Sunteți pe pagina 1din 16

Heat-treatment Processing of Austenitic Manganese Steels

Seluk Kuyucak, Renata Zavadil, Val Gertsman


CANMET Materials Technology Laboratory, Ottawa, Ontario, Canada
ABSTRACT
Hadfields austenitic manganese steels are best described as retained austenites.
Steels heat-sensitized at 400 800C range, such as slowly cooled castings, develop intergranular embrittlement caused by hypereutectoid carbide precipitation.
A solution annealing and water quenching heat treatment restores the mechanical
properties. The present paper is an overview of research results describing the
quantitative effect of grain boundary constituents on impact toughness and effect
of section size on heat treatment, and a method of measuring of quench time using
quench water temperatures.
Keywords: manganese steel castings, heat-treatment, embrittlement
INTRODUCTION
Hadfields steels with a basic composition of 1.2% C, 13% Mn are metastable
austenites. This basic composition has remained the same since its discovery by
Robert Hadfield in 1882. Both carbon and manganese help stabilize austenite
against martensite formation. The martensite start temperature Ms is typically
below the temperature of liquid nitrogen (196C). The embrittling carbides
present in the as-cast structure are removed by solution annealing above 1000C
and quenching. The resulting retained austenite structure with a large amount
of carbon in solution gives these steels their high impact toughness and high
strain-hardening rate and capacity, which give them their excellent impact-abrasion resistance.
This overview of our recent studies on manganese steels [1-10] was funded by the
American Foundry Society and the North American metal casting industry. It
encompasses the nature of embrittling phases and their quantitative relationship to
impact toughness, the behaviour of carbides and phosphide eutectic, structural
soundness and microporosity distribution, macro- and micro- segregation, and a
method of measuring quench time as a process control tool. The overview fo-

cuses on results, the specific details of procedures and materials are found in
individual research papers.
BEHAVIOUR OF CARBIDES
The embrittling intergranular carbides in manganese steels form during slow
cooling or reheating through the 400 800C range. They are removed by solution annealing above 1000C followed by rapid quenching. The kinetics of carbide formation follow the typical C-curve of an isothermal transformation diagram, with the fastest growth (carbide nose) occurring at 600 650C. Essentially,
two types of carbides form: the thin carbides form very rapidly and appear as
grain boundary delineations in the etched structure at 200-1000 (Fig. 1a-d).
Later, thick carbides nucleate on thin carbides and grow along the grain boundaries. The thin carbides do not embrittle the steel significantly, whereas the thick
carbides embrittle severely. The latter are distinguished from the thin carbides by
having a resolvable cementite interior with a clear austenite / carbide interphase
boundary on both sides of the cementite film. The thin carbides remain as delineations even at high magnifications, their cementite is not resolved at 1000.
In regular manganese steels (ASTM A 128, Grade B), the thin carbide delineations are less than 0.2 m thick, and the thick carbides have a starting thickness of
0.5 1.5 m, and appear as a step where they meet the thin carbides (arrows in
Fig. 1). In Cr-bearing manganese steels (Grade C), intermediate thickened carbides are also observed (Fig. 1e,f). These are effectively similar to thin carbides.
They appear as thick diffuse delineations and the cementite is not resolved by
light microscope.
Figures 2a-d show extracted replica images of grain boundary carbides by transmission electron microscopy (TEM). The thickened carbides in Grade C steel are
seen to have a straight interface with one austenite grain and a serrated interface
with the other. Although, their total width in the thickened areas have reached 3
4 m, their centers are 1.0 1.5 m thick when their finger-like growth is excluded.
Figures 2e-g show thin foil TEM micrographs of the three carbides. The thin carbide delineations are seen to be discreet particles. They nucleate on a grain boundary, but growth usually takes place into only one austenite grain. The thickened carbides in Grade C steel also consist of many discreet particles but with greater protrusions into austenite matrix. There are indications of stress on the lattice
particularly as the particles get larger (Fig. 2f). The thick carbide films, on the
other hand, are continuous (Fig. 2g).
Figure 3 shows energy dispersive X-ray analysis (EDX) line scans across a thin
and a thick carbide in a Grade C steel using the TEMs nanoprobe. The thin carbides Mn and Cr contents are close to those of austenite matrix, 13% and 2.5%,
respectively. The thick carbide has significantly higher concentrations of Mn and
Cr in solid solution (18% and 6%, respectively) and the diffusion of these elements into the carbide causes a depleted layer in the surrounding matrix. The lite2

