Sunteți pe pagina 1din 18

Oxygen Pyrometallurgy at Copper Cliff

A Half Century of Progress


Paul E. Queneau and Samuel W. Marcuson

CONTENTS
INTRODUCTION
OXYGEN PYROMETALLURGY
o THE DEVELOPMENT OF OXYGEN FLASH SMELTING
o OXYGEN ENRICHMENT
o OXYGEN TOP-BLOWN ROTARY CONVERTER
o OXYGEN SMELTING
o OXYGEN FLASH-CONVERTING CHALCOCITE CONCENTRATES
o OXYGEN TOP BLOWING/NITROGEN BOTTOM STIRRING
o OXYGEN CONVERTING CHALCOCITE CONCENTRATE
OXYGEN PRODUCTION
MEASURING PROCESS PARAMETERS
WINDOW ON THE FUTURE
REFERENCES

INTRODUCTION
"All nonferrous metallurgy will be benefited by the use of cheap oxygen .
. .
the application of oxygen will revolutionize the art of smelting and it
will
probably change the whole operation and equipment."
So wrote F.W. Davis of the U.S. Bureau of Mines almost 75 years ago.
1
He cautioned that
the existing cost of oxygen was much too high for metallurgical processes, but he
forthrightly stated that "the oxygen industry is now able to make plants for supplying large
quantities of oxygen to metallurgical industries at low cost." His committee's convincing
Report of Investigationsprimarily concerned with steelmakingwas aimed at major
decreases in metal production cost. There was no consideration of environmental impact
this subject was not considered important in 1923, and the opinion would not change for
decades. As pioneer environmentalist, scientist Rachel Carson warned, "A grim specter has
crept upon us almost unnoticed."
2
Today, engineers must heed the bell that she first tolled.
3
,4

Davis' prescient words fell on deaf ears throughout the world's nonferrous industry.
Ultimately, however, they sounded in Inco's Copper Cliff Research Laboratory in 1941 and
were amplified by the pertinent 1936 study of Telfer Norman
5
but there was a war to win.
When the veterans returned, the "innocently beloved proud plumes of heavy industry" that
were billowing out of the Copper Cliff Smelter stacks were recognized for what they were.
Most notably, their sulfur content was wounding Ontario and Quebec forest lands.
Simultaneously, the paper-making sulfite pulp mills were importing large quantities of
elemental sulfur, and the smelter was importing large quantities of coal.
Economic application of oxygen pyrometallurgy would permit replacement of this sulfur by
liquid sulfur dioxide, with simultaneous replacement of the coal by low-cost tonnage
oxygen, produced using low-cost hydroelectric power. Reaching this grail would achieve
not only a significant decrease in metal production cost, but it would also enable an
significant decrease in environmental degradation, a goal that would be increasingly
stressed by government.
Inco's resulting half century of progress in developing oxygen pyrometallurgy (Figure 1) is
a paradigm of the long-term teamwork that is necessary to attain such an operational shift
laboratory theoreticians, hot-metal operators, and management all working together with
mutual respect.









Figure 1a. Flowsheets of nickel-copper extraction at Copper Cliff1945.










Figure 1b. Flowsheets of nickel-copper extraction at Copper Cliff1995.


OXYGEN PYROMETALLURGY
Intensive laboratory and tonnage pilot-plant R&D on the oxygen flash smelting of sulfide
concentrates was initiated at the end of December 1945. It proceeded in classical stages:
first in two-dimensional paper studies of smelting, using solely sulfur and iron as fuels;
then in increasingly large three-dimensional apparatus design and operation. Excellent
validating metallurgical and economic databases were established by the end of June 1947.
Unfortunately, the apparent high cost of tonnage oxygen proved forbidding: the price
stipulated by the dominant supplier, Linde (USA), was prohibitive.
Assistance was, therefore, sought from sound, though relatively inexperienced, sources, and
the low bidder was accepted. It was the parent of Canadian Liquid AirAir Liquide of
Paris, which was then a builder of small oxygen plants; it is now the builder of the largest
oxygen plants in the world. The fearful cost barrier was hurdled by the determining
contributions of their brilliant chief engineer: Maurice Gobert. A worthy bid was also
submitted by enterprising Leonard Pool, founder of Air Products, the supplier of reliable
mobile dwarf oxygen generators for U.S. Army Engineers in World War II; Air Products is
now a giant tonnage oxygen supplier.
In January 1948, Copper Cliff managers Roy Gordon (plant) and Paul Queneau (R&D)
submitted their joint decision to top management:
"In view of the economic superiority and metallurgical potentialities of
the
new process it is recommended that a (first step) 400 t.p.d. copper
concentrate
flash smelting unit and a 300 t.p.d. oxygen plant be installed at Copper
Cliff.
. . . There would be periods of large surplus oxygen production . . .
such
oxygen could be consumed to advantage in the Smelter reverberatory
furnaces for
air enrichment so as to decrease coal consumption."
In those early post-war years, there were very long delays in equipment delivery,
particularly of special machinery for the oxygen plant. A 500 t/d copper concentrate oxygen
flash smelting furnacesupplied with low-cost oxygen by a 300 t/d plantfinally went
on-stream four years later: January 2, 1952.
6

This energy efficient and environmentally friendly reactor led the world of
pyrometallurgyincluding the oxygen steel converterin its direct, massive-scale, use of
oxygen. In the spring of 1948, convincing experiments in Gerlafingen, Switzerland,
inspired by Robert Durrer, demonstrated oxygen top-blowing of blast-furnace hot metal
into steel. In November 1952, VOEST's oxygen steelmaking Linz/Donawitz converter plant
went on-stream in Austria.
7

