis of the<br />

cement clinker raw meal. Additions of the phosphating<br />

sludge up to 5.0% significantly modify only the<br />

zinc content of the clinker that was produced. The<br />

major element of the sludge, zinc, shows an average<br />

of incorporation of 75%.<br />

4. SUMMARY<br />

This <strong>review</strong> outlines the various aspects of phosphating.<br />

Although, numerous modifications were<br />

proposed recently, on the deposition technologies<br />

to achieve different types of <strong>coatings</strong> and desirable<br />

properties such as improved corrosion resistance,<br />

wear resistance, etc., <strong>phosphate</strong> <strong>conversion</strong> coating<br />

still plays a vital part in the automobile, process<br />

and appliance industries, as it has unique advantages<br />

in the cost-wise placement among all the<br />

emerging deposition technologies and pays off the<br />

finisher with a handful of profit.<br />

ACKNOWLEDGEMENT<br />

Financial support given <strong>by</strong> the Council of Scientific<br />

and Industrial Research, New Delhi, India, is gratefully<br />

acknowledged. The author expresses his sincere<br />

thanks to Dr.S. Srikanth, Scientist-in-charge,<br />

NML Madras Centre and Prof. S.P. Mehrotra, Director,<br />

National Metallurgical Laboratory,<br />

Jamshedpur, for their constant support and encouragement.<br />

REFERENCES<br />

[1] M.G. Fontana, Corrosion Engineering, 3 rd<br />

Edition (McGraw-Hill Book<br />

Company,Singapore, 1987).<br />

[2] U.R. Evans, An Introduction to Metallic Corrosion,<br />

3 rd Edition (Edward Arnold Publishers<br />

Ltd., London, 1981).<br />

[3] H.H. Uhlig, Corrosion and Corrosion Control,<br />

2 nd Edition (John Wiley & sons, Inc., New<br />

York, 1971).<br />

[4] F.L. La Que, In: Good Painting Practice, ed.<br />

<strong>by</strong> John D. Keane, Vol. 1, 2 nd Edition, (Steel<br />

Structures Painting Council, Pittsburgh, 1973)<br />

Chap. 1.1, p. 3.<br />

[5] Henry Leidheiser, Jr., In: Metals Handbook,<br />

Vol. 13, 9 th Edition (American Society of<br />

Metals, Ohio, 1987) p. 377.<br />

[6] J.W. Gailen and E.J. Vaughan, Protective<br />

Coatings for Metals (Charles Griffin & Co. Ltd.,<br />

1979) p. 97.<br />

[7] M.A. Kuehner // Met. Finish. 83 (1985) 17.<br />

[8] Ichiro Suzuki, Corrosion Resistant Coatings<br />

Technology (Marcel Dekker, Inc., New York,<br />

1989) p. 167.<br />

[9] Phosphorous and its Compounds, Vol. II, ed.<br />

<strong>by</strong> J.R. Van Wazer (Interscience Publishers<br />

Inc., New York, 1967) p. 1869.<br />

[10] W.A. Ross, British Patent 3,119 (1869).<br />

[11] T.W. Coslett, British Patent 8,667 (1906).<br />

[12] T.W. Coslett, British Patent 15,628 (1908).<br />

[13] T.W. Coslett, British Patent 22,743 (1909).<br />

[14] T.W. Coslett, British Patent 28,131 (1909).<br />

[15] R.G. Richard, British Patent 37,563 (1911).


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

[16] W.H. Allen, U.S. Patent 1,206,075 (1916).<br />

[17] W.H. Allen, U.S. Patent 1,311,726 (1914).<br />

[18] D.B. Freeman, Phosphating and Metal<br />

Pretreatment - A Guide to Modern Processes<br />

and Practice (Industrial Press Inc., New<br />

York, 1986).<br />

[19] W. Rausch, The Phosphating of Metals<br />

(Finishing Publications Ltd., London, 1990).<br />

[20] V.M. Darsey and R.R. Tanner, U.S. Patent<br />

1,887,967 (1932).<br />

[21] R.R. Tanner and H.J. Lodeesen, U.S. Patent<br />

1,911,726 (1933).<br />

[22] Fritz Singer, German Patent 673,405 (1939).<br />

[23] The Pyrene Co. Ltd., British Patent 473,285<br />

(1937).<br />

[24] The Pyrene Co. Ltd., British Patent 517,049<br />

(1940).<br />

[25] W. Machu, Die Phosphatierung (Verlag-<br />

Chemie, Weinnharin, 1950).<br />

[26] J.S. Thompson, U.S. Patents 2,234,206<br />

(1941); 2,312,855 (1943).<br />

[27] G.W. Jernstedt, U.S. Patent 2,310,239<br />

(1943).<br />

[28] M.B. Roosa // Lubrication Engineering 6<br />

(1950) 117.<br />

[29] A.H. Jenkins and D.B. Freeman, British<br />

Patent 866,377 (1961).<br />

[30] D.B. Freeman // Prod. Finish. (London) 29<br />

(1976) 19.<br />

[31] W.C. Jones, U.S. Patent 4,140,551 (1979).<br />

[32] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// J. Electrochem. Soc. India 41 (1992) 27.<br />

[33] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Surf. Coat. Technol. 43/44 (1990) 543.<br />

[34] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Bull. Electrochem. 6 (1990) 920.<br />

[35] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Surf. Coat. Intl. (JOCCA) 74 (1991) 222.<br />

[36] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Met. Finish. 89 (1991) 39.<br />

[37] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Trans. Inst. Met. Finish. 70(2) (1992) 81.<br />

[38] T.S.N. Sankara Narayanan and M.<br />

Subbaiyan, In: Surface Engineering- Fundamentals<br />

of Coatings, ed. <strong>by</strong> P.K. Datta and<br />

J.S. Gray (Royal Society of Chemistry,<br />

London, 1993), Vol. 1, p. 132.<br />

[39] T.S.N. Sankara Narayanan, M. Panjatcharam<br />

and M. Subbaiyan // Met. Finish. 91 (1993)<br />

65.<br />

[40] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Trans. Inst. Met. Finish. 71 (1993) 52.<br />