rature on phase diagrams shows that in a 13% Mn, 1.2% C steel, the Mn content
of the cementite that forms at 600C varies between 17% Mn (Benz 1973, in [11])
and 32% Mn (Koch and Keller 1964, in [11]). The 18% measured composition of
the thick carbide in Fig. 3 falls within this range. This supports the hypothesis
that the thick carbides are the stable phase, and the thin carbides are metastable.
The thin carbides appear as delineations under light microscope, primarily because they are comprised of many small, discreet particles. The thickened carbides in Grade C steel are essentially the same, with greater protrusions into austenite matrix. Crystallographically, all carbides index to an M3C cementite structure but their chemical compositions are significantly different. It is proposed that
the following mechanism operates for carbide precipitation during quenching:

The thin carbides are less stable but have a good lattice match with the
austenite matrix. This gives rise to a low energy interface and a low nucleation barrier. The thick carbides, on the other hand, are the more stable
constituents, but have a higher energy interface and a higher nucleation
barrier.

Discrete, metastable thin carbides whose Mn and Cr contents are similar to


those of the matrix nucleate rapidly at the grain boundaries. They are
comprised of many discreet particles, which give rise to the observed delineations in the etched microstructure. These carbides have good cohesion
with austenite because of their small size and low energy, possibly coherent, interface. They often show preferred growth into one austenite grain.
In Grade C steel, this growth is more pronounced, resulting in the appearance of thickened carbides (Figs. 1f, 2c,d,f).

Soon after they form, the thin carbide growth slows down and stops because of the increased lattice mismatch and stress at the interface, and the
driving force being low, is not able to overcome the increased interfacial
energy. Eventually, a thick carbide with a larger driving force and an incoherent interface forms.

The thick carbide thus formed grows rapidly in the lateral direction in a
continuous film, along the grain boundary, incorporating any thin carbides
along its way. Growth in the normal direction slows after a micron or so
thickness because of the required diffusion of Mn and Cr from the matrix
and of Si from the carbide. The thick carbides embrittle the steel either
from their poor cohesion with austenite, in which case cracks propagate
along the carbide / austenite interface, or because of their continuous nature and the cracks propagate through the brittle carbide.

QUANTITATIVE RELATIONSHIP BETWEEN EMBRITTLING PHASES


AND IMPACT TOUGHNESS
The degree of embrittlement in austenitic manganese steels depends on the degree
of grain boundary coverage by the embrittling phases. In this respect, intergran3