The Development of Oxygen Flash Smelting
The first autogenous oxygen flash-smelting laboratory experiments were conducted in a
small horizontal, sheet metal-enclosed, refractory-lined furnace. Dry sulfide concentrates
and flux were injected into the preheated unit using 99.5% O
2
cylinder oxygen. After
months of trial and error, batch runs reached rates of up to 3 t/d, with matte grades and
metal recoveries at least equal to conventional copper and nickel reverberatory practice,
and off-gas analyzing up to 98% SO
2
. Calculations based on the test data indicated that the
cost of oxygen consumed per tonne of copper concentrate would be half that of
reverberatory furnace coal consumptionproviding the cost of oxygen was low! The total
of other costs appeared at least competitive.
A pilot-plant furnace was then designed and built for the continuous smelting of dry sulfide
flotation concentrates. A design principle was "keep it simple stupid," so both furnace and
burner were basically adaptations of the neighboring reverberatory furnaceswith mineral
concentrate and oxygen substituted for coal and air. Another difference from conventional
practice was cyclic reduction drenching of furnace slag with molten FeS from flash-smelted
pyrrhotite concentrate prior to tapping, which decreased slag oxygen potential and CuNi
loss. This practice would have been improved by the inclusion of some coal with the
pyrrhotite, but this "obvious" action was not obvious at the time. Oxygen was supplied at
95% O
2
by a Canadian Liquid Air 5 t/d oxygen plant, and furnace off-gas was treated in a
Canadian Industries Limited sulfur dioxide liquefaction pilot plant.
Pilot-plant operations started in January 1947 and were terminated late in 1948 with full
success at 25 t/d of concentrate. Furnace dimensions had to be increased four times
mainly due to refractory erosion by flying molten particulatesbefore the final size was
attained. Matte grades of up to 75% CuNi were produced, with pyrrhotite-cleaned slags
analyzing up to 0.9% CuNi. Testing of the liquid SO
2
product for sulfite pulp production by
Abitibi Paper Company indicated its superiority to SO
2
produced by sulfur burning.
The day after New Year's Day, 1952, a pioneering commercial oxygen reactor (Figure 2)
flashed into life. In accordance with prior planning, the new plant superintendentplant
coinventor Charles Youngwas the former pilot-plant superintendent; before that, he had
been a laboratory engineer involved in the initial R&D. The able, invaluably supportive,
up-from-the-ranks smelter manager, Duncan Finlayson, had originally been contemptuous
of the "black box" furnace concept: "I can't stick my head in itI don't like it!" His
understandable skepticism was overcome by early pilot-plant exposure and by the
effectiveness of the well-maintained instrumentation employed, which allowed the operator
to monitor and control the key furnace variables. After the usual birth pains, the daily
furnace charge averaged 500 tonnes of 28% Cu chalcopyrite concentrate, 100 tonnes of
pyrrhotite concentrate, and 90 tonnes of sand flux. The matte produced analyzed 45%
CuNi, the slag was 0.75% CuNi, and the off-gas was 75% SO
2
. The lastrelatively small
in volumewas condensed in a 300 t/d liquid SO
2
plant, which was a decimal order of
magnitude larger than any other in the world.
8







Figure 2. The original Inco Oxygen flash furnace (reproduced from the July 1995 Journal
of Metals).


Having well served its educational function, the innovative furnace was replaced by a
rewardingly profit-making 1,000 t/d unit two years later. The critically vital, Inco-owned
and operated Air Liquide tonnage oxygen plant delivered admirablyproducing 95% O
2

gas at a total cost of $4/t and using 0.4/kWh power. Low-cost production of pipeline
tonnage oxygen was proven! Pyrometallurgy was reborn! The Inco staff technical paper
describing this achievement,
6
and the five names in alphabetical order on the covering
patent
9
complied with the aphorism: "Share credit, share success." Now, Inco owns and
operates Air Liquide oxygen plants with a total capacity of 1,800 t/d to feed a variety of
Copper Cliff furnaces.
10

Oxygen Enrichment
L.S. Austin wrote in 1919:
11
"One can infer that the reverberatory furnace is primarily a
combustion chamber (for the waste heat boilers), with the melting, the furnace reactions
and the separation of matte secondary factors." The villain of his grievance was tonnage
nitrogen. Oxygen-enrichment of the furnace's combustion air increases fuel efficiency and
permits higher smelting rates. The amount of heat delivered to the waste heat boilers by
each tonne of nitrogen equals that required to smelt one tonne of solid charge. This was
demonstrated in Copper Cliff smelter full-scale tests on one of the seven 9.1 m X 33.5 m
nickel reverberatory furnaces. These showed that one tonne of oxygen was equivalent to at
least one half tonne of coal and that throughput rate could be increased 33% by a decrease
in the nitrogen/oxygen ratio. Oxygen-enrichment of these furnaces' combustion air was,
therefore, established as routine practice.
There were 19 air-blown 4.0 m X 10.7 m Peirce-Smith converters in the Copper Cliff main
converter aisle, treating liquid reverberatory furnace matte. Oxygen utilization efficiency,
in its exothermic reaction with iron and sulfur, was close to 100%. However, half of the
total heat developed suffered the nitrogen curse, so it was lost. Several of these converters
blew flash-furnace matte to blister copper. If blower air was enriched with oxygen (e.g., to
33% O
2
, thereby changing N
2
/O
2
volume ratios from 4:1 to 2:1), the usual heat balance in
the converter would be much improved. In addition to the increase in the conversion rate
and decrease in gas volume, cold charge could be smelted. The limiting factor would be
excessive impact on tuyere and refractory life. Converter trial operations, in which the
oxygen content of blower air was varied in the 25-35% range, were launched in 1958.
These indicated a 30% O
2
content to be optimal. The additional useful reaction heat was
employed to melt large quantities of scrap and concentrates (e.g., the 73% copper filter
cake produced by the matte flotation copper-nickel separation plant). The 30% O
2
level of
enrichment was then systematically extended to all 19 converters.
On the basis of Inco's experience with tonnage oxygen generation and utilization, one of the
authors was able to write the following in 1960:
"The pyrometallurgist will gain further benefits from the advent of low-
cost
oxygen. The dead hands of nitrogen have been lifted from oxidation
reactions
which utilize the oxygen in air. The nonferrous metal industry is on the
threshold of understanding in this connection. As one example, oxygen
enrichment of combustion air will give new life to otherwise obsolescent
or
obsolete conventional furnaces. Greatly improved reverberatory and rotary
furnace design will be employed for utilization of tonnage oxygen in
continuous
autogenous smelting and converting. The tuyereless, top-blown oxygen
steel
converter will invade and conquer the smelters and refineries of the
nonferrous
industry. Decrease in nitrogen dilution of sulfurous smelter gases will
permit
increased sulfur fixation and result in decreased atmospheric
pollution."
12