169<br />

[41] T.S.N. Sankara Narayanan and M. Subbaiyan<br />

// Prod. Finish.(London) 45 (1992) 9.<br />

[42] W. Mc Lead, D.V. Subrahmanyam and G.R.<br />

Hoey // Electrodep. Surf. Treat. 3 (1975) 335.<br />

[43] R.D. Wyvill, Proceedings of the Conference<br />

on Finish’83, 12.1-12.15, 1983.<br />

[44] Y. Matsushima, S. Tanaka and A. Niizuma,<br />

U.S. Patent 4,063,968 (1977).<br />

[45] Y. Matsushima, N. Oda and H. Terada, U.S.<br />

Patent 4,220,486, 1980.<br />

[46] M.A. Kuehner, SAE. Automobile Engineering<br />

Congress, Detroit, Paper No. 740099<br />

(1974).<br />

[47] P. Burden, British Patent 1,198,546 (1970).<br />

[48] J.K. Howell // Plating 60 (1973) 1033.<br />

[49] Y. Ayano, K. Yashiro and A. Niizuma, U.S.<br />

Patent 4,153,479 (1979).<br />

[50] H.A. Jenkins and D.B. Freeman, German<br />

Patent 1,203,087 (1965).<br />

[51] A.H. Jenkins and D.B. Freeman, French<br />

Patent 1,243,081 (1960).<br />

[52] E. Wyszomirski, German Patent 1,187,101<br />

(1965).<br />

[53] H.A. Jenkins and D.B. Freeman, German<br />

Patent 1,208,599 (1966).<br />

[54 G. Muller, W. Rausch and W. Wuttke, European<br />

Patent 0,121,274 (1984).<br />

[55] G.D. Howell and R.F. Ayres, U.S. Patent<br />

3,579,389 (1971).<br />

[56] W.N. Jones and J.W. Ellis, U.S. Patent<br />

3,515,600 (1970).<br />

[57] W.N. Jones and J.W. Ellis, French Patent<br />

1,538,274 (1968).<br />

[58] W.S. Russel, German Patent 1,072,055<br />

(1959).<br />

[59] E. Mayer and H. Rogner, German Patent<br />

1,095,624 (1960).<br />

[60] Rudolf Brodt, German Patent 1,090,048<br />

(1960).<br />

[61] Gerhard Collardin GmbH, British Patent<br />

903,151 (1962).<br />

[62] D.J. Mueller, U.S. Patent 4,451,301 (1984).<br />

[63] K. Goltz, U.S. Patent 4,427,459 (1984).<br />

[64] C. Ries and M. Prymak, U.S. Patent<br />

3,810,792 (1974).<br />

[65] P.G. Chamberlain // Met. Finish. Abs. 3<br />

(1961) 54.<br />

[66] W. Rausch, H.Y. Oei, H.J. Edler and H. Liebl,<br />

U.S. Patent 3,516,875 (1970).<br />

[67] A. Askienazy, V. Ken and J.C. Souchet, U.S.<br />

Patent 4,089,708 (1978).<br />

[68] K.J. Woods, U.S. Patent 4,086,103 (1978).<br />

[69] A.J. Hamilton, U.S. Patent 4,149,909 (1979).