ular carbides play a key role but other embrittling phases such as phosphide eutectic, non-metallic inclusions and microporosity also are involved. It is important to
note that the abundance of an embrittling phase is not as important as its distribution. As low as 5 ppm bismuth can embrittle high copper alloys [12] and the presence of 60 ppm aluminum nitride as grain boundary films is sufficient to embrittle steels [13]. These constituents, even in very thin layers, cause a loss of cohesion with the matrix. Oxides have the least cohesion with a metal matrix, but
typically their distribution is in the form of random discrete particles, hence they
are not as harmful to toughness. Phosphorous is a greater concern in manganese
steels it tends to segregate at grain boundaries, liquefies during solution annealing
and forms an embrittling phosphide eutectic film. Sulphur in low carbon steels
can form Type II grain boundary iron-sulphide eutectic during solidification [14];
but in manganese steels, tends to form globular manganese sulphides which are
not particularly detrimental.
The very heterogeneous distribution of the embrittling phases in austenitic manganese steels makes it difficult to predict their impact toughness from metallographic observations. A sufficiently large area must be characterized and averaged to
achieve adequate accuracy. This was done for three steels, regular manganese
steel (Grade B3: 1.2% C, 13% Mn), high C steel (Grade B4: 1.3% C, 13% Mn),
and Cr-bearing steel (Grade C: 1.2% C, 13% Mn, 2% Cr). Charpy specimens
from these steels were tested and their ends were prepared for metallography. The
samples were examined at 200 magnification to easily distinguish the thin,
thickened and the thick carbides from each other. Although each field presented a
0.1 mm2 area, approximately 1000 fields were scanned to qualitatively establish
the grain boundary coverage by a given constituent. For rapid characterization,
albeit with some loss of precision, a procedure based on a multi-field comparison
was employed. The fraction of grain boundary covered by a certain constituent
was mentally recorded while scanning the sample by moving the microscope
stage. The overall average was recorded and the procedure was repeated for each
constituent. A trained eye can accomplish this within 10% precision in 20 minutes each sample. As each constituent embrittles Hadfields steel differently,
each is assigned a weight. For example, we know that as-cast material, completely covered with thick carbides, has toughness less than 90% of its optimum.
Therefore, the thick carbides are given a weight of 0.9. Phosphide eutectic is
about as detrimental as the thick carbides, also is weighted at 0.9. On the other
hand, a heat-treated steel displaying only the thin carbides is very close to its
optimum toughness; therefore, the thin carbides are weighted at 0.1. Thickened
carbides are considered somewhere between and are weighted at 0.5. Microporosity, of course, creates complete decohesion, and is weighted at 1.0. Therefore,
according to this scheme, a 90% coverage by the thin carbides is equivalent to
10% coverage by thick carbides.
Figure 4 shows the room temperature CVN impact toughness plotted against the
weighted grain boundary coverage in the three steels mentioned [5]. A good correlation was obtained regardless of the steel type and processing. The results can
be summarized as follows:

Weighted g.b. Coverage


< 20 %
20 - 40%
> 50%

Toughness
Excellent
Good
Poor

CVN, J
>100
>50
< 20

Impact toughness decreased at a rate of 3.5 J per percent weighted grain boundary
coverage, from a nominal value of 180-200 J.
MACRO- AND MICRO- SEGREGATION
Macro-segregation
The kinetics of carbide precipitation in manganese steels increase sharply with
carbon content, so that steels having more than 1.3% C are difficult to produce.
Any positive macrosegregation therefore would compound this problem, as this
steels poor thermal diffusivity and now higher carbon content would make it very
difficult to heat treat the section centers. However, chemical analyses of toppoured and side-poured blocks showed negligible or inverse macrosegregation.
Carbon was richer around the perimeter of the casting and leaner towards the
center [4,7]. This was true for most of the other elements, phosphorous in particular showed a clear inverse macrosegregation (Fig. 5). The problem of heat treating the thick-sectioned austenitic manganese steels is therefore strictly a heat
transfer problem.
Micro-segregation and its Effect on Phosphide Eutectic
Steels having more than 0.06% P or solution annealed above 1150C can be susceptible to phosphide eutectic embrittlement, where a phosphide-carbide eutectic
liquefies at the grain boundaries, gradually spreads along the grain boundary
during the course of heat treatment and remains in place on quenching [3]. The
as-cast structures contain this eutectic at the centers of carbide colonies as globular constituents with a mottled appearance [2,4]. In the as-cast steels, the cooling
castings do not spend as much time at high temperatures as in heat treatment, the
eutectic does not have time to spread over the grain boundaries. Phosphide eutectic embrittlement is reversible. When castings are heat treated at the correct solution annealing temperature, it re-dissolves into the austenite matrix. In severely
embrittled steels, however, a trail of Kirkendall voids and cavities are left behind
and the toughness is not completely restored [3].
Most solutes in manganese steels, notably Mn and Cr, and to some extent P, microsegregate towards grain boundaries and intercellular regions. The degree of
microsegregation increases with section size [7]. This makes the thick sections,
especially near the casting surface, where the phosphorous content and the temperatures are higher, more susceptible to phosphide eutectic formation.