This prediction has proven accurate.
13-17
Nevertheless, low-cost oxygenthe immense
value of which was demonstrated on a commercial scale in 1952continued to be greatly
underemployed for decades.
18,19

Oxygen Top-Blown Rotary Converter
Inco took another major stride forward in the use of tonnage oxygen to enhance heat and
mass transfer in nonferrous pyrometallurgy, by pioneering the employment of a
steelmaking converter for this purpose. In the metallurgical world, it was the generally held
opinion that blowing nickel sulfide to metal in a converter presents impossible-to-solve
thermodynamic and operating problems. In fact, it is impossible to produce metallic nickel
from nickel matte in a Peirce-Smith converter (e.g., due to disastrous nickel oxide
formation). During sulfur dioxide evolution, nickel and nickel sulfide form a single solution
phase extending from Ni
3
S
2
to pure nickel, and the NiO (melting point: 1,984 C) that
forms has low solubility in matte.
However, 1941 Copper Cliff Laboratory studies had indicated that conversion was possible,
given sufficiently high bath temperature and oxygen potential. Hence, the high-
temperature, broad-range oxygen potential and excellent mixing capabilities of the
turbulent bath, characteristic of a post-war top-blown rotary converter (TBRC) steelmaking
process, appeared extremely attractive. The vessel employed provided efficient and
effective gas-liquid-solid contact throughout the bathwith concomitant extraordinarily
extensive control of temperature and oxygen potential. It enhanced heat transfer, increased
the overall rate of the chemical reactions, minimized composition gradients within each
phase, and significantly reduced diffusion barriers.
Oxygen metalmaking by tonnage nickel matte experiments in a TBRC were proposed and
opposed in heated debates within Inco. Conventional wisdom said such experiments would
fail and perhaps kill: the converter would produce nickel oxide instead of metal, and the
nickel sulfide (melting point 788C), at 1,650C, would cut through the rapidly rotating
refractory lining like a knife through butter.
In 1958, Paul Queneau and John Feick, Copper Cliff Peirce-Smith converter
superintendent, supported by John Thompson, Inco's chemical engineer chief executive
officer, explored direct nickel sulfide conversion to oxygen crude nickel in a three tonne
KALling converter at DOmnaverts Steel Works (KALDO) in Sweden. The experiment was
immediately successful. This victory in novel nickel making having been achieved,
opportunities in TBRC oxygen coppermaking, fire refining, and beyond were revealed and
successfully pursued by Inco in a seven tonne TBRC at Port Colborne.
20,21
It all seemed so
obvious after the breakthrough.
In 1971, two 50 tonne TBRCs were commissioned at Copper Cliff (Figure 3) as the first
stage in the transformation of complex metal sulfide intermediates to 99.98% pure nickel
by the Inco Pressure Carbonyl Process. Today, the operation of these converters is routine,
having produced a million tonnes of oxygen crude nickel to date.
22





Figure 3. The oxygen top-blown rotary converter
in action at Inco's Copper Cliff nickel refinery.


Commissioning and operation of the TBRCs completed development of an oxygen culture
at Copper Cliff. Management and technical staff understood the advantages oxygen
technology offered, and operators and maintenance personnel knew how to work with
oxygen as a useful ally. This culture, coupled with continuing active research and
development, enabled commercialization of new oxygen technologies as Inco responded to
the changing economic and environmental challenges of the seventies and eighties.
Oxygen Smelting
Experimentation with roof-mounted oxy-fuel burners in reverberatory furnaces commenced
in October 1977
23
using ideas developed at the Caletones smelter.
24
The first burners
generated excessive noise levels and deteriorated rapidly. Two years of development
generated effective oxy-fuel smelting capabilities and yielded a rugged burner that gave a
stable flame at acceptable noise levels. In October 1979, a reverberatory furnace equipped
with 12 oxy-fuel burners began operation. Smelting rate increased by 45%; fossil fuel
consumption and exhaust gas volume decreased by 55% and 65%, respectively.
23,25
The
increased productivity and lower gas volumes contributed to a major rationalization of the
furnaces and flue systems and concomitant improvements in the workplace environment.
Oxy-fuel fired reverberatories operated for more than a decade, treating all of the nickel
concentrates. The last such furnace was shut down in 1993 with commissioning of two
oxygen flash furnaces as part of Inco's $600 million (Canadian) SO
2
abatement
program.
10,26, 27

These second-generation Inco oxygen flash furnaces (Figure 4) are larger than the original
furnaces, employ greater amounts of water cooling, and incorporate modern gas cleaning
systems that are extremely compact, as allowed by the low-volume exhaust gas of tonnage
oxygen smelting. Cleaned gas feeds a double-contact acid plant and the original liquid SO
2

plant.
10,26
Furnace feed is a bulk copper/nickel concentrate. Petroleum coke and natural gas
are added to provide supplemental heat and to allow return of converter slag and smelting
of reverted material. Table I compares the original and new furnaces.









Figure 4. A schematic of a current Inco oxygen flash furnace at Copper Cliff.