170 T.S.N. Sankara Narayanan<br />

[70] H.Y. Oei and S. Moller, U.S. Patent<br />

3,723,192 (1973).<br />

[71] W.A. Vittands and W.M. McGowan, U.S.<br />

Patent 4,168,983 (1979).<br />

[72] Pyrene Chemical Services Ltd., British<br />

Patent 1,421,386 (1976).<br />

[73] J.I. Maurer, U.S. Patent 3,723,334 (1973).<br />

[74] T.C. Atkiss and W.E. Keen, Jr., U.S. Patent<br />

4,057,440 (1977).<br />

[75] C.T. Snee, U.S. Patent 3,676,224 (1972).<br />

[76] S. Rolf and N. Reinhard, German Patent<br />

2,516,459 (1976).<br />

[77] Fritz Schaefer, German Patent 1,109,987<br />

(1961).<br />

[78] Y. Yoshida and A. Takagi, British Patent<br />

1,425,871 (1976); U.S. Patent 3,961,991<br />

(1976).<br />

[79] R. Bennett // Polym. Paint Col. J. 174 (1984)<br />

878.<br />

[80] A. Nicholson, German Patent 1,144,991<br />

(1963).<br />

[81] H. Bertsch, F. Stork and L. Schaefer, German<br />

Patent 1,289,714 (1969).<br />

[82] M. Brock and B.A. Cooke, U.S. Patent<br />

4,052,232 (1977).<br />

[83] M. Brock and B.A. Cooke, U.S. Patent<br />

4,147,567 (1979).<br />

[84] A. Stoch, Cz. Paluszkiewicz and E. Dlugon<br />

// J. Mol. Struct., 511<strong>–</strong>512 (1999) 295.<br />

[85] V. Burokas, A. Martusiene and G. Bikulcius<br />

// Surf. Coat. Technol. 102 (1998) 233.<br />

[86] V.D. Shah, French Patent 1,487,183 (1967).<br />

[87] J.I. Maurer, R.E. Palmer and V.D. Shah, U.S.<br />

Patent 3,222,226 (1965).<br />

[88] L. Schiffman, U.S. Patent 3,063,877 (1962).<br />

[89] E. Jung, V. Vossius and G.W. Coldewey,<br />

German Patent 1,939,302 (1970).<br />

[90] P. Ajenherc, G. Garnier and G. Lorin, European<br />

Patent 0,154,384 (1985).<br />

[91] K. Goltz and W.A. Blum, British Patent<br />

1,417,376 (1975).<br />

[92] K. Goltz and W.A. Blum, U.S. Patent<br />

3,864,175 (1975).<br />

[93] G. Schneider, U.S. Patent 3,647,569 (1972).<br />

[94] W. Herbst and H. Ludwig, U.S. Patent<br />

3,272,662 (1966).<br />

[95] W.S. Carey and D.W. Reichgott, U.S. Patent<br />

4,917,737 (1990).<br />

[96] W. Herbst, F. Rochlitz and H. Vilcsek,<br />

German Patent 1,207,760 (1965).<br />

[97] W. Herbst, F. Rochlitz and H. Vilcsek, U.S.<br />

Patent 3,200,004 (1965).<br />

[98] E. Duch, W. Herbst, F. Rochlitz, H. Scherer<br />

and H. Vilcsek, U.S. Patent 3,202,534<br />

(1965).<br />

[99] R.A. Ashdown, U.S. Patent 3,493,440<br />

(1970).<br />

[100] R.A. Ashdown, British Patent 1,050,860<br />

(1966).<br />

[101] W.O. Crawford, Army Weapons Command<br />

Rock Island, Tech. Rep. Jan. 1966; Met.<br />

Finish. Abs., 8 (1966) 137.<br />

[102] J.V. Otrhalek, R.M. Ajluni and G.S. Gomes,<br />

U.S. Patent 4,124,414 (1978).<br />

[103] J.V. Otrhalek, R.M. Ajluni and G.S. Gomes,<br />

U.S. Patent 4,182,637 (1980).<br />

[104] B. Da Fonte Jr., U.S. Patent 4,359,347<br />

(1982).<br />

[105] D.J. Melotik, U.S. Patent 3,720,547 (1973).<br />

[106] G.D. Kent, Paint Conf.’87, Chicago, Paper<br />

25-1, March, 1987.<br />

[107] D.J. Guhde, U.S. Patent 3,877,998 (1975).<br />

[108] Y. Miyazaki, M. Suzuki and H. Kaneko,<br />

U.S. Patent 4,180,406 (1979).<br />

[109] L. Kulick and J.K. Howell, Jr., U.S. Patent<br />

4,039,353 (1977).<br />

[110] B.S. Tuttle, Swiss Patent 362,290 (1962).<br />

[111] B.S. Tuttle, W.A. Vitlands and O.L. Walsch,<br />

German Patent 1,170,751 (1964).<br />

[112] B.S. Tuttle and W.A. Vitlands, Brit. Patent<br />

1,012,267 (1965).<br />

[113] J.N. Tuttle Jr. and O.P. Jaboin, U.S. Patent<br />

4,673,445 (1987).<br />

[114] J. Schapira, V. Ken and J. Dauptain, U.S.<br />

Patent 4,497,666 (1985).<br />

[115] J. Schapira, V. Ken and J.L. Dauptain,<br />

European Patent 0,085,626 (1982)<br />

[116] K. Yashiro and Y. Moriya, U.S. Patent<br />

4,341,558 (1982).<br />

[117] R. Murakami, Y. Mino and K. Saito, European<br />

Patent 0,061,911 (1982); U.K. Patent<br />

Appl. 2,097,429 (1982).<br />

[118] H.R. Charles and D.L. Miles, U.S. Patent<br />

3,819,423 (1974).<br />

[119] J.P. Jones, U.S. Patent 4,362,577 (1982).<br />

[120] R.C. Gray, M.J. Pawlik, P.J. Prucnal and<br />

C.J. Baldy, U.S. Patent 5,294,265 (1994).<br />

[121] N. Das and J.P. Jandrists, U.S. Patent<br />

6,027,579 (2000).<br />

[122] W. Wichelhaus, H. Endres, K.H. Gottwald,<br />

H.D. Speckmann and J.W. Brouwer, U.S.<br />

Patent 6,090,224 (2000).<br />

[123] W. Wichelhaus, H. Endres, K.H. Gottwald,<br />

H.D. Speckmann and J.W. Brouwer,<br />

U.S.Patent 6,395,105 (2002).


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

[124] G. Buttner, M. Kimpel and K. Klein, U.S.<br />

Patent 5,773,090 (1998).<br />

[125] R. Opitz, K. Hosemann and H. Portz, U.S<br />

Patent 4,600,447 (1986).<br />

[126] J.N. Tuttle Jr. and O.P. Jaboin, U.S. Patent<br />

4,673,445 (1987).<br />

[127] Guy Lorin, Phosphating of Metals (Finishing<br />

Publications Ltd., London, 1974).<br />

[128] C. Rajagopal and K.I. Vasu, Conversion<br />

Coatings: A Reference for Phosphating,<br />

Chromating and Anodizing (Tata McGraw-<br />

Hill Publishing Company Ltd., New Delhi,<br />

2000).<br />

[129] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Plat. Surf. Finish. 80<br />

(1993) 72.<br />

[130] T.S.N. Sankara Narayanan // Met. Finish. 94<br />

(1996) 40.<br />

[131] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Prod. Finish. (London) 45<br />

(1992) 6.<br />

[132] American Chemical Paint Co., British<br />

Patent, 501,739 (1939).<br />

[133] Pyrene Co. Ltd., British Patent, 551,261<br />

(1943).<br />

[134] D. James and D.B. Freeman // Trans. Inst.<br />

Met. Finish. 49 (1971) 79.<br />

[135] I.G. Farbenindustrie Aktiengesellschaft,<br />

French Patent, 801,033 (1936).<br />

[136] R.J. Kahn, British Patent, 507,355 (1939).<br />

[137] Soc. Continentale Parker, Belgium Patent,<br />

443,128 (1942).<br />

[138] Societe Continentale Parker, French Patent,<br />

849,856 (1939); Pyrene Company Ltd.,<br />

British Patent 510,684 (1939)<br />

[139] D.R. Vonk and J.A. Greene, U.S. Patent<br />

5,588,989 (1996).<br />

[140] B. Mayer, P. Kuhm, P. Balboni, M. Senner,<br />

H.D. Speckmann, J. Geke, J.W. Brouwer<br />

and A. Willer, U.S. Patent 6,379,474 (2002).<br />

[141] R.C. Gibson, U.S. Patent 2,301,209 (1942).<br />

[142] J.V. Laukonis, In: Interface Conversion for<br />

Polymer Coatings, ed. <strong>by</strong> P. Weiss and<br />

G.D. Cheever (American Elsevier Publishing<br />

Company Inc., New York, 1968) p.182.<br />

[143] The Pyrene Co. Ltd., British Patent,<br />

561,504 (1944).<br />

[144] T.W. Coslett, British Patent, 16,300 (1909).<br />

[145] E.L. Ghali and J.R. Potvin // Corros. Sci. 12<br />

(1972) 583.<br />

[146] K.A. Akanni, C.P.S. Johal and D.R. Gabe //<br />

Trans. Inst. Met. Finish. 62 (1984) 64.<br />

171<br />

[147] K.A. Akanni, C.P.S. Johal and D.R. Gabe //<br />

Met. Finish. 83(4) (1985) 41.<br />

[148] B. Zantout and D.R. Gabe // Trans. Inst.<br />

Met. Finish. 61 (1983) 88.<br />

[149] K.S. Rajagopalan, B. Dhandapani and<br />

A. Jayaraman , In: Proceedings of the 3 rd<br />

International Congress on Metallic Corrosion<br />

(1966) Vol. 1, p. 365.<br />

[150] B. Dandapani, A. Jayaraman and K.S.<br />

Rajagopalan, British Patent 1,090,743<br />

(1967).<br />

[151] C. Rajagopal, B. Dandapani, A. Jeyaraman<br />

and K.S. Rajagopalan, U.S. Patent<br />

3,586,612 (1971).<br />

[152] K.S. Rajagopalan, C. Rajagopal,<br />

N. Krithivasan, M. Tajudeen and M.E.<br />

Kochu Janaki // Werkstoffe und Korrosion<br />

23 (1971) 347.<br />

[153] F. Zucchi and G. Tranbelli // Corros. Sci. 11<br />

(1971) 141.<br />

[154] K.S. Rajagopalan, C. Rajagopal,<br />

N. Krithivasan, M. Tajudeen and M.E.<br />

Kochu Janaki // Met. Finish. J. 19 (1973)<br />

188.<br />

[155] K.S. Rajagopalan and R. Srinivasan // Sheet<br />

Met. Indus. 55 (1978) 642.<br />

[156] P. Bala Srinivasan, S. Sathyanarayanan,<br />

C. Marikannu and K. Balakrishnan //<br />

Surf.Coat. Technol. 64 (1994) 161.<br />

[157] K. Ravichandran, Harihar Sivanandh,<br />

S. Ganesh, T. Hariharasudan and T.S.N.<br />

Sankara Narayanan // Met. Finish. 98(9)<br />

(2000) 48.<br />

[158] K. Ravichandran, Deepa Balasubramanium,<br />

J. Srividya, S. Poornima and T.S.N.<br />

Sankara Narayanan // Trans. Met. Finish.<br />

Asso. India 9 (2000) 205.<br />

[159] K. Ravichandran and T.S.N. Sankara<br />

Narayanan //Trans. Inst. Met. Finish. 79<br />

(2001) 143.<br />

[160] N.J. Bjerrum, E. Christensen and<br />

T. Steenberg, U.S. Patent 6,346,186 (2002).<br />

[161] P.K. Sinha and R. Feser // Surf. Coat.<br />

Technol. 161 (2002) 158.<br />

[162] M. Gebhardt, In: Proceedings of the Conference<br />

on “Surface ’66" (1966) p. 323.<br />

[163] J.B. Lakeman, D.R. Gabe and M.O.W.<br />

Richardson // Trans. Inst. Met. Finish. 55<br />

(1977) 47.<br />

[164] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Surf. Coat. Intl. (JOCCA)<br />

75(5) (1992) 184.