MICROPOROSITY
Austenitic manganese steels have a long freezing range, and microporosity is a
common feature in these castings, regardless of feeding. Microporosity occurs in
clusters as random pockets, and its occurrence does not correlate with any thermal
criteria such as Niyamas [8]. Figure 6 shows the typical distributions of porosity
in 1-in. dia. sections in side-poured and well-fed ([Niy]>0.5)1, 4-8 in. thick, 15 in.
wide, 20 in. long wedge-block castings in regular manganese steel. The random
distribution of microporosity in clusters suggests heterogeneous nucleation. Each
cluster is interconnected through intercellular regions. Microporosity can nucleate
when the pressure drop ahead of the solidification front caused by the need to feed
the solidification shrinkage leads to the supersaturation by the gaseous solutes.
Usually, a large supersaturation and a favorable site (a non-wetting surface such
as an oxide inclusion) are required to nucleate a pore [15]. When a pore nucleates, it quickly grows to fill the interdendritic channels by the exsolution of gas
until the liquid is no longer supersaturated. This process may repeat itself at other
nucleation sites, leading to the observed microporosity clusters in a 2-D section.
Gaseous solute content, oxide inclusion content and distribution, and the solidification rate (pouring temperature, molding media, use of chills, and section size)
primarily affect the formation of microporosity. Among these, the gaseous solute
content and solidification rate are probably the more important as oxide inclusions
are a common feature in aluminum-killed steels.
Microporosity forms preferentially in interdendritic areas (Fig. 6e,f); therefore, its
grain boundary fraction is much greater than its volume fraction. Indeed, the
grain boundary fraction of microporosity as an embrittling constituent in this steel
was estimated to be 5-10% [5]. Since the embrittling phases decrease Charpy impact toughness at a rate of 3.5 J per percent coverage, the contribution of microporosity to embrittlement is expected to be a 15-30 J decrease in impact toughness
from an expected maximum of 180-200 J. Microporosity is a secondary but a significant concern for loss of toughness next to intergranular carbides.
The microporosity of 1-in. dia. core sections varied from 0.04 to 1.5% [8]. The
average porosity in the two blocks was 0.37%. As this is not large enough to
compensate for the 3% solidification shrinkage, risers are required to feed the castings.
QUENCH TIME MEASUREMENT
The quenching operation is critical in the production and heat treatment of austenitic manganese steels. The quench speed depends on the load distribution on a
pallet, initial quench water temperature, section size and bath agitation. Among

[Niy] is Niyamas criterion value (G/T/t) in (C s) mm-1, where G is the


local thermal gradient and T/t is the cooling rate.