Oxygen Flash-Converting Chalcocite Concentrates
From 1965 to 1985, the smelter processed its -325 mesh nickel-containing chalcocite
flotation concentrate (Cu
2
S derived from the matte separation process) by Garr gun addition
to blowing Peirce Smith converters.
9,28
This procedure led to long converting cycles and
was a source of large dust emissions.
Development of a novel oxygen-based flash-converting process gave a short-term, low-
capital improvement.
28, 29
Due to the rapid kinetics of oxygen reactions, the smelter was
able to use a surplus Peirce-Smith converter shell as the vessel. The in-house development
of a suitable feed system and unique oxygen flash gun that could simultaneously fire
natural gas and filter cake completed the process. This process, the first commercial
application of flash copper converting, started in 1985 and operated for eight years, treating
8% moisture filter cake at rates of 250 t/d to 300 t/d. More than 300,000 tonnes of molten
semiblister assaying 2-3% sulfur were produced.
Oxygen Top Blowing/Nitrogen Bottom Stirring
Since 1993, Inco has commercialized several innovative techniques for oxygen converting
to blister copper, all based around the top blowing of oxygen accompanied by gentle
nitrogen bottom stirring.
30,31
In 1984, crucible experiments revealed that extraordinarily
high oxygen efficiencies could be obtained during blister finishing by blowing oxygen onto
the melt while sparging with nitrogen. Moreover, this mixing promotes desulfurization of
the molten blister and enhances the approach to chemical equilibrium. Exhaustive
laboratory tests demonstrated that the process was effective at low and high top-blowing
rates, was insensitive to lance position, and required only small flows of sparging gas.
Pilot-plant studies at the 3-5 tonne scale confirmed the results.
30
Importantly, this work
demonstrated the usefulness of ceramic porous plugs (Figure 5) for nitrogen injection into
copper and gave the confidence needed to install them into a commercial vessel.








Figure 5. A cross-section of a porous plug.


Full scale tests began in 1989 using a Peirce Smith converter shell equipped with two
porous plugs and an oxygen lance. The combined blowing approach yielded oxygen
efficiencies of 85% during blister finishing, although the subsonic open pipe lance was
mounted 1.8-3.7 m from the bath and blew gently to minimize splashing. By using oxygen,
the converter consumed scrap at a rate of 20% of the semiblister charge (2-3% sulfur). The
porous plugs performed well in copper service, and elimination of tuyeres minimized
fugitive emission generation when the converter rolled into and out of stack. Finishing
blister by oxygen top blowing/nitrogen bottom stirring was incorporated into the new
flowsheet of the Copper Cliff Smelter in 1990.
28
To conserve capital, the process was
implemented in existing Peirce-Smith converter shells (Figure 6). Commercial operation
began in November 1993. *








Figure 6. A schematic of a commercial oxygen top-blown/ nitrogen bottom-stirred
converter.


Oxygen Converting Chalcocite Concentrate
Continuing research into the flash converting of chalcocite showed that the 10-20% dusting
rate experienced at all scales of operation was due to particle fragmentation during
ignition.
33, 34
Moreover, removal of this large quantity of dust in a gas cleaning system
feeding an acid plant involved major handling problems.
35
Hence, the search for a better
way of oxygen converting chalcocite began.
Plant tests showed that tuyere injection of chalcocite accompanied by oxygen top
blowing/nitrogen stirring was effective, and commercial operation commenced in 1993.
36

The reactor vessel is a 18 m long, 4.5 m diameter cylinder with oxygen lances mounted on
each endwall. Each of two blow-tank conveying systems is connected to a single tuyere and
injects chalcocite at a rate of 25 t/h. Dusting rate is about 1%, and oxygen efficiency is in
the 90% range.
Pilot-plant studies in 1994 showed the feasibility of combining top blowing/bottom stirring
with a simplified feeding technique. Full-scale tests began in August 1995. The application
of nitrogen stirring through porous plugs has been extended a further step. Feeding is
accomplished by gravity introduction of dry, nonagglomerated concentrate (90% -44 m)
through a water-cooled pipe onto the "eye" created by the nitrogen.
37
Supplemental heat is
provided in the area to promote melting. Feeding zones and converting zones are separated
so that gas velocity around the feed stream is minimal (Figure 7). As a result, a dusting rate
of 1.5-2% is achieved. Full-scale development continues. As demonstrated in the pilot-plant
work, this simple approach can be useful for other continuous converting applications.









Figure 7. The gravity-fed oxygen reactor for chalcocite converting.


OXYGEN PRODUCTION
The continuing developments in oxygen pyrometallurgy have been assisted by major
improvements in tonnage oxygen production. In contrast to 1946, the industry of today is
highly competitive with several suppliers. Cryogenic oxygen production remains the
preferred technology for large tonnage applications. Developments in centrifugal
compressors, improvements in the fractionation cycle, and the application of computer
controls have greatly increased the energy efficiency and productivity of modern plants.
38-41

Molecular sieve front-end purification eliminates the cold box, enhances gas purities, and
obviates the yearly plant shutdown for deriming.
41
Typically, today's oxygen plants
incorporate one day's storage of liquid oxygen to ensure against plant shutdowns and are
equipped with computerized load following to minimize energy consumption and costs.
Capital and production costs for such plants are summarized in Figure 8.
42
Aside from
capital, energy remains by far the major factor in production costs; other supply, operating,
and maintenance labor costs are relatively small.









Figure 8. Oxygen production requirements: (a) plant investment for oxygen at 20 psig; (b)
power requirements for oxygen at 20 psig; (c) power requirements for oxygen at pressures
above 20 psig; and (d) operating costs (MIT-maintenance, insurance, and taxes).


