Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/301292769

SONIC INJECTION IN BATH SMELTING AND CONVERTING: MYTHS, FACTS AND


DREAMS

Conference Paper · October 2013

CITATIONS READS

7 1,124

1 author:

Joel Kapusta
BBA
24 PUBLICATIONS   152 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Optimized Sonic Injectors for Bath Smelting and Converting View project

All content following this page was uploaded by Joel Kapusta on 17 April 2016.

The user has requested enhancement of the downloaded file.


RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 267

SONIC INJECTION IN BATH SMELTING AND CONVERTING:


MYTHS, FACTS AND DREAMS
Joël P. Kapusta

BBA Inc.
630, René-Lévesque Blvd. West, Suite 1900
Montréal, Québec, H3B 4V5, Canada
(joel.kapusta@bba.ca)

Keywords: Converting, Gas Injection, Sonic Velocity, Concentric Tuyere, Plant Trials

Abstract

The steelmaking and non-ferrous metals industries have often been viewed as ‘two solitudes’ in
reference to their different approaches to similar processing issues. Submerged gas injection
presents one such example. While sonic injection has revolutionized steelmaking (e.g. Q-BOP,
K-OBM, and AOD processes), it has essentially not been adopted in sulfide bath smelting and
converting which are still plagued with operational problems due to the limitations of low
pressure injection. This paper provides a historical perspective on Canadian efforts in research
and development in the field of sonic injection in copper and nickel converting – including plant
trials conducted in the 80s – and briefly reviews its commercial implementation in the late 90s
and 2000s. Throughout the paper, the author will attempt to dispel some myths, elaborate on key
facts, and also share a few dreams and a view of the future.

Introduction

“It is unlikely that current blowing practice in converters can be tolerated indefinitely […] It
may well be that the nonferrous converter of tomorrow bears a striking resemblance to the
steelmaking furnace of today.” These were the words of Hoefele and Brimacombe in 1979 in
their ground breaking paper entitled “Flow Regimes in Submerged Gas Injection” [1]. More than
30 years later, the blowing practice of nonferrous bath smelting and converting vessels has
essentially remained unchanged: low pressure, single-pipe, submerged tuyere injection with its
associated problems of accretion buildup, tuyere blockage and punching, bath slopping and
splashing, and limited forward penetration of horizontally injected gases.

New technologies such as flash or top-submerged-lancing (TSL) have certainly earned a growing
share of the smelting capacity, and to some extent of the converting capacity. Yet, Peirce-Smith
converters have retained the largest share for copper and nickel converting. One can only wonder
why such little progress has seemingly been achieved by the nonferrous metals industry over the
last three decades while the steel industry has practiced high velocity gas injection for over five
decades. The answer does not lie in a lack of research on gas injection dynamics. On the
contrary, a wealth of published research exists on the topic as demonstrated by the exhaustive
review paper by Brimacombe, Nakanishi, Anagbo and Richards [1] in 1990 with more than 270
references!
268 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

Sonic injection trials in nonferrous converters have also been conducted and the concept of free
blowing (or punchless operation) has been proved. Adoption of the technology has not occurred,
most often on the basis that replacing a 103 kPag (15 psig) blower with a 345 kPag (50 psig)
compressor would increase the annual operating cost beyond the perceived or estimated potential
benefits of the technology. As far back as the late 70s, the consensus in the nonferrous industry
was that for a new converter the capital and operating costs of a blower and punching machine
would be similar to those of a 50 psig compressor, and therefore, implementing sonic injection
for new converters might be attractive. Numerous new vessels have been built and commissioned
in the last 30 years or so, and yet, to the best knowledge of the author, none has been equipped
with sonic tuyeres outside of China and apart from Xstrata Nickel in Canada (now Glencore
Xstrata) and Thai Copper Industries in Thailand.

A more accurate reason than solely economics, as suggested by Mackey and Brimacombe [3],
might be that the growth in oxygen usage in nonferrous pyrometallurgy has been at a slower pace
compared to the steel industry. As much as the steel converter needed a high oxygen tuyere to
overcome the productivity and efficiency limits it had reached, the autogenous nonferrous
converter had not hit a similar productivity barrier. Moreover, enriching the blast air of
converters by a few percent of oxygen has generally been sufficient for debottlenecking.

Times have changed. Nonferrous smelters are undermined by economic pressures and lower
profit margins. Sustaining the viability of the copper and nickel bath smelting furnaces of today
may require the same leap forward than the steel converter of the 60s experienced with the
Savard-Lee tuyere. The time may have come for bath smelting and converting of sulfide mattes
to re-evaluate the potential of high oxygen sonic injection and consider the technology as part of
the drive (feasibility studies) to boost productivity, improve energy efficiency, or lower the
environmental footprint by reducing off-gas volumes. The pressure on existing bath smelting
operations to achieve those goals is intensified by the need to do so at a time when available
capital is at particularly low levels.

In the 60s, understanding gas injection phenomena to improve pyrometallurgical operations


became a critical need of the industry as demonstrated by the universality of research on the
topic. A wealth of knowledge exists in the chemical engineering literature. Metallurgists brought
their own issues of hot molten metals and created further knowledge specific to pyrometallurgy.
Notable research was conducted by A. Wraith and F.D. Richardson in the UK, A. Spesivtsev and
L. Shalygin in Russia, D. Robertson in the UK and USA, K. Mori, M. Sano and T. Kimura in
Japan, E. Burström and G. Carlsson in Sweden, N. Gray and M. Nilmani in Australia, to name
only a few. The focus of this paper, however, is not to review the global wealth of knowledge.
Rather, the aim is to provide a historical perspective on the R&D efforts, and notably the plant
trials, conducted by Canadian researchers, engineers and operators. Their know-how should not
be lost and must be transmitted to the new generation of metallurgists.

During my years of collaboration with Professor Ralph Harris as an adjunct professor at McGill
University, I was very fortunate to interact with Ralph’s graduate students in extractive
metallurgy. Ralph was so genuinely committed to his students, always striving to motivate and
support their inquisitiveness in the course of their research. In this stimulating climate, I felt
compelled to share my own passion and contribute with Ralph, even if modestly, to keeping
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 269

alive the Canadian legacy in gas injection phenomena. Such efforts in pyrometallurgy should
continue in Canadian universities so that the Canadian know-how can be passed to new
generations and put to practice in plants around the world. The Ralph L. Harris Memorial
Symposium is in my opinion one step towards this goal and dream.

Early Gas Injection Research – The Pioneers in Canada

Canadian Liquid Air and Noranda pioneered research on gas injection in Canada from the 40s
through to the 60s. Highlights of their key efforts are briefly summarized in this section.