these, the load distribution is the least effectively controlled. If the water permeation through the load is poor, then the effective section size of the load becomes
much larger, and with the thermal diffusivity of the steel being poor, the quenching operation can slow down considerably. The other process variable that is of
interest is the initial quench water temperature. Although each metal caster has its
own specifications for maximum initial and final water temperatures, in hot summer months and with large loads, these can become difficult to attain. Typically,
specifications apply to worst case scenarios (large section sizes). It is desirable to
know if a certain load can be quenched (or has been quenched) even though the
initial water temperature is (or has been) above the specified limit.
The steel chemistry plays an important role in determining the required quenching
speed. Increasing carbon content increases the rate of carbide precipitation, and
steels having more that 1.3% C are difficult to make. Chromium also increases
the carbide precipitation rate by increasing the carbide stability and causes carbides to start forming at higher temperatures.
The measurement of quench water temperatures provides a cost-effective means
to evaluate a quenching operation. As castings cool, the quench water warms;
hence, there is a one-to-one temperature relationship between the two. Quench
water temperatures and heat capacity data of the steel, can be used to derive the
average temperature of the castings at a given time during the quench, starting
from their furnace (solution annealing) temperature. A useful, single-valued determination that lends itself to comparison would be the time to extract 90% of the
heat from the load. This is schematically shown in Fig. 7 for a laboratory quench
[3]. The total temperature rise in the quench water corresponds to the total heat
taken from the castings. Therefore, 90% of that rise corresponds to 90% heat extraction. From the construction in the graphic, 90% heat extraction (t90%) in this
quench took 4.2 minutes. In regular manganese steels having 1.2% C, satisfactory
toughness (>100 J) is obtained if t90% is kept below 10 mins [3]. In higher carbon
manganese steels (>1.3% C), the required t90% decreases to less than 5 minutes.
Figure 8 shows an industrial quench where there is continuous water circulation
from a cooling tower [6]. The cooling rate of the quench tank ( T ) attributable to
the cooling water is given by:

M W CW T m CW (T TCW )
m
T
(T TCW )
MW

(1)

where m is the cooling water flow rate, MW is the mass of quench water, TCW is
the cooling water temperature as it enters the tank, and T is the tank temperature.
The cooling water flow rate is usually known, but if not, it can be estimated from
the data. At the end of a quench, when the castings are sufficiently cool, their
contribution to heating the tank could be neglected and the tank temperature is
solely affected by the cooling water. In Figure 8, 30 minutes after the quench, the
tank cooling rate was 0.136C/min., the tank temperature was 23.4C, the cooling
water temperature could be taken to be the tank temperature prior to the quench,
7

20C. The 171213 ft. quench tank contained 75,100 kg water. Using this information in Eqn. 1 yields a cooling water flow rate of 3000 kg/min. This figure
is probably more reliable than a nominal design value and could be used to double
check for any blockages in the system. The water temperatures must be adjusted
by adding the effect of cooling water:
Ti

m
(Ti TCW ,i ) ti
MW

(2)

The adjusted water temperature curve thus obtained represents the water temperature in the absence of any cooling. The quench times can now be found by a
similar construction as in Fig. 7.
The precision of the quench time analysis can be improved by taking temperatures
in more than one location in the tank and averaging the results, the cooling water
temperature as it enters the tank, and the ambient temperature. The data can be
further processed to obtain cooling rates that show transitions from vapor blanket
to nucleate boiling, and lastly convective cooling, and to conduct quench factor
analysis using information from an isothermal transformation diagram [9].
CONCLUSIONS
Heat treatment processing of austenitic manganese steels has been studied with
the following results:
Intergranular embrittlement by hypereutectoid carbide precipitation. The detrimental thick carbides have been distinguished from the relatively harmless
thin carbides. A model has been proposed for the different types of carbide formation. A quantitative relationship has been established between impact toughness and grain boundary coverage by the embrittling phases.
Chemical segregation. Macrosegregation was either absent or negative (concentrations were higher around the perimeter of the castings than in the center). As
carbon and chromium are not any richer in the center, the thick section problem in
quenching the manganese steels is strictly a heat transfer problem. Microsegregation increases with section size, and together with inverse macrosegregation, compounds the phosphide eutectic embrittlement around the casting perimeter.
Microporosity and structural soundness. Microporosity was observed as randomly and heterogeneously distributed colonies in castings. Although its volume
fraction averaged 0.37%, it had a preferred location at intergranular and intercellular regions, where its grain boundary fraction was 5-10%. Its embrittling
effect was secondary to carbides, decreasing the impact toughness by an estimated
15-30 J.
Quench time measurement. A cost-effective procedure was developed making
use of quench water temperatures, to quantitatively assess the quenching operation.
8

