Sie sind auf Seite 1von 5

FEATURE | Material

Characterization & Testing

Microstructure
of Nitrided Steels
George F. Vander Voort Struers Inc.; Wadsworth, Ill.
Nitriding is one of the most interesting and useful surface-hardening techniques.
It is unique in that during the nitriding process, the specimen is not heated into
the austenite phase, and it does not rely upon the formation of martensite to
achieve high hardness and useful properties. It is heat treated prior to nitriding,
forming tempered martensite to obtain the desired core properties unlike all other
surface heat-treatment processes.

h processing associated with


he
nitriding
does have some adn
vantages
in
avoiding problems
v
such
as
quench
cracking and
s
distortion. It also has some side benets
in improved corrosion resistance and
generation of benecial residual compressive stresses, which improves fatigue resistance. Nitrided surfaces do exhibit high
surface hardness, leading to improved
wear resistance.
Of course, like any process, there
are disadvantages. One that has always
plagued nitriding is the very long cycle
time to achieve a decent case depth, about
20 times longer than for carburizing considering equal case depths. The other
problem has been the generation of the
compound layer, often erroneously called
the white-etching layer, which is brittle
and is generally deleterious if present.
Etching does not color this layer white; it
is white, just as it was as-polished, since

the etchant had no effect on the layer.


Literature Review
The literature on the microstructure of nitrided steels does, unfortunately, contain
numerous errors and could be improved.
Examples of the microstructures of nitrided steels in the literature are often subpar
and even poor or false. Phase identication of the compound layer and the underlying diffusion zone need work.
In the past, bulk X-ray diffraction on
the OD surface has been the chief tool for
phase identication. While a useful tool,
it needs augmentation by microanalytical
methods with very small spatial resolution
for phase identication. Limited work has
been done using well-established tools such
as the transmission electron microscope.
This is partly due to the difculty in preparing thin foils from relatively small regions at the surface compared to the ease in
preparing foils from bulk specimens. Also,

it is somewhat a reection of the types of


research studies being funded at universities in the U.S. The writer is currently exploring the use of electron backscattered
diffraction (EBSD) with the scanning
electron microscope (SEM) to see if this
technique can provide an easier approach
for analysis of the compound layer.
In the literature, one can see nitrided
examples of low-carbon steels; low-carbon, low-alloy steels; and HSLA (highstrength, low-alloy) steels (none of these
contain signicant amounts of alloying elements that form hard nitrides, such as Al,
Cr, V or Mo). They exhibit an outer layer,
called the compound layer, which is reported to be composed of iron nitrides: av
(Fe4N) and (Fe2N1-x) phases.
Little, if any, solid-solution strengthening occurs from the diffused nitrogen into
the steel. Some needle-like intragranular
iron nitrides, probably av, may be seen in
ferrite grains below the compound layer,

400

Hardness, Knoop

350
300
250
200
150

Fig. 1. Microstructure of salt-bath nitrided, resulfurized 1215 carbon steel with a compound
zone (black arrow), no diffusion zone and some
nitride needles (white arrows) in the ferrite
grains. Etched with a 10:1 mixture of 4% picral
and 2% nital (1250X, oil immersion).

100
0

0.2

0.4

0.6

0.8

Depth, mm

Fig. 2. Knoop hardness profile (100 gf load) starting as close as possible to the compound layer to a depth of 1 mm.
IndustrialHeating.com - November 2011 51

FEATURE | Material
Characterization & Testing

0.002"

Fig. 4. Nitralloy 135 with a nitrided surface


case free of any harmful compound layer.
Etched with 2% nital followed by 10%
sodium metabisulfite. Note the white
grain-boundary films (white arrows) in the
500X micrograph.

0.002"

Fig. 5. The white grain-boundary films


that are relatively parallel to the surface
were darkened (white arrows) by etching
with alkaline sodium picrate at ~90C for
60 seconds. This etch colors cementite,
Fe3C, when used in this manner (original
at 500X).