Savard-Lee Developments at Canadian Liquid Air – 1948 to 1968

The discovery by Bessemer in 1855 that pig iron could be autogenously blown with air led to the
Bessemer converter, a first major milestone in the steel industry [4]. Although Bessemer
included the use of oxygen in his patent, its implementation was not possible at that time as
oxygen would cause severe erosion of the bottom of the reactor. In fact, submerged oxygen
injection was not used for over a century, until the development in Canada of the concentric
tuyere by Messieurs Guy Savard and Robert Lee of Canadian Liquid Air [5-6]. In the words of
Mackey and Brimacombe [3], Savard and Lee discovered how to turn into reality the hundred
year old “dream to remove the dead hand of nitrogen from the Bessemer converter.” The first
industrial implementation of the Savard-Lee tuyere went on stream in Germany in 1967 at the
Eisenwerk-Gesellschaft Maximilianshütte (Maxhütte). The process was called OBM for Oxygen
Bottom Metallurgy Maxhütte, and was patented by Maxhütte and Air Liquide [7]. In the Savard-
Lee concentric tuyere, the oxygen stream at sonic velocity is shrouded with a hydrocarbon gas
that cracks at steelmaking temperatures and provides local cooling at the tuyere tip. By
shrouding the oxygen jet with a medium that is non-reactive with the melt but reactive with
oxygen, the injector is protected, in particular in the critical zone at the injector-refractory
interface where refractory wear occurs.

Details of the two-decade long development of the Savard-Lee concentric tuyere are out of the
scope of this paper and have been published elsewhere [3, 8-9]; only a brief summary is
presented. The quest for bottom oxygen injection started in 1947 at the DOSCO steel mill in
Sydney, Nova Scotia. The first trials were actually an attempt to desiliconize molten iron via
supersonic top lancing of oxygen in a ladle. Over a period of 20 years, a series of trials and
demonstrations took place in Canada and Europe with oxygen injection via top lancing, high
pressure around 4 135 kPag (600 psig), single pipe bottom tuyere, as shown on Figure 1 (cooling
by Joule-Thompson effect with protective mushroom formation), and finally concentric tuyere
with a hydrocarbon as coolant and oxygen getter allowing the generation and control of
protective accretions (mushrooms) at the tuyere tips. The early tests of the concentric tuyere
occurred in 1964 in the K-Vessel (named after its designer Bob Kottmeier of Canadian Liquid
Air and shown on Figure 2). The first industrial 20-tonne heat was poured on 17 December 1967
in a modified Thomas converter at the Eisenwerk-Gesellschaft Maximilianshütte in Sulzbach-
Rosenberg, Germany, just four months after Dr. Karl Brotzmann, Director of Research at
Maxhütte, had witnessed two small scale demonstrations in Canada. Those tests were conducted
at the Freeman Corporation foundry in the town of Cap-de-la-Madeleine, some 150 km east of
Montreal.
270 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

The concentric tuyere has spawned a family of steel processes such as the Quick Basic Oxygen
Process (Q-BOP), the US counterpart to the OBM, the Argon Oxygen Decarburization (AOD),
the Klockner Oxygen Blown Maxhütte (K-OBM), and has also found application in non-ferrous
metal processing. An international symposium was held in Montreal in 1992 to honor Savard and
Lee and provide a record of their achievements and of those who followed their lead in
concentric injection [10].

Figure 1. The first bottom blown vessel (modified ladle) used in 1956 for desiliconizing
trials at DOSCO in Sydney, Nova Scotia (Archives of Air Liquide Canada).

Figure 2. The experimental K-vessel built for tests and demonstrations of the Savard-
Lee concentric tuyere between 1964 and 1967 (Archives of Air Liquide Canada).

Developments at the Noranda Research Centre in the 60s

A great deal of research was conducted by Noranda in the 60s, at a time when major
developments were underway in both matte smelting and converting. The Noranda Reactor is
one example of a major Canadian contribution to the pyrometallurgy of copper. In 1969,
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 271

Themelis, Tarassoff and Szekely [11] noted that very few studies had been reported on the
dynamics of submerged gas injection into converters. They believed that a better understanding
of gas-matte interactions would provide a more rational basis to operate and design converters.
Their experimental setup consisted of air injection into water in Plexiglass models of a Peirce-
Smith converter. Using a photographic technique, they were able to visualize and measure the
trajectory of the air jets into water at various tuyere diameters and gas velocities. Some of their
photographs and graphs are reproduced for illustration in Figure 3 and Figure 4.

(a) (b)

(c)
Figure 3. Photographs of an air jet in water: (a) short time exposure (0.6 milliseconds)
illustrating the unstable jet, (b) long time exposure (5 seconds) visualizing the jet cone,
and (c) long time exposure visualizing influence of buoyancy [11].

Figure 4. Experimental results and predicted jet trajectory under the influence of
buoyancy [11].
272 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

In their experiments they measured an average jet cone angle of 20 degrees and observed that
buoyancy had little effect on the radial distribution of velocities and concentration (axial
symmetry of the jet). Their next step was to provide a complete formulation based on continuity
and momentum balances that mathematically described the air jet trajectory (Figure 5). Their
equation was a function of the Froude number of the air-water system and was in very good
agreement with their experiments. Based upon these observations, they extrapolated their water
model results to the air-matte system of a copper converter. A comparison of the air jet trajectory
and penetration in water and copper matte are presented in Figure 6 and the famed model for
copper converting in Figure 7 with a calculated modified Froude number of 157.

Figure 5. Diagram of a jet trajectory with its mathematical characterization [11].

Figure 6. Effect of the modified Froude number on Figure 7. Idealized air jet trajectory
the trajectory of a horizontally injected air jet into and penetration in a copper
water (top) and copper matte (bottom) [11]. converter [11].
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 273

Interestingly, a re-examination of the calculations from Themelis et al. [11] showed that their
modified Froude number value of 157 in Figure 7 is in fact incorrect under the conditions given
with their figure. The correct value should have been 15.7, a more typical number for Peirce-
Smith converting, which according to Figure 6, would have indicated that air injected
horizontally into copper matte did not penetrate into the bath but rose immediately above the
tuyere line: a fact that would later be ‘discovered’ and proven by Brimacombe and Hoefele [1].
A summary of their discovery is provided in the next section.

