Hydrogen in Aluminum Lithium Alloys

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

3 Hydrogen in Aluminum–

Lithium Alloys
During melting process of aluminum alloy containing lithium, without any type of
protection of the surface against a contact with the atmosphere of the furnace, there
is an interaction with oxygen, water vapor, carbon oxide, and dioxide (Table 2.5). As
it has been discussed in Chapter 2, films formed on the surface are not protective in
their nature [76,77], which is why hydrogen created by the reaction with water vapors
will be enriching the molten. Lithium, being a hydride-forming element, should
increase the solubility of hydrogen in aluminum [78]. Magnesium also increases the
solubility of hydrogen in aluminum [79].
During the first years of mastering the aluminum–lithium production alloy, a
melting unit in a composition with reverberatory furnace and holding furnace with
flame heating was used to prepare them [40,41,66]. Processes of gas saturation of
the molten in such a unit do go in a more intensive way in the beginning moment
of the charge-melting, when as a result of creation and dripping of the liquid metal
drips the contact area with the atmosphere does increases. Due to the technology
of melting of the aluminum alloys with lithium, it is anticipated that there will be
some scrap of these alloys in the charge, up to 50%–70%, in the flame furnace, and
despite the rational charging process, during the heating-up process and melting of
the scrap, there will be an intensive process of interaction of alloys in the furnace.
A contact between an opened surface of the molten with the atmosphere can also
be done in the moment of breaking the evenness of the flux covering in the holding
furnace when lithium is introduced into the molten and when taking off the slag.
During all these operations, alloys will be enriched with hydrogen and other ele-
ments coming from the furnace atmosphere. In this chapter, we present the results
of the research of how hydrogen behave in aluminum alloys with lithium and the
ways to diminish it.

3.1  INTERACTION OF ALUMINUM ALLOYS WITH HYDROGEN


A lot of research projects are dedicated to studying how aluminum and its alloys
interact with hydrogen [80–82]. Aluminum in solid state begins to interact with
hydrogen; the speed of this interaction increases as the temperature increases, and
aluminum easily dissolve hydrogen in liquid form in conjunction with the following
formula:

Al liquid (solid) + H2 ↔ 2H (soluble in Al) (3.1)

37
38 Aluminum–Lithium Alloys

While the temperature is constant, this process can be described by the formula of
Siverts:

S = Ks (P)½ (3.2)

Solubility of hydrogen in liquid aluminum is being substantially affected by an


additional agent. According to the project [82], the solubility of hydrogen in liq-
uid alloys of aluminum–magnesium system is steadily increasing from aluminum
to magnesium.
It was shown later [62,83] that at temperatures 660°C–700°C, the solubility of
hydrogen in aluminum increases almost lineally with the content of magnesium. At
temperature 700°C for aluminum solubility equals to 0.9 cm3/100 g, in Al + 4% Mg
alloy—1.6 cm3/100 g, and in Al + 6% Mg alloy—2.0 cm3/100 g.
The solubility of hydrogen in solid-state aluminum is very little and contains
0.015 cm3/100 g, but it increases due to the presence of magnesium in aluminum.
For example, aluminum alloy with 5.25% Mg, the content of hydrogen increases
to 0.06 cm3/100 g, which is firstly related to high solubility of hydrogen in Mg2Al3
composition, which amounts to 1.45 cm3/100 g at melting temperature in solid state
and 5.9 cm3/100 g in liquid state [80]. The solubility of hydrogen in this intermetallic
state is subject to Siverts law, at least up to 600 atm. This proves that hydrides in this
intermetallic state are not created until such pressures are reached.
The equilibrium solubility of hydrogen in liquid and solid metal was shown ear-
lier. In real production of semifinished aluminum alloy products, hydrogen generally
creates supersaturated solid solutions [80]. But in solid solution, not all the hydrogen
present in liquid metal is fixated since a part of it has time to separate during crystal-
lization in shrinkage cavities or creates primary gas porosity.
Thus, hydrogen in hard aluminum and its alloys can be present in aluminum
solid solution, being dissolved in intermetallic phases, creating civilities filled with
molecular hydrogen.
The main source of hydrogen can be both atmospheric moisture and moisture cre-
ated during fuel combustion in the process of open flame metal heating. Partial pres-
sure of the water vapor amounts to 13–15 mm of mercury column in a lab electrical
furnace and 60–120 mm of mercury column in a gas furnace.
Hydrogen enrichment of aluminum is done while there is an interaction with
water vapor according to the following formula:

3H2O + 2Al = Al2O3 + 6H (3.3)

Atomic hydrogen is partially absorbed by aluminum and partially recombined into


molecules and goes into the atmosphere:

3H2O + 2Al = Al2O3 + 3H2 (3.4)

Oxide film (hydroxide) is created simultaneously with hydrogen, which slows down
the interaction reaction, due to its impenetrability for water molecules. Atomic
Hydrogen in Aluminum–Lithium Alloys 39

hydrogen from water vapors easily penetrates into the molten. It has been defined
that the concentration of hydrogen in the solution (H) is proportional to the pressure
of atomic hydrogen:

[H] = КP ½ H2O (3.5)

where
К is constant dependable on temperature
PH2O is partial pressure of water vapors in the atmosphere

This formula is similar to Siverts formula for solution of hydrogen in aluminum at its
interaction with molecular hydrogen at K coefficient being another value.
In a number of projects, a justification test of the Siverts correlation for hydrogen
solubility in aluminum and its alloys during their interaction with water vapor was
carried out, where a fact was established that hydrogen content increases in metal
along with the increase in air humidity [80,64,83].
The content of hydrogen in moltened aluminum alloy changes in the direction of
reaching the equalization with atmosphere surrounding the molten, which depends
on alloy content, atmosphere temperature, and humidity. The speed of achieving the
balance, besides that, is being affected by properties and condition of the film sur-
face on the molten and also by the depth of the bath.
Electrical holding furnaces are good for getting the molten with lower hydrogen
content in comparison with flame holding furnaces, especially in the case of alloys
with weak protective properties of the surface films, for example, for alloys with
magnesium.
As shown in projects [59,64], the process of hydrogen elimination and its absorp-
tion by the molten is being largely affected by both surface film on the molten and
aluminum oxides present in the alloy in view of the suspension. The degassing pro-
cess during storage is more intensive, the lesser is the content of the suspended oxide.
Due to higher density of oxides (ρ = 3.6 g/cm3), aluminum alloys settle in a bath of
aluminum molten (ρ = 3.6 g/cm3, viscosity η = 1.2 Pa at 700°C). The rough inclusion
particles settle, while smaller ones go up on the surface.
Calculations made [62,84] have shown that the most dispersive particles (dэ =
2–10 μm) are located evenly within the alloy mass and that allows us to consider
the liquid–metal system, as colloid system, which is thermodynamically unstable.
Insoluble disperse nonmetallic inclusions particles which have high interphase
energy at the border with liquid-metal, while being stored should by itself coagulate,
which is restricted by their lesser mobility, wedging pressure due to interaction of
double electric layers and viscosity increase of surface layers by means of absorption
of covering-active substances on them. Turbulization of liquid metal creates good
conditions for increasing the size of nonmetal components in aluminum molten.
In the project [64], it was established that increasing the content of oxide alumi-
num in aluminum and its alloys increases the content of hydrogen. The presence
of liquid structure in small-dispersed oxide aluminum substantially slows degas-
sing of aluminum moltens [62,84]. Different researchers have different views on the
mechanism of hydrogen and aluminum oxide interaction. Some of them look at that
40 Aluminum–Lithium Alloys