172 T.S.N. Sankara Narayanan<br />

[165] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Trans. Met. Finish. Assoc.<br />

India 1 (1992) 9.<br />

[166] T.S.N. Sankara Narayanan // Met. Finish. 91<br />

(1993) 57.<br />

[167] T.S.N. Sankara Narayanan // Plat. Surf.<br />

Finish. 83 (1996) 69.<br />

[168] T.S.N. Sankara Narayanan //Prod.<br />

Finish.(London), 48(4) (1995) 16.<br />

[169] T.S.N. Sankara Narayanan // Prod. Finish.<br />

(London) 49(2) (1996) 31.<br />

[170] T.S.N. Sankara Narayanan // Prod. Finish.<br />

(London) 50(10) (1997) 4.<br />

[171] Samuel Spring, Metal Cleaning (Reinhold<br />

Publishing Corporation, New York, 1963).<br />

[172] William P. Kripps, In: Metals Handbook<br />

(American Society of Materials, Ohio,<br />

1987), Vol. 13, 9 th Edition, p. 380.<br />

[173] A.J. Leibman, In: Steel Structures Painting<br />

Manual, ed. <strong>by</strong> J. Bigos (Steel Structures<br />

Painting Council, Pittsburgh, 1954), Vol.1,<br />

p. 6.<br />

[174] A.G. Roberts, Organic Coatings, Properties,<br />

Selection and Use (U.S. Dept. of<br />

Commerce, NB of Standards, 1968), p. 105.<br />

[175] M. Straschill, Modern Practice in the<br />

Pickling of Metals (Robert Draper,<br />

Teddington, Middlesex, England, 1968).<br />

[176] W. Bullough // Metallurgical Reviews 2<br />

(1957) 391.<br />

[177] J.F. Andrew and P.D. Donovan // Trans. Inst.<br />

Met. Finish. 48 (1970) 152; 49 (1971) 162.<br />

[178] R.D. Wyvill and T. Cape // Prod. Finish.<br />

(Cincinnati) 52(1) (1987) 68.<br />

[179] S. Sankara Pandian and S. Guruviah // Met.<br />

Finish. 88(2) (1990) 73.<br />

[180] A. Zavarella, U.S. Patent 2,476,345 (1949).<br />

[181] M. Green, US Patent 2086712 (1937); US<br />

Patent 2082950 (1937).<br />

[182] F.P. Spruance Jr., U.S. Patent, 2,438,877<br />

(1948).<br />

[183] V.S. Lapatukhin, Phosphating of Metals<br />

(Mashgiz, Moscow, 1958).<br />

[184] R.D. Wyvill // Prod. Finish. (Cincinnati) 49(2)<br />

(1984) 50.<br />

[185] R. Murakami, H. Shimizu, T. Yoshii,<br />

M. Ishida and H. Yonekura, U.S. Patent<br />

4,419,147 (1983).<br />

[186] H.L. Pinkerton, In: Electroplating Engineering<br />

Handbook (Reinhold Publishing<br />

Corporation, New York, 1962), p. 705.<br />

[187] S. Maeda and M.Yamamoto // Prog. Org.<br />

Coat. 33 (1998) 83.<br />

[188] Samuel Spring, Preparation of Metals for<br />

Painting (Reinhold Publishing Corporation,<br />

New York, 1965), p. 204.<br />

[189] G.D. Cheever // J. Paint Technol. 39(504)<br />

(1967) 1.<br />

[190] R.P. Wenz and J.J. Claus // Mater. Perform.<br />

(12) (1981) 7.<br />

[191] T.S.N. Sankara Narayanan // Prod. Finish.<br />

(London) 50(4) (1997) 18.<br />

[192] A. Neuhaus and M. Gebhardt // Werkstoff<br />

und Korrosion 17 (1966) 493.<br />

[193] A. Neuhaus, E. Jumpertz and M. Gebhardt<br />

// Korros. 16 (1963) 155.<br />

[194] P. Chamberlain and S. Eisler // Met. Finish.<br />

50(6) (1952) 113.<br />

[195] G.D. Cheever, SAE Technical paper No.<br />

700460 (1970).<br />

[196] W. Machu, In: Interface Conversion for<br />

Polymer Coatings, ed. <strong>by</strong> P. Weiss and<br />

G.D. Cheever (American Elsevier Publishing<br />

Company Inc., New York, 1968), p.128.<br />

[197] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan, In: Proceedings of the<br />

National Conference on Modern Analytical<br />

Techniques in Materials Science (Indian<br />

Society for Analytical Scientists, Mumbai,<br />

1991), p. 15.<br />

[198] G.D. Cheever, In: Interface Conversion for<br />

Polymer Coatings, ed. <strong>by</strong> P. Weiss and<br />

G.D. Cheever (American Elsevier Publishing<br />

Co. Inc., New York, 1968), p. 150.<br />

[199] T. Miyawaki, H. Okita, S. Umehara and<br />

M. Okabe, In: Proceedings of the conference<br />

on Interfinish ‘80 (Kyoto, 1980).<br />

[200] M.O.W. Richardson, D.B. Freeman,<br />

K. Brown and N. Djaroud // Trans. Inst. Met.<br />

Finish. 61 (1983) 183.<br />

[201] F. Kaysser // Galvanotechnik (Saulgau)<br />

63(11) (1972) 11.<br />

[202] G.D. Cheever // J.Coat Technol. 50 (1978)<br />

78.<br />

[203] G.D. Cheever // Am. Paint Coat. J. 62(14)<br />

(1977) 64.<br />

[204] Tony Mansour // Mater. Eval. 41 (1983) 302.<br />

[205] C. T. Yap, T. L. Tan, L. M. Gan and H. W. K.<br />

Ong // Appl. Surf. Sci. 27 (1986) 247.<br />

[206] T.S.N. Sankara Narayanan and M.<br />

Subbaiyan // Prod. Finish. (London) 45(9)<br />

(1993) 7.<br />

[207] V. Hospadaruk, J. Huff, R.W. Zurilla and H.T.<br />

Greenwood, National SAE meeting (Detroit,<br />

March, 1978), Paper No. 780186


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

[208] G.D. Cheever // J. Paint Technol. 41 (1969)<br />

289.<br />

[209] R.W. Zurilla and V. Hospadaruk // SAE<br />

Transactions (1978) Paper No. 780187.<br />

[210] A. Losch, J.W. Schultze and H.D.<br />

Speckmann // Appl. Surf. Sci. 52 (1991) 29.<br />

[211] T.S.N. Sankara Narayanan // Met. Finish.<br />

91(5) (1993) 51.<br />

[212] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Met. Finish. 91(6) (1993)<br />