Hardness, Knoop

1200
1000
800
600
400
200
0
Depth, mm

Fig. 6. Knoop hardness profile plot (100 gf load) for Nitralloy 135, free of any compound
zone, showing a very high hardness case.
0.010"

Fig. 3. Composite image made from several


contiguous fields to show the case/core microstructure of the Nitralloy 135 specimen,
etched with a 10:1 mix of 4% picral and 2%
nital (originals at 200X).

but they have little inuence on the case


hardness. Consequently, nitriding has
been mainly centered upon steels that
contain elements that form very ne, very
hard nitrides: Al, Cr, V and Mo. Elements
such as Ti and Zr do form very hard nitrides in steels, but they are comparatively
quite large (in the m range) and do not
create a case-hardness prole.
The original experiments [1] on nitriding performed by Adolph Machlet at the
American Gas Company in Elizabeth,
N.J., (see U.S. Patent 1,092,925, dated 24
June 1913) focused on nitriding carbon
52 November 2011 - IndustrialHeating.com

steels. In 1906, Adolph Fry of the Krupp


Steel Works in Essen, Germany, began a
similar study of nitriding.[1] However, Fry
realized early in his work that alloying
elements were necessary to develop commercially useful nitrided steels. Frys U.S.
patent (1,487,554) was granted on March
18, 1924. Fry learned that only steels containing additions of Cr, Mo, Al, V or W
could achieve a high surface hardness
when nitrided. His work at Krupp led to
the development of the Nitralloy grades.
Etchants

A perusal of publications regarding nitriding reveals a range of etchants that


have been used. Nital, by far the most
widely used etchant for steels, has been
commonly used for nitrided steels. McQuaid and Ketcham used 4% nital to
etch nitrided Cr-Al and Mo-Al steels and
AISI/SAE 4615 in their study published

in 1928.[1] Robert Sergeson, a research


metallurgist at the Central Alloy Steel
Corporation in Canton, Ohio, and an
early researcher of the nitriding process,
introduced the use of a 10-to-1 solution of
4% picral plus 4% nital in 1929 (although
many people have used 2% nital instead
of 4%).[2] This is an excellent etchant.
Lightfoot and Jack[3] studied nitriding with and without formation of the
compound layer. They noted that during
nitriding a carbon-rich layer is created
ahead of the nitrided case. This carbon
precipitates as cementite in grain boundaries that are roughly parallel to the surface. This carbon accumulation caused by
the inward diffusion of nitrogen has been
veried by others.[4-6] Jegou et al.[6] used
the electron microprobe to measure the
nitrogen and carbon proles through the
nitrided case. These authors showed the
grain-boundary lms darken in specimens

nitrided for 10 and 100 hours when they


etched with boiling picral. They probably used boiling alkaline sodium picrate.
It is well known that alkaline sodium
picrate used at 80-100C will darken cementite (Fe3C) carbide. Albert Sauveur,
dean of American metallurgists, attributed [7] the boiling alkaline sodium-picrate etch for identication of cementite
to Kourbatoff in 1906. Despite studies
proving that the white grain-boundary
lms in the diffusion zone are cementite,
numerous authors have stated that they
are nitride (Reference 8, for example).
This seems to be a common natural
error, assuming that the white grainboundary lm is a nitride just like the
compound layer.
The only systematic study on the use
of etchants to identify the phases in the
compound zone and the white grainboundary lms is by Mridha and Jack.[9]
They evaluated 10 different etchants and
showed that for nitrided pure iron, nital
does not distinguish the phases in the
compound zone. They concluded that
the best reagents to distinguish av from
are picral (etches boundaries in the nitride phases), Vilellas reagent (attacks
boundaries in and stains av), a sulfatechloride solution (stains only ) and
Oberhoffers reagent (a short 2-5 second
etch dissolves ).
For alloy steels (chiey 3% Cr steels),
the boundary of the nitrided zone was
best revealed using picral, Marbles and
Oberhoffers reagents. The latter two
etchants best revealed the extent of the
carbide-enriched region beneath the
nitrided region. The sulfate-chloride reagent was sensitive to all constituents.
The best etchant for revealing the white
grain-boundary lms of cementite was alkaline sodium picrate used at 85C for 2
minutes. This etch also revealed the presence of cementite in the compound zone.
Cementite was conrmed using X-ray
diffraction. They recommended etching
nitrided steels rst with alkaline sodium
picrate and then with Oberhoffers reagent for 3 seconds. Reference 10 by the
same authors covers the characterization