ACKNOWLEDGEMENTS
The authors would like to thank Mark Charest for TEM replica study and Cathy
Bibby for sample preparation.
REFERENCES
1) KUYUCAK S, Newcombe P and Zavadil R On the heat treatment of Hadfield's austenitic manganese steels, Part I: step-down temperature and residuals. AFS Trans., 108, Paper No. 00-154, (2000).
2) KUYUCAK, S and Zavadil, R On the heat treatment of Hadfield's austenitic manganese steels - Part II: metallographic studies. AFS Trans., 108,
Paper No. 00-126, (2000).
3) KUYUCAK S, Zavadil R and Newcombe On the heat treatment of Hadfields austenitic manganese steels, Part III: heat transfer model, macrosegregation and phosphide eutectic. AFS Trans., 109, Paper No. 01-117, (2001).
4) KUYUCAK S and Zavadil R On the heat treatment processing of austenitic
manganese steels. Proc. 21st Heat Treat Conf., 5-8 Nov. 2001, ASM International, Paper No. 16.
5) KUYUCAK S and Zavadil R On the heat-treatment of Hadfields austenitic
manganese steels, Part IV: microstructure vs. impact toughness relationship.
AFS Trans., 110, Paper No. 02-116, (2002).
6) KUYUCAK S and Newcombe P On the heat-treatment of Hadfields austenitic manganese steels; Part V: measuring quench time in industrial castings.
AFS Trans., 111, Paper No. 03-102, (2003).
7) KUYUCAK S and Zavadil R On the heat-treatment of Hadfields austenitic
manganese steels; Part VI: impact toughness, microstructure, macro- and
micro- segregation in large wedge-block castings. AFS Trans., 111, Paper
No. 03-133, (2003).
8) KUYUCAK S, Zavadil R and Ouellet G On the heat-treatment of Hadfields austenitic manganese steels; Part VII: casting soundness and microporosity in large wedge-block castings. AFS Trans., 111, Paper No. 03-134,
(2003).
9) KUYUCAK S, Newcombe P, Bruno P, Grozdanich R and Looney G Measurement of quench time as a process control tool. Proc. Technical and Operating Mtg., Nov. 2003, Steel Founders Society of America.
10) KUYUCAK S, Gertsman V Y and Zavadil R On the heat-treatment of Hadfields austenitic manganese steels; Part VIII: studies on microcharacterization. AFS Trans., 112, Paper No. 04-129, (2004).
11) RAYNOR G V and Rivlin V G in Phase Equilibrium in Iron Ternary Alloys, The Institute of Metals, London, (1988), pp. 172-173.

12) SADAYAPPAN M, Zavadil R and Sahoo M Effect of impurity elements on


the mechanical properties of aluminum bronze alloy C95800. AFS Trans.,
107, pp. 329-336, (1999).
13) WOODFINE B C and Quarrell A G . Effect of Al and N on the occurrence
of intergranular fracture in steel castings. J Iron and Steel Inst., 195, (1960),
pp. 409-414.
14) EEGHEM V J and DeSy A Side effects of cast steel deoxidation. AFS
Trans., 72, (1964), pp. 142-148.
15) CAMPBELL J in Castings, Butterworth-Heinemann, (1991), p. 203.
16) PIWONKA T S, Kuyucak S and Davis K Shrinkage-related porosity in
steel castings. AFS Trans., 110, Paper No. 02-113, (2002).

10

Fig. 1. Intergranular, hypereutectoid carbides in austenitic manganese steels [2,5].


(1a,b) regular manganese steel. Arrows show transition from thin to thick carbides where a double austenite / carbide phase boundary on the carbide film becomes visible. (1c,d) Further examples from regular manganese steel: carbidefree and thin carbide grain boundaries (1c), thick carbide laterally growing into a
thin carbide (1d). (1e,f) Cr-bearing manganese steel. The thickened carbides
in (1e) do not show a clear austenite / carbide phase boundary, whereas the
thick carbides in (1f) do. Etched in equal parts of water, conc. HCl and conc.
HNO3.