The vacuum swing adsorption process has found application for oxygen requirements of
less than 100 t/d at purities of 90%. In this technique, nitrogen is removed in two molecular
sieve adsorption trains connected in parallel and operating in sequence. While incoming air
is purified in one train, the other train is regenerated by pressure reversal.
40
Such an
installation can be used to supply oxygen to a relatively small user or to top-up a large
cryogenic plant that cannot meet ever-increasing smelter demands.
MEASURING PROCESS PARAMETERS
As the oxygen pyrometallurgy revolution continues, the reaction rates and complexity of
the processes generally increase. Conventional methods of monitoring and controlling
pyrometallurgical processes cannot meet these challenges. Moreover, the closed nature of
modern reactors prevents the use of traditional techniques such as visually monitoring
flame color or bath appearance. Necessary sensors, signal translators, and data processors
must give correct and timely information to the operator. If betrayed, the operator is
automatically wrong, and the process can automatically failpossibly disastrously. All
concerned must understand the vulnerability of modern "invisible" processes to
uncontrolled, characteristically fast, and potentially dangerous reactions. This is especially
true for single-vessel, multistaged processes. The operator needs to know not only what
was going on in the closed vessel, but what is going on and what will be going on!
Incorporation of useful, well-maintained, reliable "blind flying" instruments is
indispensable as is precise metering of inputs and outputs. Effective blending of feeds
minerals, scrap, and residuesis essential for steady-state operation.
Algorithms that account for both mass and heat effects in autogenous or semiautogenous
reactors require comprehensive information about input and output streams. Using
distributed control systems, the solid-feed rate, typically controlled with impact-type
meters, can be systematically calibrated with more accurate weight loss readings from dry
feed bins. These bins are subject to both filling and emptying cycles; thus, direct use of
weight loss is not possible.
Recent developments in analytical techniques promise a revolution in the determination of
solids composition. Prompt gamma neutron activation spectroscopy is employed in power
and cement plants to provide on-line analysis.
43
The solids are irradiated with neutrons and,
in turn, emit gamma rays characteristic of the nuclei present and independent of matrix
effects. Analysis of the spectra involves significant data processing. However, once set up,
the technique can be used for continuous measurements over a moving belt. Alternatively,
the technique may provide quick chemical assays with simplified sample preparation steps
amenable to the shop floor.
Monitoring of pyrometallurgical processes is seriously hampered by the vulnerability of
sensors to high temperatures. Moreover, liquid and solid particulates in the reactor
atmosphere cause corrosion and erosion. Direct temperature measurement by insertion of
thermocouples into the reactor freeboard is often impractical because they burn or short out.
The development of two-wavelength pyrometers has improved temperature measurement,
but even these pyrometers can be affected by the atmosphere. Thus, temperature in many
pyrometallurgical reactors is currently determined by manual immersions during skimming
and tapping. An interesting approach, developed by Noranda, is temperature measurement
through tuyeres.
44
A retractable periscope mounted on the back of a tuyere transmits light
via a fiber-optic cable to a two-wavelength pyrometer located remotely from the reactor.
Accurately measuring matte and slag levels in a closed reactor is difficult. The widely used
technique of bar immersion is distinctly limited with respect to both accuracy and
applicability. Determination of reactor weight,
45
gamma radiation, lasers, and microwaves
can be used to measure levels of molten systems.
Gamma gauges for remote determination of bath levels have found application in the glass
industry.
46
The apparatus is mounted in the narrow forehearth area of the furnace and
comprises a transmitter located on one side and a receiver on the other. Application of the
technique to nonferrous applications (e.g., to indicate slag, matte, and metal levels) has so
far been limited due to large reactor widths.
47

Laser-based systems have found application in casting operations. In the aluminum
industry, lasers measure liquid levels in furnaces and also the rate of mold fillage.
48
Similar
applications apply to cast iron operations.
49
The application of lasers in nonferrous reactors
may be limited by the presence of dust and fume.
Electromagnetic microwaves hold promise for determining bath levels in continuous
oxygen pyrometallurgy. Microwaves are relatively insensitive to smoke and dust and are
not affected by high temperatures or temperature gradients as are ultrasonics. In the steel
industry, radar has been employed to measure metal level in basic oxygen furnaces and
torpedo cars.
50,51
Microwaves can also quantify the rate of rise during bottom teeming.
52
A
method of measuring slag thickness during casting operations has been identified.
53

Monitoring the progress of a steel converter with disposable oxygen probes based on
stabilized zirconia has long been an accepted part of the process. Commercial applications
in nonferrous systems are more limited. Today, highly reliable probes for measuring bath
oxygen potential in copper converters and anode furnaces are available,
54
and their use will
increase. However, the much desired continuously operating oxygen probe remains elusive
due to the sensitivity of the electrolyte. Other solid electrolyte systems are sensitive to CO
2
,
SOx, and NOx and may find commercial application in pyrometallurgy.
55
Still other
potential methods of monitoring reactor conditions include optical spectroscopy
56
and
continuous analysis of internal and exhaust reactor gases (e.g., O
2
, CO, CO
2
, H
2
, SO
2
) by
employing in-situ probes or sample withdrawals.
WINDOW ON THE FUTURE
A window on the future of pyrometallurgy is provided by metal making directly from
mineral concentrate in a single, closed, continuous oxygen converter. The impossible
dreams of continuous metalmaking directly from mineral feed have a long history. In 1870,
a textbook on metallurgy gave a detailed description of the "Siemens Process of Producing
Steel Direct from the Ore". It confidently stated: "The experiments on this important
process are now so far completed, that it is expected that the process will soon be
introduced into practice."
57
In 1896, Oliver Garretson described a logical process for
continuous copper smelting, converting, and slag cleanup, but it remained in two
dimensions.
58
In 1968, Howard Worner described his WORCRA concepts, "which seek to
maximize energy conservation" by "direct smelting-converting in one furnace in which
both smelting, dispersed-phase refining and slag conditioning and settling are combined in
distinct but communicating zones or branches." The genuine merit of his thinking was
demonstrated in years of pilot-plant operations, but commercial operations did not follow.
59