process only from a thermodynamical point of view, believing that heterogenic par-
ticles are a ready-made surface, which makes it easy to create a new phase—a bubble
of hydrogen [86]. Others think that aluminum oxide particles make an electrostatic
impact on hydrogen and create complexes such as H2–Al2O3 [78]. The amount of gas
hydrogen, held in molten on oxide and other nonmetallic particles surface, is defined
by its morphology [85].
New ideas about the behavior and interaction of nonmetallic impurities in alumi-
num moltens were developed by Makarov [62,84]. It is stated there that most of the
hydrogen in aluminum molten is in the dissolved state, the percentage of molecular
hydrogen is negligible, but its role is significant as, being on the inclusion particles,
molecular hydrogen is a major factor in stability of the system “molten aluminum–
nonmetallic impurities.”
The number of oxides in aluminum and its alloys (Table 2.10) can create a
substantial effect on the system “molten hydrogen” [34,88]. The concentration
of hydrogen met in aluminum molten during preparation of alloys, as a general
rule, is always lesser than its solubility at regular pressure rate, which is why
the main hydrogen mass is in the form of a solution. However, at the surface of
the hard particle, there is an absorption monolayer of atomic hydrogen, which
creates a field of adsorption forces. At 0.01% oxide concentration in the molten,
these power fields are closed, which slows down the process of decomposition of
supersaturated solution, making discharge of the hydrogen as hard as absorption.
The amount of hydrogen absorbed by the surface of oxide, with contents of 0.01%,
amounts to 0.014 cm 3/100 g in molten. Besides that, molecular hydrogen resides in
pores, cracks, and other dents of oxide film, in the amount of ~5 × 10 –4%, which
delays sedimentation of oxide and stabilizes the system aluminum molten–oxide
phase.
The furnace atmosphere impacts substantially on the level of hydrogen con-
tent and contamination of nonmetallic particles. Furnace content depends on the
type of the melting machine: in electrical furnaces, it is close to the content of
the air atmosphere; in flame reverberatory furnaces, it is defined by the content
of the fuel, mode of combustion, and amount of air let in for the fuel combustion
[34]. The analysis of the furnace atmosphere, which can have oxygen, hydrogen,
nitrogen, gas compounds CO, CO2, H 2O, Cm H n, SO2, witnesses about its oxidative
nature.

3.2 CHANGING OF HYDROGEN CONTENT IN


ALUMINUM–LITHIUM ALLOYS DEPENDING
ON THE CONDITIONS OF MELTING
Due to the high oxidation of aluminum alloys with lithium, it appeared impossible
to try a Dardell–Gudchenco liquid test method to find out the content of hydrogen
in the molten. Sampling of test material from molten and sample casting into the
moulds of existing constructions to carry out the hard test analysis yielded no posi-
tive result due to the very large range of values until the latest times. All data about
hydrogen content in aluminum–lithium alloys have been received during analysis of
Hydrogen in Aluminum–Lithium Alloys 41

TABLE 3.1
Hydrogen Content in 1230 and 1240 Alloy Billets Depending on Melting
Conditions
Melting Conditions Hydrogen
Size of Content
Alloy Type of Furnace Atmosphere Additional Treatment Billet (cm3/100 g)
1230 Gas forge Gas combustion No 70 1.03–0.87
products
Induction furnace, Helium No 90 0.5–0.4
16 kg capacity
Induction furnace, Helium Vacuuming with residual 90 0.3
16 kg capacity pressure 1 mmHg;
soaking 30 min
1420 Gas forge Gas combustion No 70 1.3–0.9
products
Induction furnace, Helium No 50 0.4–0.3
16 kg capacity
Induction furnace, Helium Vacuuming with residual 50 0.3
16 kg capacity pressure 1 mmHg;
soaking 30 min

samples cut out of billets, by vacuum extraction method, both in the present research
and in consequent ones in industrial conditions.
Lack of analysis methods, allowing to control the changing of hydrogen content
during the technological process, creates obstacles on the way of researchers study-
ing kinetics of hydrogen content changing in alloys during the process of melting
depending on modes of molten preparation and their treatment methods.
First data received in the beginning of the 1970s have showed (Table 3.1) that in
experimental billets of 1230 and 1420 alloys, the content of hydrogen substantially
exceeds the level, which is indicative for serial aluminum alloys [85]. In the mean-
time, during experimental hot melting, a possibility to lower that level was defined
using remelting of alloys in inert atmosphere or using additional vacuum treatment
(Table 3.1).
To explain the data received and to define the ways to lower the content of hydro-
gen in aluminum–lithium alloys, the analysis of sources of hydrogen enrichment
was carried out and a hypothesis was put out about special terms of enrichment
of aluminum alloys, containing lithium, by hydrogen and possible forms of their
presence.