Another noteworthy achievement of Noranda in the 60s is the research into gaseous copper
refining to eliminate the slow, inconvenient and polluting practice of poling with birch poles.
After much experimentation, a high pressure tuyere was developed for gaseous propane or
natural gas injection and implemented in the anode furnaces of the Horne Smelter [12]. Although
the injection of air and natural gas was done through only one tuyere per anode furnace, that
tuyere was operated at high pressure to avoid blockage. This tuyere may have been the first
recorded commercial implementation of sonic injection in the non-ferrous industry. In a recent
discussion, Phillip Mackey put this achievement back into context: the development was just for
one single tuyere operating in a copper bath with little or no slag [13]. Although an achievement
in terms of sonic injection for the time – the late 60s – the non-ferrous industry was not yet ready
to consider implementing, without further research, the high pressure injection concept in a
converter with a complete tuyere line blowing into a 3-phase molten bath of matte, slag, and
copper. This technical challenge was for metallurgists to tackle in the coming next two decades.

Breakthrough in the 70s – The Brimacombe Group at the University of British Columbia

In the 50s and 60s, researchers in the chemical engineering field devoted great efforts in
characterizing the size and velocity of gas bubbles in liquids. Metallurgists carried those research
efforts into the 70s and 80s and developed mathematical models of varying complexity to
calculate the theoretical volume of submerged gas bubbles as a function of gas flow rate, tuyere
diameter and submergence, molten bath density, and other significant parameters. They also
developed and used novel techniques to examine the dynamics of submerged gas jets in the
laboratory and in operating copper and nickel operations.

Inspired by his time spent in 1971 at the Noranda Research Centre in Pointe-Claire, Quebec,
Brimacombe spear-headed new research in gas injection at the University of British Columbia
(UBC) in Vancouver. In the mid-70s, Oryall and Brimacombe carried out similar experiments as
those of Themelis et al. [11] in the late 60s, and compared air injection in water and
mercury [14]. They demonstrated that the properties of the bath had a major influence on the gas
jet penetration. They confirmed that the model of Themelis et al. [11] was indeed valid for
injection in water but was over estimating the actual gas penetration for injection into liquid
mercury. In fact their measurements revealed that the jet expansion angle in mercury ranged
between 150 and 155 degrees, a value more than seven times greater than the angle of air jets in
water. This finding strongly indicated that the physical properties of the gas and liquid (or molten
bath) had a considerable impact on the jet dynamics. Using an electroresistivity probe combined
with an oscilloscope, Oryall and Brimacombe [14] found that the air jets penetrated both forward
and backward and that the forward penetration in mercury was very limited. Although injected
horizontally, the air jets appeared as if vertically injected.
274 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

In the 70s, a new technique adopted (developed) for both laboratory and plant investigation was
to measure the pressure fluctuations at the back of tuyeres using very sensitive sensors with a fast
response (in the range of 1/10th to 1/100th of a second). Hoefele and Brimacombe [1] adopted the
technique and used piezoelectric transducers as sensors that they coupled, in their experiments,
with an amplifier and a recording oscilloscope. With this approach, they successfully recorded
different types of oscillations for increasing gas flow rate. Using high speed films, they
correlated the pressure oscillations to the gas dynamic events occurring at the tuyere tip and
delineated two regimes of flow, bubbling and steady jetting, as a function of the modified Froude
number and the ratio of gas to liquid densities. At low gas flow rate and with a Mach number
around 0.3 (bubbling regime), they measured oscillation frequencies in the order of 10 to 12 s-1
corresponding to a frequency of bubble discharge of 10 to 12 bubbles per second, as shown on
Figure 8. At higher gas flow rate approaching underexpanded flow (Mach number close to 1.0),
the shape and frequency of the pulses become irregular, while for fully underexpanded flow
(jetting), the pressure is practically steady with only a few sharp pulses, as shown on Figure 9.

Hoefele and Brimacombe completed their research by measuring the back-pressure fluctuations
of a tuyere during normal operation of a nickel converter at Inco’s Thompson Smelter (now
Vale) in Manitoba. The pressure traces they observed at the campaign start of a newly relined
converter were similar to those measured in the laboratory with clear and regular pulses with a
frequency of 10 to 12 s-1 (see Figure 10). Hoefele and Brimacombe concluded that horizontally
injected air into a converter was actually discharging as discreet bubbles, revoking the myth that
air injected at low pressure in a bath discharged as a jet. This validation that normal blowing of a
converter was a bubbling rather than jetting system was a major breakthrough for the
understanding of converter operation and also submerged gas injection phenomena.

Figure 8. Pressure traces measured in the Figure 9. Pressure traces measured in the
laboratory under bubbling regime (NFr’ = 52 laboratory under jetting regime (NFr’ = 900
and NMa = 0.34) [1]. and NMao = 1.28) [1].
ρ g u 02
Note: NFr’ is the modified Froude number where ρg and ρl are the gas and liquid densities (g. cm3), uo
g (ρ l − ρ g ) d 0
is the bulk gas velocity in the tuyere (cm. s-1), do is the tuyere diameter (cm), and g is the constant of gravitational
acceleration (cm. s-2). NMa is the Mach number. NMao is the nominal Mach number for underexpanded flow defined
as the Mach number that would be obtained just beyond the tip of the tuyere if the gas discharging from the tip
accelerated uniformly in the flow direction and attained the local pressure measured at the tip. This is a fictitious
number which still gives a measure of the degree of underexpansion of the gas jet.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 275

Using the normal blowing rate of the Thompson Smelter, about 1 200 Nm3/h at the time of their
plant measurements, Hoefele and Brimacombe calculated the volume of individual bubbles (at
10 bubbles per second) to be very large: 33 000 cm3 at 20°C, corresponding to a 40 cm diameter
bubble, or 166 000 cm3 if the air is equilibrated with the bath at 1200°C, corresponding to a
68 cm diameter bubble (see Figure 11).

Figure 10. Pressure traces measured in the Figure 11. Schematic cross section of a
Thomson Smelter nickel converter under nickel converter showing bubble sizes at
normal blowing conditions (middle of the 20, 323, and 1200°C rising in close
bank tuyere) [1]. proximity of the back wall [1].

The finding that converting operations are characterized by large discrete bubbles of oxygen
containing gas rising vertically above the tuyere tips was indeed a milestone in the 70s. This gas
behavior with tuyeres mounted flush with the refractory of converters suggested that the high
refractory erosion at the back wall was directly related to the dynamics of gas injection.
Measuring the back pressure of a steel pipe inserted in a tuyere and connected to high pressure
air, Hoefele and Brimacombe also validated their laboratory experiments that above a critical
back pressure, about 345 kPag (50 psig) in their plant measurements, the air injected in the
converter becomes underexpanded and discharges as a steady jet with a greater penetration into
the bath. They stipulated that this jetting regime at higher injection pressure may offer major
benefits to converter operation, particularly in back wall erosion as previously suggested by
Oryall and Brimacombe [14]. They speculated that underexpanded jetting injection may lower
the need for tuyere punching due to the greater momentum of the air jet.