89.<br />

[213] T.S.N. Sankara Narayanan // Met. Finish.<br />

92(1) (1994) 31.<br />

[214] T.S.N. Sankara Narayanan // European<br />

Coat. J. (March 1998) 162.<br />

[215] T.S.N. Sankara Narayanan // European<br />

Coat. J. (June 1998) 466.<br />

[216] T. Sugama, L.E. Kukacka, N. Carciello and<br />

J.B. Warren // J. Mater. Sci. 26 (1991)<br />

1045.<br />

[217] P.J. Gardner. I.W. McArn, V. Barton and<br />

G.M. Seydt // J. Oil Colour Chem. Assoc.<br />

(JOCCA) 73 (1990) 16.<br />

[218] N. Sato // SAE Transactions (1990) Paper<br />

No. 900838.<br />

[219] J.P. Servais, B. Schmitz and V. Leroy//<br />

Mater. Perform. 27(11) (1988) 56.<br />

[220] A.J. Sommer and H. Leidheiser, Jr.// Corros.<br />

43 (1987) 661.<br />

[221] R.R. Wiggle, A.G.Smith and J.R.Petrocelli<br />

// J. Paint Technol. 40 (1968) 174.<br />

[222] T.R Roberts, J. Kolts and J.H. Steele, Jr. //<br />

Society of Automotive Engineers Transactions<br />

(1980) Paper No. 891749.<br />

[223] W.J. van Ooij and O.T. De Vries, In: Tenth<br />

International Conference on Organic Coatings<br />

Science and Technology, ed. <strong>by</strong> A.G.<br />

Patsis (Marcel Dekker, Inc., New York<br />

1984).<br />

[224] W.J. Van Ooij and A. Sabata // Surf. Coat.<br />

Technol. 39/40 (1989) 667.<br />

[225] D.D. Davidson, M.L. Stephens, L.E. Soveide<br />

and R.J. Shaffer // SAE Technical Paper<br />

(1986) No. 862006.<br />

[226] T.S.N. Sankara Narayanan // Met. Finish.<br />

94(6) (1996) 86.<br />

[227] R. Kojima, K. Nomura and Y. Ujikara //<br />

J. Japanese Soc. Colour. Mater. 55 (1982)<br />

365.<br />

[228] I. Sugaya and S. Kondo // J. Japanese Soc.<br />

Colour. Mater. 36 (1963) 283.<br />

[229] H. Odashima, M. Kitayama and T. Saito //<br />

SAE Trans. (1986) Paper No. 860115.<br />

173<br />

[230] L. Kwiatkowski, A. Sadkowski and<br />

A. Kozlowski, In: Proceedings of the 11 th<br />

International Congress on Metallic Corrosion<br />

(Italy, 1990), p. 2: 395-2.400.<br />

[231] L. Fedrizzi, A.Tomasi, S. Pedrotti, P.L.<br />

Bonora and P. Balboni // J. Mater. Sci. 24<br />

(1989) 3928.<br />

[232] P.L. Bonora, A. Barbucci, G. Busca,<br />

V. Lorenzelli, E. Miglio and G.G. Ramis, In:<br />

Proceedings of the XXth National Congress<br />

of Inorganic Chemistry (Pavia, Italy,1987).<br />

[233] F. Liebau // Acta Crystallographica 18<br />

(1965) 352.<br />

[234] K. Yamato, T. Ichida and T. Irie // Kawasaki<br />

Steel Technical Report 22 (1990) 57.<br />

[235] N. Sato // J. Met. Finish. Soc. Japan 37<br />

(1986) 758.<br />

[236] N. Sato // Surf. Coat. Technol. 30 (1987)<br />

171.<br />

[237] N. Sato, T. Minami and H. Kono // Surf.<br />

Coat. Technol. 37 (1989) 23.<br />

[238] N. Sato and T. Minami // J. Met. Finish.<br />

Soc. Japan 38 (1987) 108.<br />

[239] N. Sato and T. Minami // J. Chem. Soc.<br />

Japan (1987) 1741.<br />

[240] N. Sato and T. Minami // J. Chem. Soc.<br />

Japan (1988) 1727.<br />

[241] N. Sato and T. Minami // J. Mater. Sci. 24<br />

(1989) 4419.<br />

[242] N. Sato and T. Minami // J. Chem. Soc.<br />

Japan (1988) 1891.<br />

[243] N. Sato and T. Minami // J. Mater. Sci. 24<br />

(1989) 3375.<br />

[244] K. Yamato, T. Honjo, T. Ichida, H. Ishitobi<br />

and M. Kawaki // Kawasaki Steel Technical<br />

Report 12 (1985) 75.<br />

[245] H.J. Kim // Surf. Engg. 14 (1998) 265.<br />

[246] W. Wiederholt, The Chemical Surface<br />

Treatment of Metals (Robert Draper Ltd.,<br />

1965).<br />

[247] R.P. Wenz, In: Organic Coatings Science<br />

and Technology, ed. <strong>by</strong> G.D. Parditt and<br />

A.V. Patsis (Marcell Dekker, Inc., New<br />

York, 1984), Vol. 6, p.373.<br />

[248] G.W. Grossmann // Paint Industry Magazine<br />

7 (1961) 7.<br />

[249]V. Hospadaruk, J. Huff, R.W. Zurialla and<br />

H.T. Greenwood // SAE Technical paper No.<br />

702944 (1977).<br />

[250] J.J. Wojtkowiak and H.S. Bender // J. Coat.<br />

Technol. 50 (1978) 86.