3.2.1  Sources of Hydrogen in Aluminum–Lithium Alloys


It is shown in Table 3.2 that the basic main sources of hydrogen in aluminum–lithium
alloys could be flux agent, magnesium, ligatures, scraps of lithium-containing alloys,
42 Aluminum–Lithium Alloys

TABLE 3.2
Sources of Hydrogen, Sodium, and Alkaline Earth Metals in Aluminum–
Lithium Alloys
Hydrogen Sodium Calcium
Moisture Increment Increment
Content in Alloy Content Increment Content in Alloy
Source (g/m3) (cm3/100 g) (%) in Alloy (%) (%) (%)
Atmosphere of 30–100 0.5–1.5 — — — —
flame
reverberatory
furnace
Air atmosphere 2–15 0.2–0.3 — — — —
Carnallite 2 0.1–0.2 Not
(3–5 kg/t) specified
— — 3–5 0.001–0.003 0.03
Technical 1% — Not
potassium specified
chloride — — Up to 3.7 0.001–0.005 1.0
(15 kg/t)
Lithium Not — 0.05 0.001 0.01
chloride specified 0.1 0.002 0.03
(15 kg/t) — — 0.002–0.027 0.03
Dry lithium 0.36 0.04 0.008 0.005
chloride
Flux type VI-2 3 0.1–0.3 10
(0.5–1 kg/t) — — 3.5 0.0004
Argon 0.01 — — — — —
200 PPm
Chlorine 1.6 0.1–0.2 — — — —
Aluminum — 0.1–0.2 0.002 0.0006–0.002 0.001 0.001
0.1–0.2 0.001 0.0004–0.001
0.1–0.2
Lithium LE1 — Up to 0.1 0.02 0.0004 0.03 0.0006
— Up to 0.1 0.005 0.0001 0.03 0.0006
Magnesium — Up to 0.5 0.01 0.0005
0.005 0.0002
Ligatures — 0.1–0.3 0.001–0.002 Up to 0.0004

specially feedheads and butt ends, covered by hygroscopic flux. The difference in
quality of charging materials has not been pointed out in Table 3.2. This will be
reflected in the later text.
Increased inclination for hydrogen enrichment of aluminum–lithium alloys is due
to the following factors:
Hydrogen in Aluminum–Lithium Alloys 43

• The film created on the melt surface is not protecting the metal from a con-
tact with furnace atmosphere moisture.
• Lithium and manganese increase the solubility of hydrogen in liquid aluminum.
• Lithium hydride is created in the presence of hydrogen.

This gave us a reason to suppose that hydrogen in aluminum alloys with lithium
can be in three states: in solved state, bound to hydrides, and in molecular form. At
that in the melt- in solved state and in bound in to hydrides, in solid—dissolved and
bound in to hydrides, in molecular form.

3.2.2 Hydrogen in 1420 Alloy during Melting in Reverberatory Furnace


Statistics data analysis, received from 1976 to 1986 at Kamensk Uralsky Metallurgical
Works, has shown that in 1420 alloy the hydrogen content in melt in a flame heat-
ing furnace is a lot higher than in commercial aluminum alloys melt at the same
conditions [85–87]. It varies within the wide range (Figure 3.1) and has a seasonal

100

80
Number of cases, %

60

5 4 3 2 1
40

20

(a)

100

80
2
Number of cases, %

5 3 4 1
60

40

20

0.3 0.5 0.7 0.9 1.1 1.3 1.5


(b) Hydrogen content, cm3/100 g

FIGURE 3.1  Statistical data about the hydrogen content in 1420 alloy, melt at flame heating
furnace: 1—1976 year, 2—1977 year, 3—1979 year, 4—1981 year, 5—1985 year. (a) begin-
ning of melting (b) after 16 h of settling.
44 Aluminum–Lithium Alloys

100

80

Number of cases, %
60

40
4 3 2 1

20

0.3 0.5 0.7 0.9 1.1


Hydrogen content, cm3/100 g

FIGURE 3.2  Seasonal changes in hydrogen content in 1420 alloy during melting under flux
in a reflecting holding furnace with flame heating: (1, 3)—summer, (2, 4)—winter, (1, 2)—
beginning of melting, (3, 4)—after 16 h of settling.

dependence (Figure 3.2) It was also pointed out that a substantial decrease of hydro-
gen content from 1976 to 1985 years was found while melting in the same furnace.
The observed decrease of hydrogen content in 1420 alloy year by year (Figure
3.1) is related to the improving technological process, aimed at lowering hydrogen
content in aluminum–lithium alloys, such as argon dehydration, flux remelting and
dehydration, development of new flux mixes, changing of the technology of pouring
of molten, vacuum treatment of the molten, and usage of charging material of higher
quality.
Decreasing hydrogen content, achieved in 1977 as compared with 1976, is con-
nected with transition to apply liquid flux in the holding furnace. At the end of 1978,
introduction of zirconium began to take place using aluminum–zirconium ligature
instead of magnesium–zirconium ligature, which has a higher pollution rate. It
caused a substantial decrease of hydrogen content in 1420 alloy in 1979.
Since 1981, metallic lithium is supplied in the form of small, up to 4 kg, bil-
lets, being put inside metal containers, filled with argon under excess pressure
of 0.07 MPa. Before lithium was supplied in the form of 150 g billets, it is put in
white cans, filled with the mix of paraffin with transformer oil. It was necessary
to clean up billets from grease. This is the reason why grease was a significant
source for hydrogen enrichment. Besides this weighing of lithium, its operation of
introduction into the molten was long-lasting and caused substantial quality loss
of flux covering.
At the same time with purity increase of metal lithium, the quality of lithium
chloride became better, which began to be supplied dehydrated in granular form.
Hydrogen in Aluminum–Lithium Alloys 45

However, having lowered the content of hydrogen in 1420 alloy, these events did
not exclude the cases of getting the metal with higher hydrogen content in spite of
molten chorine treatment in the holding furnace.
A wide range of values within the period of 1 year is related to the number of
factors, such as seasonal change in atmosphere humidity of working space in fur-
nace and charging material (Figure 3.2) and temperature changes of molten in the
melting unit. Besides that, a great influence was born by usage of nonmelt flux to
reinstate the outside flux covering in the holding furnace, and usage not sufficiently
dried tools.
Work experience in the conditions of commercial production shows that alumi-
num alloys with lithium, due to its high activity, are incomparably more sensible to
any violations in technology than regular commercial alloys.
From the analysis carried out, it follows that the technology of lithium alloys
preparation, as well as preparation of commercial alloys in machines with flame
heating without the following refining treatment, does not guarantee getting a
semifinished product with regulated hydrogen content [62]. As a general rule,
it was correctly supposed that the more molten is contaminated, the lesser its
cleanliness becomes after refining in the same conditions (Figure 3.3). To provide
execution of exclusive cleanliness requirements as to the nonmetallic inclusions
and hydrogen, can be achieved by excluding the contact of melt with furnace
atmosphere and by applying of effective refining and degassing methods.