Those findings elucidated many questions as well as raised more new questions. Would a high
pressure tuyere be self-blowing and not require punching? Would jetting injection influence the
oxygen efficiency? Would jetting increase slopping or splashing of the bath? Only full scale
plant trials would provide all the answers. Those were to come in the following decade.

Plan Trials in the 80s – Proving the Sonic Injection Concept

The much anticipated plant trials of sonic injection in converting occurred at three smelters in
North America: first, in April 1981 at the Asarco copper smelter in Tacoma, Washington State,
276 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

USA (smelter closed in 1985), then in August 1985 at the Noranda Horne smelter in Rouyn-
Noranda, Quebec, Canada (now Glencore Xstrata), and lastly in November 1985 at the Inco
Copper Cliff smelter in Sudbury, Ontario (now Vale). Those three documented plant trials are
briefly described and their outcome summarized.

Asarco Tacoma Smelter – Plant Trials (April-August 1981)

In 1980, Brimacombe and Hoefele were granted a patent for their non-ferrous metal converting
process using a high pressure (or sonic) tuyere [15]. With Canadian Liquid Air the assignee of
the patent, Robert Lee approached Asarco in May 1980 with the idea of testing the high pressure
tuyere as a joint venture. Under the agreement, Air Liquide was to provide technical assistance
as well as the oxygen installation for the tests. Asarco was to install the necessary piping and
instrumentation. The partnership agreement also addressed the matters of royalties on possible
future use of the technology by Asarco, should the tests be positive, conclusive and should
Asarco decide to implement the technology. Only highlights of the trials are presented in the
section below as complete details were published by Brimacombe, Meredith and Lee [16].
Converter No. 4 was selected for the trials and some details of its typical practice in 1981 are
provided in Table I.

Table I. Normal Converter Practice at Asarco Tacoma Smelter in April-August 1981


Converter No. 4 Details
Converter diameter and length 4.0 m by 10.7 m (13 ft. by 35 ft.)
Number / Size of tuyeres 52 with 44.5 mm I.D. (1.75 in.)
305 mm (12 in.) for slag blows
Tuyere submergence
610 mm (24 in.) for finishing blows
Punching machine Gaspé type
Injection pressure 103 kPag (15 psig) for slag and finishing blows
750 Nm3/h/tuyere for slag blows
Injection flow rate
620 Nm3/h/tuyere for finishing blows
Oxygen enrichment (efficiency) 26% (63%)
Campaign life (No. of charges/campaign) 375 days (288 charges)
Copper produced/campaign 25 000 tonnes

For the initial tests, four 12.7 mm pipes were inserted into the first four regular tuyeres at the
south end of Converter No. 4 and refractory mix was packed in the annulus between the inserted
pipes and the regular tuyere pipes. The trials started on 21 April 1981 on charge No. 14. They
were very short lived (only one charge). Towards the end of the charge, one of the high pressure
tuyeres failed and copper ran through the back. The remaining three high pressure tuyeres were
examined. They were both warped and twisted, although not blocked. The consensus amongst
those involved in the trials was that the construction of the high pressure tuyeres – inserted pipes
and refractory mix – and the insufficient cooling due to a low flow rate per tuyere were the
reasons for the failure. Thick-walled, larger diameter tuyeres (44.4 mm O.D. and 19.1 mm I.D.)
were designed and installed, replacing four regular tuyeres.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 277

The second sonic injection trials started on 5 May 1981 on charge No. 28 (same lining as for the
first test). Details of the operating conditions for the sonic injection trials are summarized in
Table II. At the end of charge No. 28, the back of the high pressure tuyeres were opened (cap
removed) and the pipes inspected. They were found clean and showed no blockage. Inspection of
the air flow chart, presented on Figure 12, showed that the blast flow rate remained constant
during the charge. This first successful sonic injection converting cycle was accomplished
without punching, a success that warrantied continuing the trials.

Figure 12. Original air flow chart recording from 5 May 1981 (charge No. 28) at the
Asarco Tacoma Smelter (Archives of Air Liquide Canada).
278 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

The next charges were similarly blown without any trouble from the test sonic tuyeres. After
each charge, the back of the tuyeres were opened and at times, some buildup at the tip of the
tuyeres was observed. This buildup was however light, easily removed with a cleanout bar (as a
precaution), and did not hinder the blast flow. As the oxygen delivery system became
operational, oxygen enrichment of the blast was initiated on 14 May 1981. At enrichment levels
around 28%, no buildup formed around the tuyere tip and individual bricks became visible. The
enrichment was reduced and operation of the sonic tuyeres continued without any problem for
another 31 charges. The buildup that did reappear at lower oxygen enrichment may have been
viewed as problematic at the time; with hindsight and more trials, we now know this buildup
actually provides protection at the tuyere tips and surrounding refractory.

Although the sonic tuyeres operated at constant flow rate for several more weeks, the
conventional low pressure tuyeres adjacent to the sonic tuyeres started to break out and had to be
plugged with refractory mud. Since visual inspection through the converter mouth did not reveal
any crack, hole or damage to the refractory, the trials were continued until 2 August 1981 after
completion of charge No. 116. Since the sonic tuyeres were affecting the adjacent low pressure
tuyeres, the trials were stopped to ensure integrity of the converter. The high pressure trial period
lasted 89 days and 88 charges. Details of the test conditions are summarized in Table II.

Table II. High Pressure Trial Parameters at Asarco Tacoma Smelter (April-August 1981)
First Trial Second Trial
Number of sonic tuyeres 4 4
Pipes inserted into first Sonic tuyeres replacing
Arrangement details
4 regular tuyeres 4 regular tuyeres
19.1 mm I.D. (0.75 in)
Size of sonic pipes or tuyeres 12.7 mm I.D. (0.5 in)
44.4 mm O.D. (1.75 in)
Sonic injection pressure 427 kPag (62 psig) 414 kPag (60 psig)
Blast injection flow rate 255 Nm3/h/tuyere 680 Nm3/h/tuyere
Trial duration (No. of charges) 1 day (1 charge) 89 days (88 charges)

Conclusions – Asarco Tacoma Trials. The goal of the trials was to assess the merits of sonic
tuyere injection. The conclusions drawn from the trials were as follows:

• The sonic tuyeres do not block and operate without punching as suggested by Hoefele
and Brimacombe in 1979 [1];
• The sonic tuyeres are expected to operate in punchless mode throughout the life of the
refractory lining;
• The sonic tuyeres operate at constant blast flow rate during the entire blowing cycle in
contrast to conventional low-pressure tuyeres experiencing variable flow rates as a
consequence of tuyere blockages;
• The sonic tuyeres are silent and require little or no maintenance;
• The sonic tuyeres generate accretions of varying sizes depending on oxygen enrichment.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 279

Unfortunately, the impact of sonic injection on refractory erosion at the tuyere line could not be
assessed conclusively due to the limited number of sonic tuyeres tested. If the technical and
operational feasibility of sonic tuyeres was demonstrated by the trials, a major accomplishment
in itself, the economic viability of the sonic injection practice could not be proven. Only a full
scale demonstration with a converter solely equipped with sonic tuyeres and with the refractory
wear and productivity continuously monitored would provide the answer.