of nitrided 3% Cr steels using etchants


selected based upon this study.[9]
Microstructures
An example of a low-carbon, resulfurized
steel AISI/SAE 1215 that was saltbath nitrided is shown in Figure 1. This
specimen was etched with a 10-to-1 mix of
4% picral to 2% nital, an etchant that has
often been used to reveal the structure of
nitrided steels. Note that we see a welldeveloped compound layer. Because 1215
has no appreciable content of alloying elements that would form alloy nitrides (Al,
Cr, V, Mo or W), there is no diffusion
zone (as shown later for Nitralloy 135 and
41B50). Note the iron-nitride needles
(arrows) intragranular within the ferrite
grains beneath the compound layer.
Figure 2 shows a plot of Knoop hardness (100 gf load) versus depth curve
starting as close as possible to the compound layer. This layer is too thin to actually test, so the rst indent is not able to
evaluate the actual hardness of the compound layer. But one can see from the rest
of the data that nitriding has had only a
minor inuence on the case hardness.
Figure 3 shows the case/core microstructure of gas-nitrided Nitralloy 135,
an alloy developed by Frys work reported
above, which contains ~1.1% Al, ~1.6%
Cr and ~0.2% Mo good nitride formers.
Note that this specimen does not exhibit
a compound layer. With todays digital
technology, we can create a mosaic image from a series of images of contiguous
elds, where we control the lighting and
then weld the digital images together
with appropriate software. This is marvelous technology compared to the older,
very painful practice of trying to glue
prints taken in a similar alignment, which
can no longer be done.
Figure 4 shows the white grain-boundary lms in the diffusion zone that have often been erroneously identied as nitrides
(in studies without any analytical work)
because they are white. Figure 5 shows
that these grain boundaries are darkened
when etched with hot alkaline sodium
picrate, proving that they are cement-

Electroless Ni plating
1
2
3

Fig. 7. Microstructure (as-polished) of a


failed nitrided 41B50, lightly resulfurized
(note MnS stringers white arrows) chuck
jaw for a lathe. The surface was plated with
electroless Ni for edge support. The black
arrows point to the complex compound
layer (original at 500X).

Electroless Ni plating
1
2
3

Fig. 8. Failed 41B50 chuck jaw with a nitrided surface exhibiting a massive compound
layer etched with 10:1 mix of 4% picral and
2% nital (original at 500X).
1 2

20 m

Fig. 9. Failed 41B50 chuck jaw etched with


10% Na2S2O5. The black arrow points to
the Ni plating . The green, red and blue
arrows point to the complex compound
layer, while the white arrows point to the
white grain-boundary films approximately
parallel to the surface. Note the cracks in
the complex brittle compound layer that
failed in service (original at 500X).
IndustrialHeating.com - November 2011 53

FEATURE | Material
Characterization & Testing

Mns
Electroless nickel

Mns
0.002"

0.002"

Fig. 10. Views of the surface (left) and diffusion zone (right) of the nitrided 41B50 specimen
after etching with alkaline sodium picrate (90C for 90 seconds) that colors cementite (red
arrows originals at 500X).
900
Hardness, Knoop

800
700
600
500
400
300
200
100
0

0.1

0.2

0.3

0.4
Depth, mm

0.5

0.6

0.7

0.8

Fig. 11. Knoop hardness (100 gf load) profile of the failed nitrided 41B50 chuck-jaw specimen revealing a low hardness in the complex compound layer.