0.8

0.6
Holding effect, cm3/100 g Me

0.4

0.2

–0.2

–0.4

0.2 0.4 0.6 0.8 1.0 1.2


Horig, cm3/100 g Me

FIGURE 3.3  Dependence of hydrogen content decreasing effect from its original content at
1420 alloy storage in flame heating furnace during 16 h at temperatures 710°C–740°C.
46 Aluminum–Lithium Alloys

3.2.3 Effectiveness of Vacuuming of Aluminum–Lithium Alloys


As has been defined by tests and has been confirmed in production environment
[22,56,57,85,87], the best protection of aluminum–lithium alloys while melting is the
environment of dry argon, and an effective method to decrease hydrogen is to treat
the molten with vacuum.
To get aluminum–lithium alloys with regulated hydrogen content at Kamensk
Uralsky Metallurgical Works, a melting–casting unit was erected accompanied by
coreless induction furnace and vacuum holding furnace. Coreless inductive furnace
provides better preparation conditions for alloys in comparison with gas-heated
reverberatory furnace. In coreless furnace, charging material is introduced into a
liquid swamp, and the flashing process of charging material begins at the bottom.
This sharply decreases drop and run-off of liquid metal, there are no surface over-
heating of metal in the area of contact with furnace atmosphere, in content furnace
atmosphere there are no products of organic fuel combustion, this provides smaller
ratio of molten area to the bath depth. All these reduce metal contact with furnace
atmosphere, decrease metal losses in comparison with flame machine from 10% to
5%, and increase metal cleanliness as regards to nonmetal impurities (Table 2.9).
Design of the furnace, which has a massive lid, allows doing melting either under
flux or in inert gas atmosphere.
In technical literature, the question of vacuum treatment of aluminum alloys was
widely discussed and was generalized by Makarov [62,84]. It was demonstrated that
holding of the molten both in inert gas and in vacuum, in general, was in the area of
diffusive mode. Faster degassing can be achieved by rabbling of the molten, of which
the degassing process approaches the area of the kinetic mode. Special design of lin-
ing of the vacuum holding furnace provides barbotage of discharged gas through the
molten in the process of vacuuming.
The presence of oxygen or water vapor in barbotaging refining gas lowers the
cleaning effect. The presence of chlorine and carbon-active additions in refining
gas speeds up the degassing process. Cleaning of aluminum molten from hard non-
metallic components is produced by the extraction and development of molecular
hydrogen on them, and also by floating of bubbles of refining gas during barbotage.
Despite the large number of published research works, it should seem that in
quite detailed analysis of available data, not everything can be considered as exactly
resolved. For example, regarding the vacuuming temperature, we see data requir-
ing to conduct that process at higher temperatures, which increases the amount
of removed hydrogen by 2.5–3 times, while temperature increases from 700°C to
900°C. However, according to data [89], greater density of metal was observed after
vacuum degassing at 760°C, rather than at 800°C, but the authors of the work [24]
stated that the highest effectiveness of vacuuming was achieved at temperatures,
close to equilibrium solidus.
The best conditions for degassing are provided at relatively small depth and big
molten surface because the time of hydrogen transition in the bath space decreases,
which limits the time of molten soaking in vacuum conditions [84]. The litera-
ture data stated relates to aluminum alloys, which are less chemically active than
­aluminum–lithium alloys, in which the hydrogen is not bound into hydrides.
Hydrogen in Aluminum–Lithium Alloys 47

100
3

80

Number of cases, %
60
1

40 2
4

20
0.2–0.3

0.3–0.4

0.4–0.5

0.5–0.6

0.6–0.7

0.7–0.8

>0.8
Hydrogen content, cm3/100 g

FIGURE 3.4  Dependence of hydrogen content in 1420 alloy on the treatment mode in vac-
uum holding furnace: 1—vacuuming at P = 15–20 mmHg, settling in nondried argon; 2—
vacuuming at P = 2–5 mmHg, settling in nondried argon; 3—vacuuming at P = 2–5 mmHg,
settling in dried argon; 4—vacuuming at P = 2–5 mmHg, introduction of dry argon, pumping
out to P = 2–5 mmHg, settling in dry argon until the end of molten pouring.

During the first 34 moltens of the 1420 alloy, in a new unit consisting of inductive
coreless furnace and vacuum holding furnace at residual pressure 15–20 mmHg, the
level of hydrogen content in cast billets was high (Figure 3.4).
Decreasing of residual pressure to 2–5 mmHg in holding furnace at vacuuming
did not allow getting the 1420 alloy billets with stable low hydrogen content, which
was increasing while settling of the molten in the holding furnace at argon atmo-
sphere during 12 h (Figure 3.4, curve 2).
Determination of argon humidity with stationary hygrometer has shown that
humidity of argon, introduced into the holding furnace, is periodically increasing up
to 100–200 ppm, and which caused enrichment of the molten by hydrogen.
It is reasonable to mention that argon used for Amg6 alloy treatment did not
lead to significant increase of hydrogen content. This can be due to the fact that
1420 alloy is considerably more active than AlMg6 alloy owing to the presence
of lithium in it, also in beryllium present in AlMg6 alloy in thousandths per-
cent assist in creation of protective film, isolating the molten from contact with
holding furnace atmosphere. The film on 1420 alloy does not have any protective
characteristics.
The data received showed the necessity to submit exclusive requests to the
humidity of argon. To dry up the argon at the entrance of the holding furnace, a
drying device was installed, filled with silica gel or zeolite, and an uninterrupted
humidity control of the argon being introduced into the holding furnace was put
into place. In the following years, there was a great work done on choosing the
48 Aluminum–Lithium Alloys

heightened quality of argon supplier, and beginning the year 2005, a need in its
drying up was cancelled. At present, argon humidity is not higher than 3 ppm and
is actually less than 1 ppm.
Usage of dry argon helped to lessen the concentration of hydrogen in molten after
the vacuum treatment. But there was a repeat enrichment of the molten with hydro-
gen during settlement in a holding furnace in the process of pouring (Figure 3.4,
curve 3).
There was an operation of letting in the argon after the vacuuming and its repeated
pumping out up to residual pressure less than 2 mmHg with consecutive letting in
the argon up to working pressure of 0.0–1.0 atm and molten settling before pouring
during 40 min. Using this mode, we were successful in getting the lowest hydrogen
content and eliminating the repetitive enrichment of molten by hydrogen during its
settlement in a holding furnace while pouring it (Figure 3.4, curve 4). The lowest
values of hydrogen content (at level 0.2–0.3 sm3/100 g) were obtained as a result of
vacuum treatment at higher temperatures (750°C–760°C).
Unfortunately, it is impossible to evaluate the effectiveness of each technological
operation due to the lack of methodology of hydrogen express analysis.
All data stated regarding the content of hydrogen have been received from hard
samples, taken from cast billets. During casting, a change of hydrogen content in
molten can occur due to repeated enrichment of molten by hydrogen in inlet tray
and in casting mold, where a protection of the metal mirror is done by flux of
a developed compound [116] and then lowered due to intensive extraction while
crystallizing.
Research works done allowed creating a commercial technology of aluminum–
lithium alloys refining. In Figure 3.5, a statistics of hydrogen content in 1420 alloy
for 2008–2011 years is shown (881 cast billets) while in Figure 3.6, a statistics of
hydrogen content in 1441 alloy for the year 2011 is presented.