Copper Converting – Plant Observations and Measurements (1982 to 1985)

In the early 80s, Alejandro Bustos joined the Brimacombe team at UBC and focused his research
on copper and nickel converting. As an integral part of his Ph.D. activities, Bustos visited several
smelters, included Asarco Tacoma, Noranda Horne, Utah Copper and Inspiration Copper. During
his stay in the converting aisle of those smelters, he made a number of observations on accretions
formation using a custom-designed “tuyerescope”, as shown on Figure 13, and carefully
documented his observations photographically. Figure 14 presents a sequence of photographs
used by Bustos, Richards, Gray and Brimacombe [1] to illustrate the dynamics of accretion
formation.

Bustos also measured tuyere back-pressure at several of those smelters, using the same
piezoelectric sensor than Hoefele and Brimacombe [1] and providing further evidence of the
bubbling nature of conventional, low pressure, gas injection into nonferrous metal converters.
Combining plant observations with the wealth of results from laboratory work at UBC, including
research led by G.G. Richards on bath motion, bath slopping and splashing [18], Brimacombe,
Bustos, Jorgensen and Richards published their comprehensive paper titled “Towards a Basic
Understanding of Injection Phenomena in the Copper Converter,” a major compilation of the
new knowledge gained at UBC on submerged gas injection in converters [19].

(a) (b)
Figure 13. (a) Schematics of “Tuyerescope” mounted on tuyere body of a Peirce-Smith
converter [1], and (b) “Tuyerescope” and camera assembly attached at back of tuyere body for
photographic recording of accretion dynamics at the tips of tuyeres. Note the two pipe-type
accretions on the table (Archives of A.A. Bustos).
280 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

1 2 3

4 5 6

Figure 14. Sequence of photographs taken with the “tuyerescope” showing an accretion
(image 1) that is dislodged (image 2) and grows again (images 3 to 6) in a period of 3 minutes
at Asarco Tacoma Smelter (Archives of A.A. Bustos; also previously published in
Bustos et al. [1] and Bustos [20]).

Noranda Horne Smelter – Plant Trials (September-October 1985)

Sonic injection plant trials were conducted at the Noranda Horne smelter (now Glencore Xstrata)
in the period between 3 September and 2 October 1985. If the objectives of the Asarco Tacoma
smelter trials were to assess the merit of sonic tuyeres, those of the Horne smelter tests were
more ambitious in terms of quantification. Their goals were to 1) define the minimum pressure
needed for punchless operation as a function of tuyere diameter, oxygen enrichment and matte
grade, and 2) assess the impact of pressure and oxygen enrichment on accretion formation.

Two regular tuyeres of Converter No. 6 were replaced by two sonic tuyeres with internal
diameters of 18.8 and 24.3 mm (0.75 and 1.0 in.), respectively. Two normal tuyeres on each side
of the sonic tuyeres were plugged to avoid any interactions between the high- and low-pressure
tuyeres. The converter was operated normally (practice summarized in Table III) and one of the
sonic tuyeres was tested (the other being plugged) by varying the injection pressure between 207
and 415 kPag (30 to 60 psig). Oxygen enrichment of the sonic tuyere was also tested. Pressure
and flow rate were measured at various points of the gas system and recorded. Visual inspections
were regularly made at the back of the tuyere mounted with the “tuyerescope”.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 281

Table III. Normal Converter Practice at Noranda Horne Smelter in September 1985
Converter No. 6 Details
Converter diameter and length 4.0 m by 9.1 m (13 ft. by 30 ft.)
Number / Size of tuyeres 46 with 48.3 mm I.D. (1.9 in.)
965 mm (38 in.) for slag blows
Tuyere submergence
1 148 mm (44 in.) for finishing blows
Punching machine Gaspé type
Injection pressure 103 kPag (15 psig) for slag and finishing blows
850 Nm3/h/tuyere for slag blows
Injection flow rate
812 Nm3/h/tuyere for finishing blows
Oxygen enrichment (efficiency) None (95%)
Campaign life (No. of charges/campaign) 85 days (170-210 charges)
Copper produced/campaign 18 700 tonnes

Conclusions – Noranda Horne Trials. Experimental difficulties prevented the 19.1 mm tuyere to
be tested for the whole pressure range. However, trials with the 24.3 mm tuyere, which was fully
tested, provided insights into the transition between bubbling and jetting. Three different
injection regimes were delineated by Bustos, Brimacombe, Richards, Vahed, and Pelletier [21]:

1. At injection pressures above 276 kPag (40 psig), the flow rate through the tuyere
remained constant with no pressure fluctuations. The measured flow rate (~ 680 Nm3/h)
were very close to the calculated flow rates for underexpanded flow and a back pressure
of 276 kPag agreed well with the conditions of choked flow at the tuyere tip (sonic flow);
2. At an injection pressure of 241 kPag (35 psig), the flow rate was also constant when
oxygen enrichment was used; with pure air, however, the flow rate sharply declined, the
pressure fluctuated, and at times, the tuyere almost blocked;
3. At an injection pressure of 207 kPag (30 psig), the gas flow rate and tuyere back pressure
were both strongly dependent on the accretion conditions present at the tuyere tip. The
tuyere essentially behaved in the same way as the normal, low-pressure tuyeres.

The trials demonstrated that the accretion formation process is not prevented by high-pressure
blowing and that the degree of tuyere blockage was highly dependent on injection pressure. For
conventional low pressure tuyeres, even a minor accretion build up would significantly reduce
the air flow rate: when more than 50% of the tuyere area was open, a 30% flow rate reduction
was observed. For the high-pressure tuyere, a comparable or larger accretion build up at the tip
did not affect the flow rate or the back pressure. The size and shape of the accretions at the tip of
the high pressure tuyere varied with tuyere diameter and bath composition. Figure 15 provides
examples of pipe-type accretions for two tuyere diameters (18.8 and 24.3 mm I.D.).
Bustos et al. [21] noted that accretion growth and size could not be controlled even with oxygen
enrichment as was the case with the Asarco Tacoma trials. They tentatively attributed this
difference by the lower quality of the flux used at the Horne (67% SiO2 compared to 98% at
Tacoma). According to Casley, Middlin and White [22], a lower silica content in the flux lowers
282 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

the conversion rate, increases magnetite levels in the slag and typically results in higher tuyere
punching frequency (harder and larger accretions).