ite. Figure 6 shows the Knoop hardness


(100 gf load) prole for this specimen
markedly better than for the 1215 carbon
steel shown in Figure 2.
A more complex example of a nitrided
alloy steel, resulfurized 41B50, is shown in
Figure 7. This was a chuck jaw made for a
lathe that broke as soon as it was put into
service due to the brittle nature of the surface layer. The specimen was electroless
nickel-plated to enhance edge retention.
Figure 7 shows the surface in the aspolished condition. Note the MnS stringers in this lightly resulfurized alloy steel.
Three zones can be seen in the complex
compound layer. This type of complex
compound zone has been reported in the
literature, but it is not common. Zone 1
is believed to be epsilon phase formed by
outward diffusion of Fe along pore channels and reaction with N at the surface.
Zone 2 is believed to be porosity in the epsilon phase (according to the literature),
which may be lled with oxide.
Examination of this zone with dark54 November 2011 - IndustrialHeating.com

eld illumination did not conrm that


the black spots are holes, however, nor
did the inital SEM examination at high
magnication. So, more work is needed to
positively identify this dark portion of the
compound zone.
Zone 3 is the classic mixture of epsilon
and gamma-prime phase. In the microscope there is a slight dark/light contrast
difference between the two phases, which
can be faintly seen in the micrograph. Figure 8 shows the complex compound zone
after etching with the 10:1 mixture of 4%
picral to 2% nital. Note that detail is revealed in the compound zone (Zone 3).
Figure 9 shows the compound zone and
diffusion zone after etching with 10% sodium metabisulte. There is what appears
to be oxidation between the electroless
nickel plating and zone 1, as also shown
in Figures 7 and 8. Figure 10 shows two
views of the complex compound layer and
the diffusion zone of the failed nitrided
part after etching with alkaline sodium
picrate at 90C for 90 seconds to color the

cementite. Note that the lower edge of the


compound zone (area 3) contains cementite, as reported by previous researchers.
The white grain-boundary lms are darkened by the alkaline sodium-picrate etch
while the coarser cementite in zone 3 of
the compound layer is a mix of blue and
blackish particles.
Figure 11 shows the Knoop hardness
(100 gf load) prole for the failed nitrided
41B50 chuck-jaw specimen. Note that the
hardness in the outer dark surface (zone 2)
of the compound zone is lower than the inner (zone 3) layer. Zone 1 is far too thin to
determine its hardness accurately. 41B50
contains ~0.95% Cr and ~0.20% Mo, but
neglible Al (a small amount may be present for grain renement). Note that the
maximum case hardness obtained in the
nitrided 41B50 specimen is much lower
(~620-690 HK) than obtained for the nitrided Nitralloy 135 specimen (~1000-1040
HK) with ~1.1% Al, 1.6% Cr and ~0.2%
Mo (Fig. 11 compared to Fig. 6). Both cases, however, are markedly harder than that
of the nitrided 1215 carbon steel (~300-340
HK) in Fig. 2. This shows just how critical
it is to develop an alloy composition that
will form very hard, very small nitrides.
Conclusions
Metallography, when properly performed,
is an exceptionally important tool for
studying the microstructure of nitrided
steels, as well as other heat-treated metals and alloys. All etchants are not equal,
and nital is not always the best etch for all
steels, despite its wide usage. Different steel
compositions do respond quite differently
to nitriding, as illustrated by the comparison of a nitrided carbon steel with two alloy
steels one with a greater concentration of
alloying elements that will form very ne,
hard nitrides compared to the leaner alloy. The Knoop hardness proles for these
three steels were markedly different.
Nitriding processes must be controlled
to eliminate the brittle compound layer,
which has been known to cause failures
when present. The work shows that the
compound layer can be rather variable in
appearance. It is also common to see white

grain-boundary lms only in the boundaries that are parallel or nearly parallel to
the specimen surface. These lms have
frequently been claimed to be nitrides, but
numerous studies have proven that they
are cementite.
The exact mechanism for the formation
of these lms has not been fully dened,
although there are a few good preliminary
studies. It appears that as nitrogen is diffused into the steel, carbon is pushed from
the surface inward. Only limited electron
microprobe (EMPA) work has been done
to study the C and N case proles, but
these show that the C is depleted at the
surface and pushed inward while the N
content is highest at the surface and drops
as the case hardness decreases. Application of good analytical techniques, such
as EBSD and the EMPA, in future studies
should enhance our understanding of the
nitriding process. IH