100
90
80
Number of cases, %

70
60
50
40
30
20
10
0
0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00
Hydrogen content, cm3/100 g

FIGURE 3.5  Hydrogen content in cast billets of 1420 alloy for 2008–2011 years.
Hydrogen in Aluminum–Lithium Alloys 49

16
14
48%
12
Quantity of events 10
40%

8
6
4 12%
2
0
0.13–0.22 0.221–0.31 0.311–0.45
Measurement range

FIGURE 3.6  Hydrogen content in cast billets from 1441 alloy for 2011.

3.2.4 Electroflux Refining of Aluminum–Lithium Alloys


The system of outside-furnace electroflux refining (EFR) developed in the 1970th
was widely used while pouring of traditional aluminum alloys allowing to decrease
the pollution of the molten by 10–15 times and reduce hydrogen content by 20%–
30% [33]. EFR method is ideal for molten refining at flux aluminum–lithium alloys
billet casting technology due to the flux being present as a protective substance in a
tray and casting mold. While used for refining of aluminum–lithium alloys, EFR has
a list of properties determined by alloy specific properties:

• Hydrogen presence, both in solution and as a compound of lithium hydride


• Higher chemical activity of the aluminum alloy containing lithium.

Higher chemical activity of alloys with lithium results in forming of covering films
with complex compound (Table 2.4), different from films created on alloys not con-
taining lithium. Meanwhile, oxides contained in the molten define the flux mix for
refining. For example for alloys EFR polluted mainly with aluminum oxides, it’s
recommended to use sylvinite and cryolite base fluxes, and for alloys alloyed with
magnesium—carnallite based fluxes [33]. For aluminum–lithium alloys, the usage
of sodium-containing fluxes, such as sylvinite-based fluxes, is unacceptable. The
usage of pure carnallite is also undesirable due to the reaction of magnesium chlo-
ride with lithium according to the following formula:

MgCl2 + 2Li = 2LiCl + Mg. (3.6)

At the same time, magnesium chloride actively removes magnesium oxide from
the metal alloy creating oxychloride of magnesium according to the following
formula:

MgCl2 + 5MgO = MgCl2·2MgO (3.7)


50 Aluminum–Lithium Alloys

To purify aluminum–lithium alloys from such inclusions as LiAlO2, Li2O, and


Li(OH)2, it is reasonable to use lithium chloride as part of the flux for EFR.
Introduction in fluxes trifluoride of aluminum for aluminum–lithium alloys allows
lowering the content of hydrogen and sodium in the alloy.
Based on theoretical and experimental researches for alloy EFR of Al–Mg–Li и
Al–Mg–Cu–Li system, the flux of the following content was chosen: 25%–35% LiCl,
15%–20% KCl, 30%–50% of carnallite, 4%–10% AlF3; and for alloys of system Al–
Cu–Li the flux with following content: –50% LiCl, 40%–50% KCL, 4%–10% AlF3
[97,98].
When connecting traditional technology of EFR and refining of the alloy with
flux in a melting tray, we received a substantial increase of refining effectiveness
[98,99]. The metal in the casting trough, connected through a liquid metal current
with metal in the refining chamber with an electrode, is under an electric load.
As a result of this, interphase tension on the border of metal-flux decreases and
good conditions for release of oxide inclusions and hydrogen to the interphase
border and their transformation to flux is created. The adhesion work of metal
infractions into the flux environment Wa M-I is also lowered, which leads to speed
gain of inclusion transition into flux. Thus, the electrical current, flowing through
the border metal-flux, is an additional moving power in the process of infractions
delivery from the molten into the flux. The number of inclusions, transferred from
molten into flux during the unit of time, is proportional to the area of phases divi-
sion. In the traditional approach of EFR, the area of phases division equals to the
area of the current of molten while transported through the flux and amounts to
10–40 cm 2. During EFR usage on the way of tap-hole-casting mold in billet cast-
ing technology from aluminum–lithium alloys with alloy protection with flux,
area of the molten in the casting trough is added to this area, which amounts to
200–800 cm 2. Accordingly, the effectiveness of electroflux treatment of the alloy
also increases.
Data regarding electroflux refining influence on the hydrogen content are pre-
sented in Table 3.3.

TABLE 3.3
Average Content of Hydrogen in Aluminum–Lithium Alloy Billets
Without EFR EFR Treatment
Hydrogen Content Number Hydrogen Content Number Average Level of
Alloy Cm3/100 g Me of Billets cm3/100 g Me of Billets Degassing (%)
1420 0.55 304 0.38 238 31
1421 0.47 23 0.40 8 15
1441 0.36 12 0.21 24 42
1450 0.43 46 0.37 62 14
1460 0.45 20 0.35 52 22
Hydrogen in Aluminum–Lithium Alloys 51

3.2.5 Nature of Hydrogen Distribution in 1420 Alloy


Billets and Semi-Finished Products
The nature of hydrogen distribution along the section of 1420 alloys billets was ana-
lyzed; for that purpose, we sampled templates from three flat ingots with 225 ×
950 mm section and three round billets with the diameter 450 mm [85]. The sample
cutting out scheme is presented in Figure 3.7.
Along the section of a flat ingot, there’s a certain regularity of hydrogen distri-
bution, which is represented by approximate equality of extreme points to one and
the same sample numbers (Figure 3.7). The location of maximum bears probable
nature and for all ingots and billets corresponds to the zone between samples num-
ber 6 and 9.
The fact that the range of values increases when the average value of hydrogen
content in billets increases also draws the attention.
The location of the extreme points is also connected with the scheme of distribu-
tion of the liquid alloy coming into the cast mould due to the geometrical conformity
of the main extremes with the size of distribution nozzle in the round section and
with the form of inlet branch in the plane section is observed. The scheme of the
metal inlet is connected with crystallization speed distribution in the dimple. The
specific features of lithium alloys also influence on the process. In these alloys, there
is a more difference between the equilibrium content of hydrogen in liquid and solid
states than in traditional aluminum alloys. Intensive boiling of the melt in the pool
during crystallization connected with that, it can be visible during pouring without
protection of bath level by flux. The more initial content of hydrogen in liquid alloy
the more intensive the boiling is. The possibility of capture of the outgoing hydrogen
is higher in this case in those areas where the molten currents come into the dimple.
In billets with low and average hydrogen content, it is distributed more even. In a flat
ingot, the maximum hydrogen content corresponds to the point located at a distance
of 1/16 of the width and 1/8 of the thickness from the center of crosscut template
(Figures 3.8 and 3.9).