174 T.S.N. Sankara Narayanan<br />

[251] K. Takao, A. Yasuda, S. Kobayashi,<br />

T. Ichida and T. Irie // Trans. Iron Steel Inst.<br />

Japan 72(10) (1985) 258.<br />

[252] K. Takao, A. Yasuda, S. Kobayashi,<br />

T. Ichida and T. Irie // J. Iron Steel Inst.<br />

Japan 72 (1986) 1582.<br />

[253] J.A. Kargol and D.L. Jordan // Corros. 34<br />

(1982) 201.<br />

[254] S. Maeda // Corros. Abs. 22 (1983) 428.<br />

[255] N. Fujino, S. Inenaga, N. Usuki and S.<br />

Wakano // Sumitomo Search 30 (1985) 31.<br />

[256] G. Blumel and A.Vogt // Stahl und Eisen<br />

103(17) (1983) 813.<br />

[257] P. Balboni // Tratt.Finit. 27(9) (1987) 37.<br />

[258] Ph.L. Coduti // Met. Finish. 78(5) (1980) 51.<br />

[259] E.L. Ghali, Doctorate Thesis, University of<br />

Paris, 1968.<br />

[260] C. Beauvais and Y. Bary // Galvano 39(403)<br />

(1970) 625.<br />

[261] O. Ursini // Metallurgia Italiana 80(10) (1988)<br />

765.<br />

[262] O. Ursini, In:Proceedings of the XXII<br />

international Metallurgy Congress<br />

(Associazione Italiana di Metallurgia,<br />

Bologna, Italy, 1988) p. 1227.<br />

[263] E. Radomska, In: Proccedings of the<br />

Corrosion Week (Scientific Society of<br />

Mechanical Engineers, Budapest, Hungary,<br />

1988) p. 447.<br />

[264] G. Jernstedt // Trans. Electrochem. Soc. 83<br />

(1943) 361.<br />

[265] P. Tegehall // Colloids and Surfaces 42<br />

(1989) 155.<br />

[266] P. Tegehall // Colloids and Surfaces 49<br />

(1990) 373.<br />

[267] P. Tegehall // Acta Chem. Scand. 43 (1989)<br />

322.<br />

[268] T. Yamamato, A. Mochizuki, H. Okita,<br />

T. Miyawaki, Y. Shirongane and K. Mori,<br />

U.S.Patent 4,517,030 (1985).<br />

[269] J.J. Donofrio, Belgium Patent 894,432<br />

(1983).<br />

[270] A.J. Hamilton,U.S. Patent Reissue 27,662<br />

(1973).<br />

[271] A.J. Hamilton and G. Schneider, U.S.<br />

Patent 3,741,747 (1973).<br />

[272] A.R. Morrison and H.D. Hermann, U.S.<br />

Patent 3,728,163 (1973).<br />

[273] Westinghouse Electric International Co.,<br />

British Patent 560,847 (1944).<br />

[274] N. Usuki, A. Sakota, S. Wakano and M.<br />

Nishihara // J. Iron Steel Inst. Japan 77(3)<br />

(1991) 398.<br />

[275] T. Hada, S. Naito, K. Ando and M. Yoneno<br />

// J. Surf. Finish. Soc. Japan 41(8) (1990)<br />

844.<br />

[276] H. Schuemichen, In: Proceedings of the<br />

Interfinish’84 (Tel Aviv, Israel, 1984) p. 411.<br />

[277] P.E. Augusston, I. Olefjord and Y. Olefjord //<br />

Werkstoffe und Korrosion 34(11) (1983)<br />

563.<br />

[278] S. Schimada and S. Maeda // Trans. ISIJ<br />

15 (1975) 95.<br />

[279] W.R. Cavanagh and J.H. Ruble, Materials<br />

Report (Penton Publ. Co., Ohio, 1966).<br />

[280] N. F. Murphy and M.A. Streicher, In: Proceedings<br />

of the 35th Annual Convention<br />

(American Electroplaters Society, 1948)<br />

p.281.<br />

[281] T. Kanamaru, T. Kawakami, S. Tanaka and<br />

K. Arai // J. Iron Steel Inst. Japan 77(7)<br />

(1991) 1050.<br />

[282] G.P. Voorkis, In: ASTM Technical Publications,<br />

ASTM STP 962 (American Society<br />

For Testing and Materials, Philadelphia,<br />

1988) p.318.<br />

[283] T.S.N. Sankara Narayanan // Prod.<br />

Finish.(London) 48(10) (1995) 19.<br />

[284] T.A. El-Mallah and H.M Abbas // Met.<br />

Finish. 85(4) (1987) 45.<br />

[285] G. Gorecki // Met. Finish. 86(12) (1988) 15.<br />

[286] T.S.N. Sankara Narayanan // Prod.<br />

Finish.(London) 47(7) (1994) 14.<br />

[287] G. Gorecki // Corros. 48(7) (1992) 613.<br />

[288] A.V. Martushene and V. Yu.A. Burrocas //<br />

R. Zh.Korr. i Zashch. Ot Korr. (1984)<br />

3K365; Met. Finish. Abs. 27(1) (1985) 17.<br />

[289] E.A. Hill, U.S. Patent 3,874,951 (1975).<br />

[290] G. Gorecki // Met. Finish. 88(5) (1990) 37.<br />

[291] G.N. Bhar, N.C. Debnath and S. Roy // Surf.<br />

Coat. Technol. 35(1/2) (1988) 171.<br />

[292] N.A. Perekhrest, K.N. Pimenova, V.D.<br />

Litovchenko and L.I. Biryuk // Zashchita<br />

Metallov 28(1) (1992) 134.<br />

[293] N. Sato // J. Met. Finish. Soc. Japan 38<br />

(1987) 571.<br />

[294] G. Clegg // Prod. Finish. (London) 41(4)<br />

(1988) 11.<br />

[295] H. Gehmecker // Metalloberflache 44(10)<br />

(1990) 485.<br />

[296] A. Barandi // Korrozios Figyelo 31(6) (1991)<br />

154.<br />

[297] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Surf. Coat. Intl. 75(12)<br />

(1992) 483.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

[298] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Trans. Inst. Met. Finish.<br />

71(1) (1993) 37.<br />

[299] T.S.N. Sankara Narayanan // Met. Finish.<br />

94(3) (1996) 47.<br />

[300] T.S.N. Sankara Narayanan and<br />

K. Ravichandran // European Coat. J. 12<br />

(2001) 39.<br />

[301] T.S.N. Sankara Narayanan, Role of surfactants<br />

in <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong>, In:<br />

Surfactants in Polymers, Coatings, Inks<br />

and Adhesives, ed. <strong>by</strong> D. Karsa (Blackwell<br />

Publishers, Oxford, 2003), Chapter 10,<br />

p.227.<br />

[302] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Met. Finish. 93(1) (1995)<br />