Pipe-type accretion Pipe-type accretion


(tip of sonic tuyere) (tip of sonic tuyere)

Plug-type accretions
(tip of low pressure tuyeres) Plug-type accretions
(tip of low pressure tuyeres)

(a) (b)
Figure 15. Accretion growth at tip of (a) the 18.8 mm I.D. sonic tuyere at 414 kPag (60 psig),
and (b) the 24.3 mm I.D. sonic tuyere at 276 kPag (40 psig) (Archives of A.A. Bustos)

Inco Copper Cliff Smelter – Plant Trials (November-December 1985)

Lastly, sonic injection plant trials were conducted at the Inco Copper Cliff smelter (now Vale) in
the period between 12 November and 12 December 1985. Their objectives were to complete the
assessment of sonic tuyere injection and a) evaluate the conditions for punchless operation in a
converter for various matte grades, nickel inventory in the converter, and whether differences
could be observed for different stages of blowing cycles, and b) assess the impact of pressure and
oxygen enrichment on accretion formation [21, 23]. Two regular tuyeres of Converter No. 17
were replaced by two sonic tuyeres with internal diameters of 24.3 and 32.5 mm (1.0 and
1.25 in.), respectively. Two normal tuyeres on the side of the sonic tuyeres were plugged to
avoid any interactions between the high- and low-pressure tuyeres. The converter was operated
normally (practice summarized in Table IV).

Table IV. Normal Copper Converter Practice at Inco Copper Cliff Smelter in November 1985
Converter No. 17 (Copper Circuit) Details
Converter diameter and length 4.0 m by 10.7 m (13 ft. by 35 ft.)
Number / Size of tuyeres 48 with 48.3 mm I.D. (1.9 in.)
Tuyere submergence 406 mm (16 in.) for slag and finishing blows
Punching machine Timed mechanical
Injection pressure 103 kPag (15 psig) for slag and finishing blows
590 Nm3/h/tuyere for slag blows
Injection flow rate
720 Nm3/h/tuyere for finishing blows
Oxygen enrichment (efficiency) 28% – occasionally (93%)
Campaign life (No. of charges/campaign) 180 days (200 charges)
Copper produced/campaign 20 000 tonnes
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 283

Due to experimental difficulties, the 24.3 mm sonic tuyere could not be tested. The 32.5 mm
high pressure tuyere was used for the trials with an injection pressure of 276 kPag (40 psig).
Oxygen enrichment of the sonic tuyere was also tested though not extensively over a range of
conditions due to technical issues. Pressure and flow rate were measured at various points of the
gas system and recorded. Visual inspections were also regularly made at the back of the tuyere
mounted with the “tuyerescope”.

Conclusions – Inco Copper Cliff Trials. The following conclusions from the trials were drawn by
Bustos, Brimacombe and Richards [23]:

o The trials with the 32.5 mm high pressure tuyere confirmed the technical feasibility of
punchless blowing in a Peirce-Smith converter;
o At an injection pressure of 276 kPag (40 psig), the 32.5 mm tuyere remained open
throughout the entire converting cycle while the flow rate and back pressure remained
constant. The measured average flow rate – about 1 022 Nm3/h – of the punchless high
pressure tuyere indicated that the productivity of the converter could be increased by 20%
when processing flash furnace matte or 30% when treating semi-blister;
o The nickel content in the converter did not have an effect as important as initially
expected. In fact, charges with 10 tonnes of accumulated nickel behaved the same way as
charges with no nickel at all;
o Over-oxidation of charges, particularly during the end of slagmaking, greatly increased
the accretion formation process and reduced the flow of air through the tuyere, though
did not blocked it.

The trials demonstrated that the accretion formation process is not prevented by high pressure
blowing and that the size and shape of the accretions depended on bath conditions (composition
and temperature) and stage into converting (early or later, slagmaking or coppermaking) as
illustrated on Figure 16.

Plug-type accretions Plug-type accretions


(tip of low pressure tuyeres) (tip of low pressure tuyeres)

Pipe-type accretion Pipe-type accretion


(tip of sonic velocity tuyere) (tip of sonic velocity tuyere)

(a) (b)
Figure 16. Accretion growth at the tip of the 32.5 mm I.D. sonic tuyere operating at 276 kPag
(40 psig): (a) at start of slagmaking blow with pipe-shaped (horn), and (b) at end of slagmaking
blow with larger accretion reducing (not blocking) the air flow (Archives of A.A. Bustos).
284 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

High Intensity Injection in the 90s – Shrouded Injection Developments

After joining Air Liquide in 1989, Bustos applied his experience and expertise in converting and
sonic injection to the development of a technology for high oxygen injection into non-ferrous
bath converting vessels. His work led to the Air Liquide Shrouded Injector – ALSI Technology –
an innovative concept designed to take advantage of the benefits of operating at enrichment
levels above 28% oxygen without increasing the rate of refractory wear [24]. This feat is
achieved by forming a protective accretion at the injector tips (chemical protection) and by
eliminating the need for punching (no mechanical damage) as shown in Figure 17. The shrouded
injector has an inner pipe, through which oxygen enriched air is injected. The inner pipe is
surrounded by an annulus through which nitrogen (or other inert gas or hydrocarbon) flows. Both
the oxygen enriched air and the nitrogen streams are injected at pressures such that the flow
through the inner pipe and the annular space is choked [25].

Hollow-cored
N2 porous accretion
Air + O2

Sight
Window
Refractory Lining
Converter Shell

Figure 17. Shrouded injector schematics and protective accretion growth mechanism.

The first technology plant trials in copper converting took place in 1994 at the Union Minière
Hoboken Smelter in Belgium [25, 26] and the first technology demonstration in nickel
converting took place between 1997 and 1999 [25, 27]. The following sections provide a
summary of the trials and demonstration.