1
2
3
4
5
24 23 22 21 20 19 18 17 16 15 14 13 12 11 10 9 8 7 6

1 2 3 4 5 6 7 0

(a) (b)

FIGURE 3.7  The sample cutting out scheme from (a) flat ingots and (b) round billets.
52 Aluminum–Lithium Alloys

2.1
2.0
1.9
1.8
1.7
1.6
1.5
1.4
Hydrogen content, cm3/100 g

1.3
1.2
170–5
1.1
1.0
0.9 170–4
0.8
0.7
0.6
169–6
0.5
0.4
315–1–2
0.3 315–1–3
0.2
0.1

24 23 22 21 20 19 18 17 16 15 14 13 12 11 10 9 8 7 6 5
Sample number

FIGURE 3.8  Distribution of hydrogen along the section of 1420 alloy ingot.

In the round billet, there’s the lowest hydrogen content and it corresponds to the
central area, the maximum is observed in the area between samples 3 and 5, which
also as in the flat billet corresponds with the area of current outgoing from the dis-
tribution nozzle (Figure 3.10).
The study of hydrogen distribution along the length of the round billet showed
that its maximum content is concentrated in the area along the butt-end (Figure 3.11).
In the beginning of the casting process, to what butt-end corresponds, the molten
surface in the cast mold is not protected from oxidation, metal comes in the form of
falling currents, and an intensive mixing of the molten, active interaction with the
atmosphere happen. That’s why the butt part of billet always has a higher contami-
nation with oxides and hydrogen. After that, the casting process is stabilized, and
the hydrogen content in the billet decreases and doesn’t change practically along the
length of billet (Figure 3.11).
As consistent with the data received on the segregation of hydrogen along the
section of ingots, we defined areas of sampling to conduct the gas analysis—areas
with maximum hydrogen content. In flat ingots, this area is the point of section,
located at a distance of 1/16 of the width and 1/8 of the thickness from the center
Hydrogen in Aluminum–Lithium Alloys 53

1.3
1.2
170–5
1.1
1.0

Hydrogen content, cm3/100 g


0.9 170–4

0.8
0.7
0.6
169–6
0.5
0.4 315–1–3
0.3 315–1–2

0.2
0.1

1 2 3 4 5
Sample number

FIGURE 3.9  Hydrogen distribution across the flat section of 1420 alloy ingot.

1.5
1.4
1.3
1.2
1.1
Hydrogen content, cm3/100 g

1.0
0.9
0.8
0.7
0.6
0.5 185
0.4 167

0.3 167

0.2
0.1

1 2 3 4 5 6 7 0
Sample number

FIGURE 3.10  Hydrogen distribution along the section of a round billet.


54 Aluminum–Lithium Alloys

Hydrogen content, cm3/100 g 0.6

0.5
168–3

168–6
0.4 167–3
168–5
167–2
0.3

1 2 2 3 3 4 4 5 5
Butt end Head end
Template and blank number

FIGURE 3.11  Hydrogen distribution along the length of billet.

of template. In round billet, it’s a template that is adjacent to the butt end. The
sample center taken should be the half of the radius provided the sample location
is along the radius.
On melt series, the parallel definition of hydrogen content in flat ingots was car-
ried out, and also die forgings obtained from them were analyzed. Sampling from

TABLE 3.4
Content of Hydrogen in Cast-Billets
and Die Forgings from Alloy 1420
Hydrogen Content
(cm3/100 g)
Number Billet Forging
1 0.4 0.48
2 0.4 0.38
3 0.42 0.58
4 0.45 0.30
5 0.40 0.43
6 0.40 0.38
7 0.40 0.30
8 0.41 0.30
9 0.50 0.43
10 0.40 0.32
11 0.40 0.27
12 0.40 0.40
13 0.40 0.40
Hydrogen in Aluminum–Lithium Alloys 55

ingots was done before sampling scheme correction. Samples from forgings were cut
out randomly from different sections, in which structural researches and weld grad-
ing were carried out in parallel.
In Table 3.4 are shown results of hydrogen findings in cast billets and forgings
obtained from them. As a general rule, the content of hydrogen as well as in com-
mercial alloys in forgings is lower than in billets by 0.05–0.1 cm3/100 g. However,
in some forgings higher values of hydrogen content were obtained, which is related
to wrongly chosen area of sampling from billet, as shown earlier. The correction of
sampling area done eliminated these fallouts.

3.3 ABOUT FORMS OF HYDROGEN PRESENCE


IN LITHIUM-ALUMINUM ALLOYS
A comparison of integral curves characterizing hydrogen content in 1420 alloy in
the beginning and at the end of melt provided with 16 h settlement under flux shows
notable decrease of its level (Figure 3.12). As it can be assumed from the hydrogen
sediment—initial hydrogen content curve built on obtained data, it can be calcu-
lated that it can reach 30%–60%, which is substantially higher than in commercial
aluminum alloys (Figure 3.3). The molten tends to reach the concentration equal to
0.4–0.6 cm3/100 g, which is probably equilibrium for these conditions of melting in
the flame machine with flux protection of bath level for 1420 alloy.
At 24 h sediment of 1420, alloyed with beryllium, in a reverberatory electric hold-
ing furnace, the hydrogen content didn’t practically change (Figure 3.12). During
sediment of the same alloy in the same holding furnace under a layer of flux in a
form of eutectic mixture of lithium and potassium chlorides, the content of hydrogen
decreased by almost three times.

I II
Hydrogen content, cm3/100 g

2.8

2.4

1.6

1.2
1 2 3 4
Melt numbers

FIGURE 3.12  Change of hydrogen content in 1420, alloyed with beryllium, during settle-
ment in a reverberatory electric holding furnace without protection of the bath level: I, initial
contents; II, after holding within 24 h.
56 Aluminum–Lithium Alloys