"The overall cost advantages which accrue from continuitynot least in respect to
environmental conservationare manifest. . . .There is no reason why hydrocarbon-
shielded oxygen jets cannot be advantageously employed for continuous subsurface-
blowing in nonferrous converting practice."
60
In 1974 the Q-S continuous oxygen converter
was publicized throughout the United States and was illustrated on the cover of JOM. The
inventors believed it would "prove to be a contribution to maximum economic utilization of
the nation's mineral heritage, with due regard to conservation of natural resources
including the environment."
61,62
Two decades later, commercial QSL (Queneau-
Schuhman-Lurgi) oxygen converters are continuously making metal directly from mineral
feed. A dream is finally a reality!
63-66

There are, of course, other dreams being pursued. For example, industry needs to fully
harness the energy released by the oxidation of SO
2
to SO
3
during acid production. This
energy could often produce sufficient 40 ats steam, for power generation, to supply all or
most of the amount required for oxygen production. We also need to improve our ability to
control the process parameters that characterize the ideal pyrometallurgical reactors of the
future. These will rapidly and continuously convert mineral sulfide concentrate and
appropriate recycled materials to acceptable quality metal, clean slag, and sulfur dioxide-
rich gas by fully utilizing the concentrate's natural fuel content in closed, fugitive emission-
free reactors. The chemical and steel industries are making great strides in process
monitoring (e.g., tomography), and the nonferrous industry must also follow their lead.
67,68

Oxygen pyrometallurgy has revolutionized the industry. The changes it has wrought can be
compared with developments of the turn-of-the-century decades (e.g., multihearth roasters,
Dwight-Lloyd sintering machines, huge reverberatory and open hearth furnaces, by-product
coke ovens, and Peirce-Smith converters). Today, oxygen usage is ubiquitous and
addictive. Substituting oxygen for air vastly increases process productivity and cleanliness.
Revolutionary sparks were ignited at Copper Cliff and Gerlafingen half a century ago.
However, until the winds of energy conservation and environmental protection blew
compellingly, the fires were confined. Now the fires burn briskly around the worldthe
future of pyrometallurgy is bright!
69-71