Union Minière Hoboken Smelter – Plant Trials (November 1992-April 1993)

The Union Minière Hoboken operation (now Umicore) is a smelter with a complex Cu-Pb
metallurgy with several secondary as well as precious metals. Back in 1992, the feed composed
of lead, copper and precious metal-bearing concentrates as well as by-products was processed by
sinter-roasting, blast furnace smelting and converting with Hoboken type converters. Union
Minière and Air Liquide ran industrial trials of the high oxygen injection between November
1992 and April 1993 in Converter No. 6 (normal Hoboken converter practice summarized in
Table V). The objectives of the trials were to evaluate the merits of high oxygen injection and
assess the shrouded injector technology.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 285

Table V. Normal Converter Practice at Union Minière Smelter in November 1992


Converter No. 6 Details
Converter diameter and length 3.0 m by 6.0 m (20 ft. by 10 ft.)
Number / Size of tuyeres 18 with 38.0 mm I.D. (1.5 in.)
Tuyere submergence 700 mm (27.5 in.) for slag/finishing blows
Punching machine Manual
Injection pressure 100 kPag (14.5 psig) for slag/finishing blows
Injection flow rate 500-600 Nm3/h/tuyere for slag/finishing blows
Oxygen enrichment (efficiency) 26% (90%)
Campaign life (No. of charges/campaign) 200 days (140 charges)
Copper produced/campaign 7 000 tonnes

The testwork program consisted in three trial campaigns as detailed in Table VI. For the initial
test, only two shrouded injectors were installed in the converter, replacing six conventional
tuyeres (Test 1). The test started on 23 November 1992 and lasted until 29 January 1993. As no
problem were encountered during this initial test, the converter aisle personnel gained confidence
in the technology and decided to proceed with an additional campaign, operating the converter
solely with shrouded injectors. For Test 2, six shrouded injectors replaced the eighteen
conventional tuyeres. Test 2 started on 2 February and ended on 10 February 1993, at which
point the converter was taken out of operation for relining. For Test 3, two additional injectors
were installed, for a total of eight, with the aim of increasing the converter processing capacity.
Test 3 started on 25 March and ended on 28 April 1993.

Table VI. Shrouded Injector Tests at Union Minière Hoboken Smelter (1992-1993)
Test 1 Test 2 Test 3
Number of conventional tuyeres 12 0 0
Number of injectors 2 6 8
Annulus cross-sectional area, mm2 310 151 151
Maximum injection flowrate, Nm3/h/injector 550 550 550
Inner pipe enrichment, % O2 35-60 60 45-60
Overall injector enrichment, % O2 26-50 46-48 30-46

Conclusions – Union Minière Hoboken Trials. Bustos, Cardoen and Janssens [26] drew the
following conclusions from the shrouded injector demonstration campaign at the Union Minière
Hoboken smelter:

• Shrouded injection technology is well applicable and can be successfully implemented in


a Hoboken siphon-type converter;
• The initial objectives for high oxygen injection were met, namely:
o Increase the off-gas strength;
o Increase the treatment capacity for cold materials;
286 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

o Increased the converter productivity by 15%, and


o Raised the end-of-blow temperature from 1050oC to above 1200oC.
• The mushroom-type accretions formed at the tip of the injectors due to the jetting regime
of injection protect the tuyere line;
• The converter operation does not require any punching.

Although the trials were positive and conclusive, ALSI Technology was never implemented
commercially at the Union Minière Hoboken smelter as the process flowsheet was modified in
1994. An Isasmelt top lancing furnace was built and replaced the roasters, sinter plant, blast
furnaces and Hoboken converters that were decommissioned in 1995.

Falconbridge Nickel Smelter – Technology Demonstration (May 1997-January 1999)

The incentives for using oxygen in the converters in the late 90s at the Falconbridge Nickel
Smelter (now Glencore Xstrata) located in Falconbridge, Ontario, were to increase the capacity
for melting cold materials, and to increase productivity. To evaluate the potential for achieving
these objectives without causing excessive tuyere line refractory wear, Falconbridge and Air
Liquide Canada conducted a demonstration of ALSI Technology in a nickel converter starting in
May 1997. Four shrouded injectors were installed and replaced ten conventional tuyeres in
Converter No. 7, a typical Peirce-Smith converter. Table VII summarizes the details of the
converter normal practice at Falconbridge in May 1997 and Figure 18 shows the back of
Converter No. 7 with the shrouded injectors installed.

Table VII. Normal Converter Practice at Falconbridge Smelter in May 1997


Converter No. 7 Details
Converter diameter and length 4.0 m by 9.1 m (30 ft. by 13 ft.)
Number / Size of tuyeres 48 with 50.8 mm I.D. (2.0 in.)
760 mm (30 in.) for slag blows
Tuyere submergence
1 016 mm (40 in.) for finishing blows
Punching machine Kennecott type
Injection pressure 83 kPag (12 psig) for slag and finishing blows
Injection flow rate 640 Nm3/h/tuyere for slag and finishing blows
Oxygen enrichment (efficiency) 21-24% (> 90%)
Campaign life (No. of charges/campaign) 120-150 days (110-165 charges)
Nickel matte produced/campaign 10 000 to 15 000 tonnes

The injectors were designed to operate at oxygen enrichment levels of 30 to 40%. The valve
train, depicted in Figure 19, was designed to allow operating the injectors in pairs. This permitted
comparisons of injector performance at different operating set points. The compressed air
pressure required by the shrouded injectors was supplied by available plant air at 622 kPag
(90 psig). Oxygen and nitrogen requirements were satisfied with vaporized liquid product. The
flows of compressed air, nitrogen, and oxygen to each pair of injectors were regulated, controlled
and measured independently. Process data were recorded using the smelter Foxboro I/A
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 287

distributed control system (DCS). The DCS was linked to a programmable logic controller
(PLC), the valve train field devices, and the converter rotation controls to provide emergency
turn down capability in the event of pressure loss. These control and safety capabilities ensured
the coordinated operation of the shrouded injectors and conventional tuyeres.

Figure 18. Back of Converter No. 7 with four shrouded injectors installed
(one with a sight window). Remaining normal tuyeres each equipped with a
Kennecott-type puncher. (Photographic archives of J.P. Kapusta).

FCV PIT
PRV #1
PIT

#2 To Injector
O2 FCV PIT
Inner Pipe
#3
PIT

FCV #4
PRV

Air FCV

FCV

PRV
PIT
#1
#2 To Injector
N2 FCV
Annulus
PIT #3
#4

(a) (b)
Figure 19. Original gas control valve train for the plant demonstration: (a) design schematics for
independent control of compressed air, nitrogen and oxygen [27], and (b) installation at site
(Photographic archives of J.P. Kapusta).

During the technology demonstration several injector designs were tested to evaluate the effect
of changes in the cross-sectional area and the entrance geometry of the annulus. The first two
tests were short lived due to DCS/PLC communication problems. Once those programming
issues were addressed, the operation of the injectors was simple and did not require any special
care or maintenance, either between blows or between converter cycles. Notably, the shrouded
288 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

injectors did not require punching, as is the case with conventional tuyeres. Figure 20 and
Figure 21 show the typical pressures and flow rates of air, oxygen and nitrogen during a cast.