11

Fig. 2. Transmission electron microscopy images of carbides in Cr-bearing manganese steel [10]. (2a-d) carbon replica images, (2e-g) thin foil images. (2a) thin
carbide, (2b) thick carbide, (2c,d) thickened carbide, (2e) thin carbide resolved
into discreet particles (arrows indicate growth direction), (2f) thickened carbides
(the extended tips in the growth direction are caused by dislocations), (2g) thick
carbide showing as a continuous film.

12

14

a
6

Mn

Mn (wt.%)

10

4
C

Cr

0
-0.2

0
-0.1

0.1

Cr, C (wt.%)

12

0.2

Mn (wt.%)

20

10

16

12

Mn

4
Cr

Cr, C (wt.%)

Distance (m)

C
0

0
-3

-2

-1

Distance (m)

Thickness
(m)
Grade C steel
Thin carbide
Thick carbide

0.07
0.8

C
1.2
4.5 (6.8)
6.0 (6.8)

Composition (wt. pct.)


Mn
Cr
12.4
1.9
13
2.5
18
6

Fe
83
78
69

Note: carbon content measured by the TEM nanoprobe is lower than the expected
amount from the stoichiometry of cementite (given in parentheses). This is almost
always the case for EDX analysis when a light element is present along with heavier elements in the compound. Iron is from balance.
Fig. 3. EDX profiles across thin (3a) and thick (3b) carbides in Grade C steel
[10].

13

200
Regular manganese steel, Grade B3
High C manganese steel, Grade B4
Cr-bearing manganese steel, Grade C
Charpy Toughness, J

150

100

50

0
0

20

40

60

80

100

Weighted Grain Boundary Coverage, pct.

Fig. 4. Correlation between weighted grain boundary coverage by the embrittling


constituents and impact toughness [5].
0.08
5 in.

6 in.

7 in.

Block 1
Block 2

6 in.
Pct. P

0.06

5 in.
7 in.

0.04

0.02
-3.5
Drag

-2.5

-1.5

-0.5

0.5

1.5

Distance from Center, in.

2.5

3.5
Cope

Fig. 5. Macrosegregation behaviour of phosphorous in side-poured 4-8 in. thick,


15 in. wide, 20 in. long wedge-blocks (1.2% C, 13.7% Mn, 0.6% Cr, 0.3% Mo,
0.050% P) [7]. Samples were taken from 5-in., 6-in. and 7-in. sections.

14

Fig. 6. Observed microporosity in sections taken from 1-in. dia. core samples, cut
from wedge-block castings (6a-d). (6a) B1-7-4-6 (key: 1 st block, core taken from
7-in. section, 4 in. from left, sectioned at 6 in. from drag). (6b) B2-7-6-7, (6c) B16-2-2, (6d) B2-7-4-2, from [8]. (6e) Intergranular and intercellular microporosity
in a 1-in. thick block casting [16]. (6f) An extreme example of intergranular
porosity [5]. (6a-e) as-polished, (6f) etched in equal parts of water, conc. HCl and
conc. HNO3.

15

Water Temperature, C
30
25
20
15

T90%

10
5

t90%= 4.2 min.

0
0

10

Quench Time, minutes

F9034-HT.xls

Fig. 7. Construction to determine quench time from quench water temperatures in


a laboratory quench tank [3].

Block 2, Quench Severity H = 3.0 per in.

Quench water temperature, C

40

Adjusted Temp.

T max = 37C

36

T 90% = 35.5C
Actual Temp.

32

Cooling rate:
0.25C/min at 32C

28
24

T i = 22C
20
t 90% = 6.3 min
16
0

10

15

20

Quenching time, minutes

25

30
Sivyer HT.xls

Fig. 8. Quench time determination in an industrial quench tank with cooling


water recirculating from a cooling tower [6].

16

S-ar putea să vă placă și