References
1. F.W. Davis, The Use of Oxygen or Oxygenated Air in Metallurgical and Allied
Processes, Report of Investigations no. 2502 (Washington, D.C.: Bureau of Mines, July
1923).
2. Rachel Carson, Silent Spring (New York: Houghton Mifflin, 1962).
3. Paul E. Queneau, "The Recovery of Nickel from Its Ores," JOM, 22 (10) (1970).
4. Fred Kaplan, "Norilsk, Russia, Mining and Metallurgical Works," The Boston Globe (17
November 1994).
5. T.E. Norman, Eng. and Min. J., (10-11) (1936), p. 137.
6. Inco Staff, "Operations and Plants of Inco," Canadian Mining Journal (May 1946); and
Inco Staff, "Oxygen Flash Smelting," JOM, 7 (7) (1955).
7. F.W. Starratt, "LDIn the Beginning," JOM, 12 (7) (1960).
8. R.W. Allgood, "Sulphuric Acid and Liquid Sulphur Dioxide Manufactured from Smelter
Gases at Copper Cliff, Ontario," CIMM Transactions, vol. LV (1952).
9. J.R. Gordon, G.W. Norman, P.E. Queneau, W.K. Sproule, C.E. Young, U.S. patent
2,668,107 (1954).
10. C. Landolt, A. Dutton, A. Fritz, and S. Segsworth, "Nickel & Copper Smelting at Inco's
Copper Cliff Smelter," Extractive Metallurgy of Copper, Nickel and Cobalt, Proceedings of
the Paul E. Queneau International Symposium, Vol. II, Copper and Nickel Smelter
Operations, ed. C.A. Landolt (Warrendale PA: TMS, 1993).
11. L.S. Austin, The Mineral Industry, ed. G.A. Roush and A. Butts (New York: McGraw-
Hill, 1919).
12. Paul Queneau, "Foreword," Extractive Metallurgy of Copper, Nickel and Cobalt,
Proceedings of the 1960 International Symposium, ed. Paul Queneau (New York:
Interscience Publishers, TMS, 1961).
13. R.H. Saddington, W. Curlook, and Paul Queneau, "Use of Tonnage Oxygen by Inco"
and "Foreword," Pyrometallurgical Processes in Nonferrous Metallurgy, eds. J.N.
Anderson and P.E. Queneau (New York: Gordon & Breach, 1967).
14. J.R. Boldt, Jr., and Paul Queneau, The Winning of Nickel (Toronto, Canada: Van
Nostrand, 1967).
15. Paul E. Queneau, "Oxygen Technology and Conservation," Metall. Trans. 8B, 3 (1977).
16. Paul Queneau and H.R. Roorda, "Nickel," Ullmanns Encyklopadie der Technischen
Chemie (Weinheim, Germany: Verlag Chemie, 1979).
17. J.G. Eacott, "The Role of Oxygen Potential and Use of Tonnage Oxygen in Copper
Smelting," Advances in Sulfide Smelting,,Vol. 2: Technology and Practice, ed. H.Y. Sohn,
D.B. George, and A.D. Zunkel (Warrendale, PA: TMS, 1983).
18. J.C. Yannopoulos and J.C. Agarwal, eds., Extractive Metallurgy of Copper
(Warrendale, PA: TMS, 1976).
19. Paul Queneau, "Coppermaking in the EightiesProductivity in Metal Extraction from
Sulfide Concentrates" JOM, 33 (2) (1981).
20. Paul Queneau, U.S. patent 3,004,846 (1961); P.E. Queneau and B. Kalling, U.S. patent
3,030,201 (1962); Paul Queneau and L.S. Renzoni, U.S. patent 3,069,254 (1962); W.
Curlook, C.E. O'Neill, and P.E. Queneau, U.S. patent 3,468,629 (1969); C.E. O'Neill, P.E.
Queneau, and J.S. Warner, U.S. patent 3,516,818 (1970); P.E. Queneau and C.E. O'Neill,
U.S. patent 3,615,361 (1971); and J.S. Warner and P.E. Queneau, U.S. patent 3,615,362
(1971).
21. Paul Queneau, C.E. O'Neill, A. Illis, and J.S. Warner, "Some Novel Aspects of the
Pyrometallurgy and Vapometallurgy of Nickel," JOM, 21 (7) (1969); and P.E. Queneau,
S.C. Townshend, R.S. Young, U.S. patent 294,883 (1960).
22. W.J. Thoburn and P.M. Tyroler, "Optimization of TBRC Operation and Control at
Inco's Copper Cliff Nickel Refinery" (Paper presented at the 18th Annual CIM Conference
of Metallurgists, Sudbury, Ontario, August 1979).
23. J.A. Blanco, T.N. Antonioni, C.A. Landolt, and G.J. Danyliw, "Oxy-Fuel Smelting in
Reverberatory Furnaces at Inco's Copper Cliff Smelter" (Paper presented at 50th Congress
of the Chilean Institute of Mining and Metallurgical Engineers, Santiago Chile, November
1980).
24. H. Schwarze, "Oxy-Fuel Burners Save Energy at El Teniente's Caletones Smelter,"
World Mining (May 1977).
25. T.N. Antonioni, J.A. Blanco, C.A. Landolt, and W.J. Middleton, "Energy Conservation
at Inco's Copper Cliff Smelter" (Paper presented at the TMS Annual Meeting, New York,
New York, February 24-28, 1985).
26. C.A. Landolt, A. Dutton, J.D. Edwards, and R.N. McDonald, "SO
2
Abatement, Energy
Conservation, and Productivity at Copper Cliff," JOM, 44 (1992), pp. 50-54.
27. M.C. Bell, J.A. Blanco, H. Davies, and P. Garritsen, "Taking Inco into the 1990's," CIM
Bulletin, 83 (January 1990), pp. 47-50.
28. C.A. Landolt, A. Fritz, S.W. Marcuson, R. B. Cowx, and J. Miszczak, "Copper Making
at Inco's Copper Cliff Smelter," Proceedings of Copper 91-Cobre 91 International
Symposium.,Vol IV: Pyrometallurgy of Copper, ed. C. Diaz, C. Landolt, A. Luraschi, and
C.J. Newman (New York: Pergamon Press, 1991), pp. 15-29.
29. M.C. Bell, J.A. Blanco, H. Davies, and R. Sridhar, "Oxygen Flash Smelting in a
Converter," JOM, 30 (10) (1978), pp. 9-14.
30. S.W. Marcuson, C. Diaz, and H. Davies, "Top-Blowing, Bottom-Stirring Process for
Producing Blister Copper," JOM, 46 (8) (1994), pp. 61-64.
31. C. Diaz, S. Marcuson, H. Davies, and R. Stratton-Crawley, "Conversion of Nickel and
Sulfur-Containing Copper to Blister," Proceedings of Copper '87, Vol. 4: Pyrometallurgy
of Copper, ed. C. Diaz, C. Landolt, and A. Luraschi (Santiago, Chile; Universidad de Chile,
1988), pp. 293-304.
32. Robert Lee, "Innovations in Ferrous PyrometallurgyA Canadian Perspective," CIM
Bulletin, 84 (June 1991), pp. 125-131.
33. A. Otero, J.K. Brimacombe, and G.G. Richards, "Kinetics of the Flash Reaction of
Copper Concentrate," in Ref. 28, pp. 459-472.
34. G.S. Victorovich, "Oxygen Flash Converting for Production of Copper," Extractive
Metallurgy of Copper, Nickel and Cobalt: Proceedings of the Paul E. Queneau
International Symposium. Vol. I. Fundamental Aspects, ed. R.G. Reddy (Warrendale PA:
TMS, 1993), pp. 623-637.
35. H. Davies, S. Marcuson, G. Osborne, and A. Warner, "Flash Converting of Chalcocite
Concentrate at Inco's Port Colborne Pilot Plant," in Ref. 34, pp. 623-639.
36. C.A. Landolt, A. Dutton, T. Fritz, and S. Marcuson, "New Smelter Furnaces and Novel