30.<br />

[303] G. Bikulcius, V. Burokas, A. Martusiene<br />

and E. Matulionis // Surf. Coat. Technol.<br />

172 (2003) 139.<br />

[304] D.R. Gabe, K.A. Akanni and C.P.S. Johal,<br />

In: Proceedings of the Interfinish’84, (Tel<br />

Aviv, Israel; 1984), p. 474.<br />

[305] M.O.W. Richardson and M.R. Kalantary //<br />

Trans. Inst. Met. Finish. 65(4) (1987) 132.<br />

[306] J. K. Yang, J. G. Kim and J. S. Chun // Thin<br />

Solid Films 101(3) (1983) 1993.<br />

[307] T.S.N. Sankara Narayanan // Prod. Finish.<br />

(London) 48(9) (1995) 21.<br />

[308] P.E. Augustsson, I. Olefjord and Y. Olefjord<br />

// Werkstoffe und Korrosion 34 (1983) 563.<br />

[309] A. Stoch and J. Stoch // J. Solid State<br />

Ionics 34(1-2) (1989) 17.<br />

[310] W.J. van Ooij and T.H. Visser //<br />

Spectrochim. Acta (B) 39B (1984) 1541.<br />

[311] N. Sato and T. Minami // J. Met. Finish.<br />

Soc. Japan 38(4) (1987) 149.<br />

[312] J.E. de Vries, T.L. Riley, J.W. Holubka and<br />

R.A. Dickie // Surf. Interface Anal. 7(3)<br />

(1985) 111.<br />

[313] R.A. Choudhery and C.J. Vance, In: Advances<br />

in Corrosion Protection <strong>by</strong> Organic<br />

Coatings, ed. <strong>by</strong> D. Scantelbury and<br />

M. Kendig (The Electrochemical Society,<br />

NJ, 1989) p. 64.<br />

[314] J. De Laet, J. Vanhellemont, H. Terryn and<br />

J. Vereecken // Thin Solid Films 233 (1993)<br />

58.<br />

[315] T. Minami and N. Sato // J. Surf. Sci. Soc.<br />

Japan 84(4) (1988) 459.<br />

[316] N. Sato // J. Met. Finishing. Soc. Japan 38<br />

(1987) 30.<br />

175<br />

[317] N. Sato and T. Minami // J. Surf. Sci. Soc.<br />

Japan 9 (1988) 459.<br />

[318] N. Sato, K. Watanabe and T.Minami //<br />

J. Mater. Sci. 26 (1991) 865.<br />

[319] A.J. Sommer and H. Leidheiser, Jr., In:<br />

Proceedings of the 19 th Annual Conference<br />

of The Microbeam Analysis Society (Microbeam<br />

Analysis Society. San Francisco, CA,<br />

1984) p. 111.<br />

[320] N. Sato, K. Watanabe and T. Minami //<br />

J. Mater. Sci. 26 (1991) 1383.<br />

[321] K. Nomura and Y. Ujihira // J. Mater. Sci. 17<br />

(1982) 3437.<br />

[322] P.E. Tegehall and N.G. Vannerberg //<br />

Corros. Sci. 32(5/6) (1991) 635.<br />

[323] G. Rudolph and H. Hansen // Trans. Inst.<br />

Met. Finish. 50(2) (1972) 33.<br />

[324] W.J. van Ooij, T.H. Viseer and M.E.F.<br />

Biemond // Surf. Interface Anal. 9 (1984)<br />

187.<br />

[325] X. Sun, D. Susac, R. Li, K.C. Wong,<br />

T. Foster and K.A.R. Mitchell // Surf. Coat.<br />

Technol. 155 (2002) 46.<br />

[326] N. Sato // J. Surf. Finish. Soc. Japan 40<br />

(1989) 933.<br />

[327] M. Handke, A. Stoch, V. Lorenzelli and P.L.<br />

Bonova // J. Mater. Sci. 16 (1981) 307.<br />

[328] M. Wolpers and J. Angeli // Appl. Surf. Sci.<br />

170 (2001) 281.<br />

[329] D. He, F. Chen, A. Zhou, L. Nie and S. Yao<br />

// Thin Solid Films 382 (2001) 263.<br />

[330] D. He, A. Zhou, Y. Liu, L. Nie and S. Yao //<br />

Surf. Coat. Technol. 126 (2000) 225.<br />

[331] S. Maeda, M. Yamamato and K. Suzuki //<br />

Trans. Iron Steel Inst. Japan 24(9) (1984) B-<br />

301.<br />

[332] M. Suzuki, R. Kojima, K.I. Suzuki,<br />

K. Nishizaka and T. Ohtsububo // Trans Iron<br />

Steel Inst. Japan 25(9) (1985) B-220.<br />

[333] B. Maurickx, J.M. Anne and J. Foct // Surf.<br />

Interface Anal. 9(1-6) (1986) 386.<br />

[334] T.S.N. Sankara Narayanan // Prod. Finish.<br />

(London) 47(6) (1994) 17.<br />

[335] J. Flis, J. Mankowski, T. Zakroczymski and<br />

T. Bell // Corros. Sci. 43 (2001) 1711.<br />

[336] H. G. Mosle and B. Wellenkotter //<br />

Metalloberflache 34 (1980) 424.<br />

[337] H. G. Mosle and B. Wellenkotter //<br />

Metalloberflache 37 (1983) 94.<br />

[338] J. Rooum and R.D. Rawlings // J. Oil Colour<br />

Chem. Asso.(JOOCA) 71(6) (1988) 140.<br />

[339] J. Rooum and R.D. Rawlings // J. Mater.<br />

Sci. 17(6) (1982) 1745.


176 T.S.N. Sankara Narayanan<br />

[340] T.S.N. Sankara Narayanan // Pigment Resin<br />

Technol. 24(4) (1995) 12.<br />

[341] O. Pawlig, V. Schellenschlager, H.D. Lutz<br />

and R. Trettin // Spectrochim. Acta A 57<br />

(2001) 581.<br />

[342] R.J. Hill and J.B. Jones // Am. Mineral. 61<br />

(1976) 987.<br />

[343] R.J. Hill // Am. Mineral. 62 (1977) 812.<br />

[344] L. Kwiatkowski and A. Kozlowski, In:<br />

Testing of Metallic and Inorganic Coatings,<br />

ed. <strong>by</strong> W.B. Harding and G.A. Di Bari<br />

(ASTM STP 947, American Society for<br />

Testing and Materials, Philadelphia, 1987)<br />

p. 272.<br />

[345] R.St. J.Preston, R.H. Settle and J.B.L.<br />

Worthington // J. Iron Steel Inst. 170(1)<br />

(1952) 19.<br />

[346] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Met. Finish. 91(8) (1993)<br />

43.<br />

[347] L. Kwiatkowsi, A. Sadkowski and<br />

A. Kozlowski, In: Proceedings of the 11 th<br />

International Congress on Metallic Corrosion<br />

(Italy, Vol. 2, 1990) p.395.<br />

[348] ASTM D 3359, American Society for Testing<br />

and Materials, Philadelphia, 1987.<br />

[349] ASTM B 117, American Society for Testing<br />

and Materials, Philadelphia, 1985.<br />

[350] IS-3618, Indian Standards Institution, New<br />

Delhi, India, 1966.<br />

[351] R.L. Chance and W.D. France, Jr. // Corros.<br />

25(8) (1969) 329.<br />

[352] C.P. Vijayan, D. Noel and J.J. Hechler, In:<br />

Polymeric Materials for Corrosion Control,<br />

ed. <strong>by</strong> Ray A. Dickie and F.Louis Floyd<br />

(ACS Symposium series 322, 1986) p. 58.<br />

[353] C. Marrikannu, P. Subramanian,<br />

B. Sathianandham and K. Balakrishnan //<br />

Bull. Electrochem. 5(8) (1989) 578.<br />

[354] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Met. Finish. 90(10) (1992)<br />