Substantially higher effect of hydrogen decrease in alloy 1420 as compared with


commercial aluminum alloys, possibly related to hydrogen presence in the molten
in a form of lithium hydride LiH. Finding of lithium hydride in aluminum alloy via
phase x-ray structure analysis is obstructed by the fact that it has structural and size
correspondence with aluminum. To conduct the analysis, it’s required to have higher
concentration of lithium-hydride in a sample.
On a series of commercial melting in flame heating furnace in the process of
settlement during pouring on the bath-depth samples were taken from the molten
and the slag. As we could see from Table 2.6 in slag, taken from 1420 alloy surface
after 6–15 h after lithium introduction, lithium hydride was found using phase x-ray
structure analysis.
In samples, taken from molten and cast with high speed of crystallization on a
brass polished plate, lithium hydride was found only in a sample taken from the
top layers of molten after 16 h of sediment. In our opinion during that time due to
hydrides floating, its concentration in surface layers was higher, which allowed to
separate lithium hydride from solid solution of aluminum during x-ray recording
in monochromated radiation with low speed countermovement on lines with high
indices (331) and (420) (Figure 3.13) [85].
As an indirect proof of hydrogen presence in a form of lithium hydride in lithium
alloys, can be viewed through the change in flowability depending on degassing.
Comparing of this characteristic for 1230 alloy, melt in a gas-fired furnace without
any consequent degassing, with metal undergone vacuum treatment, showed that
close to 670°C temperature, in other words temperature corresponding to hydride
transfer from liquid into solid state, the metal which didn’t undergo degassing quickly
becomes less flowable, than the metal undergone vacuum degassing (Figure 3.14).
Research works [87,88] defined that soluted hydrogen doesn’t affect flowability
in liquid state, and that’s why lower content of hydrogen by itself doesn’t change
that characteristic. At the same time, it’s known that vacuuming of steels changes

AI(420) AI(331)

LiH(331)
LiH(420)

,0

58 57 56 55

FIGURE 3.13  Part of a diffractogram, showing the presence of lithium hydride in a sample,
taken from a surface layer of a commercial alloy melt.
Hydrogen in Aluminum–Lithium Alloys 57

250

200

Rod length, mm
150

2
100

1
50
690 680 670 660 650
Temperature, °C

FIGURE 3.14  Flowability of 1230 alloy, melting in a gas forge: 1—after introduction of
lithium, 2—after introduction of lithium and following vacuuming (increase = 5 mmHg).

condition of nonmetallic particles and increases flowability. The same authors didn’t
find any effect of vacuuming on flowability of pure aluminum and aluminum alloys
with zinc. It can be assumed that in our case solid inclusions reducing the fluidity
are solid hydrides of lithium. Vacuuming helps its decomposition and, as a result,
increases flowability of the molten.
A defined sharp increase of vacuum-degassing effect can serve as a confirma-
tion of hydrogen presence in the form of lithium hydrides when the temperature of
metal increased from 720°C to 760°C, which corresponds to the intensive increase of
partial pressure of hydrogen over hydride [55,61,91]. It is known that for commercial
aluminum alloys, on the contrary, the highest vacuuming effect can be reached at
temperatures near to nonequilibrium solidus [90], which is related to lowering of sol-
ubility of hydrogen in aluminum. In one of the research works [86], it was found out
that solubility of hydrogen in aluminum containing lithium dirt nonlinearly depends
on the pressure of molecular hydrogen over the molten, which allowed the author
of the research to express a supposition about banding of hydrogen part in hydride.
Thus, based on the data received we can state that in aluminum alloys with lith-
ium, one of the forms of hydrogen existence in a molten is lithium hydride.
Most possibly, we observe creation of hydride at two stages of technological process.
Firstly, during heating of lump scraps in a furnace, when hard charging mate-
rials cannot be covered by flux. At heating temperatures up to 500°C–600°C in
atmosphere of fuel combustion products, where water vapors are present, lithium
hydride is very stable and the process of its creation is thermodynamically favorable.
The fact of lithium hydride creation when heating samples of binary alloy Al–3%
Li in atmosphere, containing water vapors, at 550°C was confirmed by electronic
diffraction method. At temperatures for alloy preparation, lithium hydride is very
stable (dissociation pressure at those temperatures amounts to 30 mmHg) [55], and
isobaric-thermal potential of hydride creation at 700°C equals to 18.1 kJ/molar unit.
58 Aluminum–Lithium Alloys

Secondly, lithium hydride can be formed in a liquid pool during crystallization of


billets, when liquid phase is enriched with hydrogen discharged during crystalliza-
tion by means of solubility decrease while transforming from liquid to solid state,
and stability of the hydride of lithium increases—dissociation pressure of lithium
hydride at temperatures of the metal in the dimple within the range 630°C–520°C
decreases to 0.2 mmHg. Hydrides created during crystallizing stay in the cast-ingot
and come into the liquid alloy with scrap.
Different nature of hydrogen content change during storage of aluminum–lith-
ium alloy under flux and in the process of vacuuming with the following storage in
the argon environment, as shown earlier, related to physical-chemical properties of
lithium hydride.
Lithium hydride is a compound of ion type, in which lithium ion is positively
charged (Li+) and hydrogen ion is negative (H–).
It is known that ion link doesn’t have directional property, which is why the
interaction between ions is performed directionally independent. The system of
two charges equal by their absolute value, but opposite by their charge sign, creates
an electric field in the surrounding environment. That is why two ions of opposite
charge pulled together to each other keep the ability to interact electrostatically with
other ions. Therefore, a various number of ions of opposite charge can connect to
this ion. It specifies the inclination of ion molecules to associate with each other with
creation of unified gigantic molecules. At high temperatures, kinetic energy of mol-
ecules movement prevails over the energy of their mutual magnetism, which is why
in the gas state ion entities exist mainly in the form of nonassociated molecules. At
temperature lowering when transitioning in the liquid state, and especially in solid
state, the ion association shows itself at the most. Consequently, lithium hydride
being in liquid state can unify in “large molecules.”
Having substantially lower density than molten (0.77 g/cm3 for hydride vs.
2.13 g/cm3 for molten 1420), the hydride will float on the surface. In case of melt
preparation under flux protection, consisting of lithium and potassium chlorides,
this will be the dividing surface of metalflux. At that, removal of hydrides from the
molten is eased by the fact that viscosity of molten 1420 is very low, four times lower
than in AMG6 alloy. Besides that, lithium hydride is soluble in the flux, which is in
the form of eutectic mixture of lithium and potassium chlorides, due to the exchange
reaction with potassium chloride.