References
1. H.W. McQuaid and W.J. Ketcham, Some
Practical Aspects of the Nitriding Process,
Trans. of ASST, Vol. 14, 1928 (Republished
in the Source Book on Nitriding, American
Society for Metals, Metals Park, Ohio, 1977,
pp. 1-25).
2. R. Sergeson, Investigations in Nitriding,
ASST Nitriding Symposium, 1929 (republished in the Source Book on Nitriding,
American Society for Metals, Metals Park,
Ohio, 1977, pp. 26-55).
3. B.J. Lightfoot and D.H. Jack, Kinetics of
Nitriding With and Without White-Layer
Formation, Heat Treatment 73, The Metals Society, December 1973 (republished
in the Source Book on Nitriding, American
Society for Metals, Metals Park, Ohio, 1977,
pp.248-254).
4. L. Barrallier et al., Morphology of Intergranular Cementite Arrays in Nitrided Chromium-Alloyed Steels, Materials Science and
Engineering, Vol. A393, 2005, pp. 247-253.
5. V. Yu. Traskine et al., Physicochemical Mechanics of Structural Transformations in
Nitrided Steel, Colloid Journal, Vol. 67, No.
1, 2005, pp. 97-102.

6. S. Jegou, L. Barrallier, R. Kubler and M.A.J.


Somers, Evolution of Residual Stress in
the Diffusion Zone of a Model Fe-Cr-C Alloy During Nitriding, HTM J. Heat Treatment Mat., Vol. 66, No. 3, 2011, pp. 1-8.
7. A. Sauveur, The Metallography and Heat
Treatment of Iron and Steel, 4th ed., McGraw-Hill Book Co., NY, 1935, pgs. 482, 502
and 504.
8. R. Agnelli et al., Failure Analysis in Tool
Steels, Failure Analysis of Heat Treated Steel
Components, L.C.F Canale, R.A. Mesquita
and G.E. Totten editors, ASM International,
Materials Park, Ohio, 2008, pp. 311-350.
9. S. Mridha and D.H. Jack, Etching Techniques
for Nitrided Irons and Steels, Metallography,
Vol. 15, No. 2, May 1982, pp. 163-175.
10. S. Mridha and D.H. Jack, Characterization
of Nitrided 3% Chromium Steel, Metal Science, Vol. 16, August 1982, pp. 398-404.
For more information: Contact George F.
Vander Voort, Consultant Struers Inc., 24766
Detroit Rd., Westlake, OH 44145; tel: 847-6237648; e-mail: georgevandervoort@yahoo.
com; web: www.struers.com and www.
georgevandervoort.com.

Your Thermal Process


Deserves Special Treatment.
If you cant afford to take a
chance with a high value
load, you cant get a more
reliable furnace, oven or
kiln than an L & L Special
at any price! Every one
of our furnaces is
special!

GRAPHALLOY
BEARINGS CAN
TAKE THE HEAT.

L&L can meet the strictest provisions of


AMS2750D for aerospace applications.

F U R N A C E

C O , I N C

20 Kent Road, Aston, PA 19014


P: 877.836.8045 F: 610.459.3689
E: sales@hotfurnace.com W: hotfurnace.com

HANDLE HIGH TEMPERATURE AND HARSH


OPERATING CONDITIONS WITH EASE
GRAPHALLOY bushings,
bearings and components:
Survive when others fail
Run hot, cold, wet or dry
Excel at -450F to 1000F
Corrosion resistant
Self-lubricating
Non-galling
Low maintenance
Ovens, dryers, pumps, valves,
turbines, mixers, conveyors

GRAPHITE METALLIZING
CORPORATION
Yonkers, NY 10703 U.S.A.
ISO 9001:2008
H06a

TEL. 914.968.8400 WWW.GRAPHALLOY.COM/IH

IndustrialHeating.com - November 2011 55

Das könnte Ihnen auch gefallen