15.<br />

[355] K. Kiss and M. Coll-Palagos // Corros. 43(1)<br />

(1987) 8.<br />

[356] K.H. Ruiz and D.D. Davidson, SAE Paper<br />

No. 912299 (1991).<br />

[357] D. Wang, P. Jokiel, A. Uebleis and<br />

H. Boehni // Surf. Coat. Technol. 88 (1996)<br />

147.<br />

[358] A. Losch and J.W. Schultze //<br />

J. Electroanal. Chem. 359 (1993) 39.<br />

[359] J. Flis, Y. Tobiyama, C. Shiga and<br />

K. Mochizuki // J. Appl. Electrochem. 32<br />

(2002) 401.<br />

[360] J. Flis, Y. Tobiyama, K. Mochizuki and<br />

C. Shiga // Corros. Sci. 39(10-11) (1997)<br />

1757.<br />

[361] G. Lendvay-Gyorik, G. Meszaros and<br />

B. Lengyel // J. Appl. Electrochem. 32<br />

(2002) 891.<br />

[362] T.S.N. Sankara Narayanan and<br />

M. Subbaiyan // Met. Finish. 92(9) (1994)<br />

33.<br />

[363] T.S.N. Sankara Narayanan // European<br />

Coat. J. 4 (1999) 138.<br />

[364] Electrochemical Corrosion Testing, ed. <strong>by</strong><br />

F. Mansfeld and V.Bertocci (ASTM STP<br />

727, American Soceity for Testing and<br />

Materials, Philadelphia, 1981).<br />

[365] F. Mansfeld and M. Kendig, In: Proceedings<br />

of the 9th International Congress on Metallic<br />

Corrosion (Toronto, Vol. 3, 1984) p.74.<br />

[366] W. Wiederholt // Electroplat. Met. Finish.<br />

18(10) (1965) 334.<br />

[367] J.A. Scott // Trans. Inst. Met. Finish. 46<br />

(1968) 32.<br />

[368] L. Fedrizzi, F. Deflorian, S. Rossi, L. Fambri<br />

and P.L. Bonora // Prog. Org. Coat. 42<br />

(2001) 65.<br />

[369] R.A. Ashdown // Prod. Finish.(London)<br />

27(11) (1974) 33.<br />

[370] P. Morris // Ind. Finish.(London) 28(339)<br />

(1976) 4.<br />

[371] Y. Barry // World Surface Coatings Abstracts<br />

43 (1970) 332,199.<br />

[372] G.W. Critchlow, P.W. Webb, C.J. Tremlett<br />

and K. Brown // Intl. J. Adh. Adhes. 20<br />

(2000) 113.<br />

[373] K. Nyavor and N.O. Egiebor // The Science<br />

of the Total Environment 162 (1995) 225.<br />

[374] S. Rebeyrat, J.L. Grosseau-Poussard, J.F.<br />

Dinhut, P.O. Renault // Thin Solid Films 379<br />

(2000) 139.<br />

[375] B. Zhao, Z. Liu, Z. Zhang and L. Hu //<br />

J. Solid State Chem. 130(1) (1997) 157.<br />

[376] S. Rebeyrat, J.L. Grosseau-Poussard, J.F.<br />

Silvain, B. Panicaud and J.F. Dinhut // Appl.<br />

Surf. Sci. 199 (2002) 11.<br />

[377] H. Bala, N.M. Trepak, S. Szymura, A. A.<br />

Lukin, V.A. Gaudyn, L.A. Isaicheva,<br />

G. Pawlowska and L.A. Ilina // Intermetallics<br />

9 (2001) 515.<br />

[378] R. Ribitch, U.S. Patent 5,067,990 (1991).<br />

[379] H. Transcher // Draht 11(8) (1960) 442.


Surface <strong>pretreatment</strong> <strong>by</strong> <strong>phosphate</strong> <strong>conversion</strong> <strong>coatings</strong> <strong>–</strong> a <strong>review</strong><br />

[380] D.G. Placek and S.G. Shankwalkar // Wear<br />

173(1-2) (1994) 207.<br />

[381] A. Kozlowski and W.Czechowski //<br />

Electrodep. Surf. Treat. 3 (1975) 55.<br />

[382] P. Hivart, B. Hauw. J. Crampon and J.P.<br />

Bricout // Wear 219 (1998) 195.<br />

[383] M. Khaleghi, D. R. Gabe and M. O. W.<br />

Richardson // Wear 55(2) (1979) 277.<br />

[384] P. Hivart, B. Hauw, J.P. Bricout and J. Oudin<br />

// Tribol. Intl. 30(8) (1997) 561.<br />

[385] J. Perry and T. S. Eyre // Wear 43(2) (1977)<br />

185.<br />

[386] D. James // Wire and Wire Products 45(7)<br />

(1970) 37.<br />

[387] M. Bull // Metallurgia 51(9) (1984) 393.<br />

[388] Hanns Ketterl, German Patent 1,170,750<br />

(1964).<br />

[389] L. Bonello and M.A.H. Howes, Heat Treatment’<br />

79 (The Metals Society, London,<br />

1979).<br />

[390] L. Lazzarotto, C. Marechal, L. Dubar,<br />

A. Dubois and J. Oudin // Surf. Coat.<br />

Technol. 122 (1999) 94.<br />

[391] G. Bustamante, F.J. Fabri-Miranda, I.C.P.<br />

Margarit and O.R. Mattos // Prog. Org.<br />

Coat. 46 (2003) 84.<br />

177<br />

[392] F. Dausinger, W. Muller and P. Arnold, U.S.<br />

Patent 4,414,038 (1983).<br />

[393] P. Gay, In: Laser Surface Treatment of<br />

Metals, ed. <strong>by</strong> C.W. Draper and P. Mazzoldi<br />

(Martinus Nijhoff Publishers, Dordrecht,<br />

1986) p. 201.<br />

[394] T.S.N. Sankara Narayanan // Met. Finish.<br />

94(11) (1996) 38.<br />

[395] D.L. Clemmons and J.D. Camp //<br />

Electrochem. Technol. 2(7/8) (1964) 221.<br />

[396] D. Wenf, R. Wang and G. Zhang // Met.<br />

Finish. 96(9) (1998) 54.<br />

[397] W. Buchmeier and W.A. Roland, U.S.<br />

Patent 5,350,517 (1994).<br />

[398] D. Pearson and J.C. Wilson, British Patent<br />

1,545,515 (1979).<br />

[399] M.D. Waite, U.S. Patent 5,376,342 (1994).<br />

[400] R.F. Waters, H.E. Powell and L.N. Bollard,<br />

U.S. Patent 3,653,875 (1972).<br />

[401] C.A. Gill and C.M. Berbiglia, U.S. Patent<br />

5,273,667 (1993).<br />

[402] D.S. Peters, U.S. Patent 4,986,977 (1991).<br />

[403] C.J. Baldy // Met. Finish. 94(11) (1996) 23.<br />

[404] J. Caponero and J.A.S. Tenorio // Resources,<br />

Conservation and Recycling 29<br />

(2000) 169.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!