LiHliq + KClliq → LiClж + К (gas) ↑ + 1/2H2 (gas) ↑ (3.8)

It is known that the direction of spontaneous chemical reaction is defined by the


action of two factors: tendency to transformation of the system in the condition with
lesser inside energy and tendency to achieve the most possible condition. The latest
shows itself the more intensive the higher temperature.
In this case, the reaction takes place in the surfacing layers of molten during melt-
ing in a reflecting flame furnace in overheated condition, worsened by heat outlet as
a result of creation reactions of complex oxides, at temperatures 760°C–780°C.
Let’s evaluate the possibility of reaction with the help of table values [93,94] of
thermodynamical properties of substances taking part in the reaction.
Hydrogen in Aluminum–Lithium Alloys 59

Using additive properties of Gibbs energy, the calculation of the process can be
done by applying method of combining, representing this reaction in the form of
three reactions:

1. Li + 1/2Cl2 = LiC1
2. KCl = К + 1/2C12
3. LiH = Li + 1/2H2

∑: KCl + LiH = LiC1 + К + 1/2H2 (3.9)

For our case,

DGT (0 ) = DGT (01) + DGT (02 ) + DGT (03) (3.10)


å

In conjunction with data [95], changing of Gibbs energy for reactions (3.1) and (3.2)
at 1055 К equals

DG10055 (1) = -81,500 cal/molar unit; DG10055 ( 2 ) = +81,400 cal/molar unit (3.11)

According to data [27,33], nonconforming free Gibbs energy, enthalpy and entropy
of hydride creation equals

DH 0298 = 2166 cal/molar unit; DS0298 = cal/molar unit × degree; (3.12)


DG 0298 = 16,780 cal/molar unit

If at first close up to drop the changes of heat capacity with increasing temperature,
in other words ∆CP = f(T) = 0, then

DGT = DH 0298 - TDS298


0
(3.13)

Then for LiH at 1055 К,

DG10055 = -21,666 + 1,055 · 16.4 = -5,266 cal/molar unit (3.14)

For reaction (3.3), going the opposite direction, we take this value with the opposite
sign. Then,

DG10055( å ) = -81,500 + 81,400 + 5,266 = +5,166 cal/molar unit (3.15)


At temperature of reaction within 760°C–780°C product of potassium reaction


undergoes phase change and as a result of reaction creates a gas-state substance.
60 Aluminum–Lithium Alloys

It is known that during creation as a result of reaction of gas-state substance ∆S


equals + 30 ± 10 cal/molar unit · К. For rough calculation, we can use the correlation

DG (0T ) » DH 0298 - TDn G 30, (3.16)


where
∆nг is the number of gas molar units

which is created additionally for each transformation at the time of reaction [96].
In our case, one molar unit of potassium is created, therefore from the amount of
0
heat DG1055 we need to subtract T·30, in other words

1,055 · 30 cal/mole * K = 31,650 cal/molar unit (3.17)

Then, the heat effect of the reaction of interaction of potassium chloride with lithium
hydride will be

5,166 cal/mole * K – 31,650 cal/mole * K = – 26,484 cal/molar unit (3.18)

In other words, the reaction is thermodynamically possible and goes actively.


During 1420 alloy storage in the argon environment after the first vacuuming, we
can observe the process of second-step enrichment of alloy with hydrogen. After
repeated vacuuming, hydrogen content in the molten practically didn’t change. In
our point of view, this can be attributed with hydrogen presence in the form of
lithium hydride and be represented as follows: on the molten surface in the hold-
ing furnace a film is created, in which the prevailing phases are oxides of lithium
(Li2O), magnesium (MgO) and aluminum (χ-Al2O3), those substances are in the
form of combination of covalent form of binding with high dipolar moment. The
hydride in the form of “large molecules” floating to the surface during storage con-
sists of unions of ion type, which is drawn to oxide molecules and creates ion-dipole
complexes. Electrostatic ion–dipole interaction forces created help to decompose
the hydride into ions, freeing of hydrogen and transition of it into molten, which
leads to secondary enrichment of the molten by hydrogen. Repeated short-term
vacuuming is a very ineffective operation.
To eliminate the conditions of repeated enrichment of alloy with hydrogen after
vacuuming treatment, we need to eliminate conditions of oxides creation on the sur-
face of liquid alloy in the holding furnace by means of better sealing of the holding
furnace, increased effectiveness of argon drying. This will lessen the possibility of
ion–dipole complexes creation, which is the main factor of stabilizing the molten-
hydrogen system, during maturation of molten in the argon environment.
Increase in the effectiveness of vacuuming can be reached by increasing the tem-
perature of molten up to 750°C–760°C during vacuuming to provide conditions for
full hydrides decomposition.
An effective way to lower hydrogen content can be electric-flux refining through
flux containing potassium chloride. When calculating the temperature dependence
of linear settlement of crystallization of twin aluminum alloys with lithium and alloy
Hydrogen in Aluminum–Lithium Alloys 61

1230, substantial increase of before-settlement enlargement was registered in case of


gas enrichment of the alloy [44].
For commercial aluminum alloys, this is attributed to step-like lowering of hydro-
gen solubility during crystallizing [80]. During this process in commercial alumi-
num alloys, bubbles of molecular hydrogen are created, and part of them are floating,
and removed from molten part, and are emitted in between grains spaces, creating
gas porosity, or are captured by growing branches of dendrites with creation of gas
porosity [88]. In 1230, 1450 alloy billets of aluminum–copper–lithium system and
in alloy 1440 of aluminum–copper–lithium system, porosity is created when con-
centration of hydrogen in cast-ingot is not more than 1 cm3/100 g, what is attributed
probably with different values of maximal allowed concentration and specifics of
crystallizing. In AMG6 alloy, the porosity in cast-billets shows already when content
reaches 0.5–0.6 cm3/100 g.
Release of hydrogen during lithium-containing alloys crystallization goes on very
actively. Observing the surface of the molten in the casting mold when casting billets
from 1420 alloy beryllium showed that hydrogen evaporated breaks the film and lets
out metal in the form of fountains, which tells us about the high kinetic energy of gas
bubbles. This is most likely due to very big difference between solubility of hydrogen
in molten or solid alloy, and which defines the observed intensive hydrogen evapora-
tion during crystallizing in the form of constant boiling. Part of hydrogen has time to
evaporate from dimple, and part is tied into lithium hydride and stays in solid state
metal along with dissolved hydrogen. With a very high initial gas enrichment, not
all molecular hydrogen has time to evaporate and, as we could see, leads to creation
of porosity in billets.
In the process of many years of researches, done in experimental and commercial
conditions, the following have been developed and implemented into commercial
production:

• Technology of aluminum–lithium preparation with usage of protective


fluxes of set compositions used during 20 years for commercial production
of alloys 1230 and 1420.
• Special melting equipment was created such as melting machine as part of
coreless induction furnace and vacuum holding furnace.
• A technological process has been developed, including preparation of
molten in a furnace under flux protection made out of eutectic mixture
of lithium and potassium chlorides, blow-off of the furnace with chlorine
(chloride), vacuum treatment of the molten and its maturation in the envi-
ronment of dry argon during the process of melt pouring.

Technical process was implemented into commercial production and helps to receive
regularly lithium-aluminum alloys (1230, 1420, 1440, 1441, 1450, 1460, 1461, 1421,
1424, V-1469) with set chemical composition in main alloying elements and regu-
lated content of metal impurities and hydrogen.

You might also like