Copper Processing," The 96th Annual General Meeting of the CIM and the 1994 Mineral
Outlook Conference, ed. N. Champigny and P. Dillon (Montreal, Canada: CIM, 1994), pp.
69-71.
37. C. Diaz, S.W. Marcuson, A. Warner, and G.E. Osborne, "Reduced Dusting Bath
System for Metallurgical Treatment of Sulfide Materials," U.S. patent application
08/401081: filing date 8 March 1995.
38. D. Eyre, I. Gorup, and T. Pawulski, "Production of OxygenKeeping Pace with the
Metallurgical Demands," The Impact of Oxygen on the Productivity of Non-Ferrous
Metallurgical Processes, ed. G. Kachanivsky and C. Newman (Toronto, Canada: Pergamon
Press, 1987), pp. 77-85.
39. D.C. King, R. L. Hurchison, K.J. Murphy, and A. Odorski, "The Benefits of Optimizing
Air Separation Plant Performance,"in Ref. 38, pp. 199-208.
40. K.J. Murphy, A.P. Odorski, A.R. Smith, and T.J. Ward, "Oxygen Production
Technologies for Non-Ferrous Smelting Applications," in Ref. 38, pp. 219-235.
41. T.S. Pawulski, "Cryogenic Oxygen PlantsAn Overview," in Ref. 38, pp. 121-134.
42. D.A. Eyre, Air Liquide Engineering, private communication with authors (5 October
1995).
43. J. Makansi, "PSI Gibson Turns Compliance into a Vision for the Future," Power
(December 1993), pp. 37-40.
44. A. Pelletier, J.M. Lucas, and P.J. Mackey, "The Noranda Tuyere Pyrometer: A New
Approach to Furnace Temperature Measurement," in Ref. 31, pp. 489-508.
45. H.W. Grenfell, D.J. Bowen, and C. McQueen, "The Role of Continuous Vessel
Weighing in the Commissioning and Operation of B.S.C. Ravenscraig New No. 3 B.O.F.,"
Proc. Natl. Open Hearth Basic Oxygen Steel Conf., vol. 60 (1977), pp. 209-221.
46. "CND Continuous Level Gauge," product brochure CN-158 (Round Rock, Texas: TN
Technologies, 1995).
47. Tony Hart, TN Technologies, private communication with authors (10 March 1995).
48. "Selcom Laser Sensors and LaserPour Systems for Aluminium Level Control," product
brochure (Southfield, MI: Selective Electronics, 1995).
49. D.P. Kanicki and B.R. Krohn, "Taking the Heat Off Molten Metal Handling II-
Ferrous," Modern Casting, 74 (November 1984), pp. 27-30.
50. K.G. Crudgington and M.E. London, "Non-contact Measurement of Molten Metal in
Torpedo Ladles Using Microwaves," Measurement + Control, 23 (December/January
1990/91), pp. 303-305.
51. R.C. Novak, "BOF Bath Level Measurement at Burns Harbor," 75th Steelmaking
Conference Proceedings (Warrendale, PA: ISS, 1992), pp. 169-172.
52. A. Zeewy, L. Peltz, and A.M. Freborg, "Advanced Microwave Technology Improves
Bottom Poured Ingot Quality," I&SM (June 1993), pp. 45-49.
53. A. Zeewy, C.J. Bingel, and D.G. Hargreaves, "Microwave-Driven Slag Thickness
Measurement," 9th Process Technology Division Conference Proceedings (Warrendale,
PA: ISS, 1990), pp. 13-16.
54. S.W. Marcuson, S. Tessier, A. Vahed, A. Fritz, and C. Diaz, "Use of Oxygen Probes in
Copper Converting at Inco's Copper Cliff Smelter," Copper '95-Cobre '95 Proceedings,
Vol. 4, Pyrometallurgy of Copper, ed. W.J. Chen, C. Diaz, A. Luraschi, and P.J. Mackey
(Montreal, Canada: CIM, 1996), pp. 271-279.
55. T. Maruyama, "Solid Electrolyte Sensors for Gaseous Oxides for Pollution
Monitoring," Mater. Sci. Eng., A146, pp. 81-89.
56. W. Wendt, M. Alden, B. Bjorkman, T. Lehner, and W. Persson, "Controlling Copper
Conversion via Optical Spectroscopy," JOM, 39 (1987), pp. 14-17.
57. William Crookes and Ernst Rohrig, Practical Treatise on Metallurgy (New York: John
Wiley & Son, 1870).
58. Oliver Garretson, U.S. patent 596,992 (1896).
59. Howard Worner, "Continuous Smelting and Refining by the WORCRA Processes,"
Advances in Extractive Metallurgy (London: IMM, 1968).
60. Paul E. Queneau, "Modern Practice and Technological Innovation in the Nonferrous
Industries," JOM, 25 (1) (1973), pp. 15-18.
61. Staff Reporter, "St. Joe Minerals Corp. Has Exclusive Option on New Lead Process,"
The Wall Street Journal (22 February 1974).
62. Paul E. Queneau and Reinhardt Schuhmann, Jr., "The Q-S Oxygen Process," JOM, 26
(8) (August 1974), pp. 14-16.
63. R. Schuhmann, Jr., "Measurement, Interpretation and Control of Oxygen Activity in
Pyrometallurgical Processes," Proceedings of the Reinhardt Schuhmann International
Symposium on Innovative Technology and Reactor Design in Extraction Metallurgy, ed.
D.R. Gaskell, J.P. Hager, J.E. Hoffmann, and P.J. Mackey (Warrendale, PA: TMS, 1986).
64. H.A. Kellogg and C. Diaz, "Bath Smelting Processes in Non-ferrous Pyrometallurgy
An Overview," Proceedings of the Savard/Lee International Symposium on Bath Smelting,
ed. J.K. Brimacombe, P.J. Mackey, G.J.W. Kor, C. Beckert, and M.G. Ranada (Warrendale,
PA: TMS, 1992).
65. Paul E. Queneau, "The Coppermaking QS Continuous Oxygen Converter
Technology, Design and Offspring," in Ref. 34.
66. "Recent Metallurgical Plants," Mining Magazine (London) (August 1995).
67. G. Ondrey and G. Parkinson, "Process Tomography: Seeing is Believing," Chemical
Engineering (October 1995), pp. 30-33.
68. J. Reidel and S. Kohle, "Methods for Continuous Monitoring in Steelmaking
Processes," Metallurgical Processes for the Early Twenty-First Century, ed. H.Y. Sohn
(Warrendale, PA: TMS, 1995), pp. 799-812.
69. Anon., "Forty Years of BOP Steelmaking," 33 Metal Producing (March 1992).
70. Carlos Diaz, Hermann Schwarze, and John C. Taylor, "The Changing Landscape of
Copper Smelting in the Americas," in Ref. 54.
71. Paul E. Queneau and Martin Hirsch, "Process for the Manufacture of Steel," U.S. patent
5,466,278 (November 14, 1995); and "Process for the Continuous Manufacture of Steel,"
U.S. patent application 08/503,710; filing date: 18 July 95.
*Development of porous ceramic plugs began in 1947 and
was led by Steven Spire and Robert Lee of Canadian Liquid Air in
conjunction
with the Canadian Bureau of Mines.
3
In the 1970's, porous plug use
became widespread in the steel industry. However, their use in nonferrous
pyrometallurgy has been limited, and the usage described here represents
its
first commercial application in copper smelting.

S-ar putea să vă placă și