1400
Nitrogen
Air/Oxygen

1200

1000
Pressure, kPa

800

600

400

200

0
0 1 2 3 4 5 6 7 8 9 10
Elapsed Time, hours

Figure 20. Pressures of oxygen enriched air and nitrogen at the exit of valve train [27].

600
Nitrogen
Air
Oxygen
14
500
12

Flow Rate, Nm 3/min.


400
Flow Rate, scfm

10

300 8

6
200
4
100
2

0 0
0 1 2 3 4 5 6 7 8 9 10
Elapsed Time, hours

Figure 21. Flow rates of air, oxygen and nitrogen through a pair of injectors [27].

The refractory thickness at the tip of the shrouded injectors and conventional tuyeres was
monitored during the entire converter campaigns, refractory wear being defined as the difference
between the refractory thickness at the beginning and end of a campaign. Table VIII presents a
summary of the measurements for refractory wear. The terms “Tuyere/Injector wear ratio
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 289

(brick)” in Table VIII refers to the ratio of the refractory wear at the tuyere and the shrouded
injector tip areas, respectively. This ratio does not consider the fact that the shrouded injector
operated at 30 to 40% oxygen enrichment compared to the conventional tuyeres at 21 to 23%
oxygen. The “Tuyere/Injector wear ratio (oxygen)” takes into account this factor, providing a
measurement of the refractory consumption per unit oxygen injected, which in turn measures the
refractory consumed per unit metal produced. Figure 22 shows an example of protective
accretions formed at the tips of the injectors while Figure 23 offers a visual comparison of
refractory wear difference between the injectors and the conventional tuyeres. This photograph
was taken by Falconbridge personnel after the last test (No. 6) which ended on 1 February 1999.

Table VIII. Shrouded Injector Tests at Falconbridge Smelter (1997-1999)


Test 3a Test 3b Test 4 Test 5 Test 6
Number of casts 38 62 105 99 89
Conventional tuyere enrichment, %O2 21-23 21-23 21-23 21-23 21-23
Overall injector enrichment, %O2 30-40 30-40 30-40 40 40
Annulus cross-sectional area, mm2 263 263 93 93 93
Tuyere/Injector wear ratio (brick) 0.55 0.68 1.07 1.09 1.25
Tuyere/Injector wear ratio (oxygen) 0.88 1.08 1.70 1.98 2.27

Pipe-type accretion
(tip of sonic velocity injectors)

Plug-type accretion Conventional tuyeres Shrouded injector


(tip of low pressure tuyere) 21-23% O2 40% O2

Figure 22. Example of protective accretions Figure 23. Refractory wear pattern at tips of
formed at the tips of shrouded injectors conventional tuyeres and shrouded injectors
showing combined mushroom (from N2) and at end of Test 6 (Photograph from personnel
pipe-type (from Air-O2) accretions [27]. of Falconbridge, February 1999).

The shrouded injector technology demonstration program was successful. The key conclusions
from Bustos, Kapusta, Macnamara and Coffin [27] were as follows:

• The refractory performance of the shrouded injectors, operating at 30-40 % oxygen


enrichment, was superior or equal to the conventional tuyeres at 21-23% oxygen;
• Ensuring that both gas streams are underexpanded and that the nitrogen discharge
pressure is higher than that of the oxygen enriched air were key factors in minimizing
refractory wear;
290 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

• The shrouded injectors operated without the need for punching, unlike the conventional
tuyeres;
• The operation of the shrouded injectors was simple and did not require any special care
or maintenance either between blows or between converter cycles;
• The injector refractory wear pattern is noticeably different than that of the conventional
tuyeres. This is a result of the differences in injection mode and mechanism.
• The successful operation of a shrouded injector requires the correct combination of
injector design and operating parameters. Large changes of injector flow rate or levels of
cooling require design modifications to the injector.

The positive and conclusive results of the demonstration program, as illustrated by Figure 23, led
Falconbridge Limited to implement ALSI Technology on a commercial scale into a Peirce-Smith
type vessel that would become known as the Slag Make Converter (SMC). Some key facts on the
SMC commissioning and operation are provided in the following section.

High Oxygen in the New Millennium – Shrouded Injection Commercial Implementation

From the first trials at Union Minière in 1992 to its first commercial implementation at
Falconbridge in 1999, the high oxygen shrouded injection technology seems to have been on a
seven years fast tracking pace into commercialization. In fact, the road to implementation took
much longer for the technology main protagonists, whether developers or promoters, as this
historical perspective attempted to demonstrate. Unfortunately, J. Keith Brimacombe who
spearheaded research on gas injection phenomena in converting did not live to see sonic injection
practiced in operation. For Alejandro A. Bustos, who contributed immensely to the wealth of
knowledge on injection phenomena, this first successful commercial implementation of shrouded
injection was a dream come true. For the present writer, who started learning about gas injection
while at UBC under the supervision of Professor Greg G. Richards and was then mentored by
Alejandro Bustos at Air Liquide Canada, this was a dream career start. My own ‘dream come
true’ came with the opportunity, this time without my mentors, to design and commission
shrouded injection at the Thai Copper Industries smelter in 2005-2006. The following section
provides a summary of shrouded injection implementation in nickel converting at Falconbridge
(SMC vessel) and in copper converting at Thai Copper Industries (Hoboken converters).

Falconbridge Nickel Smelter – SMC Commissioning and Operation (May 1999-January 2000)

The last test of the shrouded injector demonstration ended in early February 1999 and just three
months later, by early May 1999, the SMC vessel was installed in the converter aisle of the
Falconbridge smelter, replacing the old Converter No. 7. The SMC was designed as an extended
Peirce-Smith type converter with a 4-meter diameter by 17-meter long (13 ft. by 55 ft.). The
purpose of the SMC was to oxidize most of the iron and associated sulfur in the metallized
furnace matte (green blows) while the remaining conventional Peirce-Smith converters became
finishing vessels. A front view of the SMC vessel is shown on Figure 24.

The original design criteria of the SMC called for a total blast rate of 9 775 Nm3/h (6 200 scfm)
with an overall oxygen enrichment of 30%. This conservative approach was chosen to take full
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 291

advantage of the experience and lessons acquired during the technology demonstration. The
decision was therefore taken to design the first shrouded injectors for the SMC with dimensions
very similar to the injectors of the last test campaign, which run from December 1998 to
February 1999 (Test 6). The SMC was then equipped with 13 shrouded injectors with a
2.5 cm O.D. (1.0 in.) and designed for a flow rate of 752 Nm3/h/injector (477 scfm/injector).
Figure 25 (a) shows the back of the SMC with the injectors in blowing position. Note that no
provision for punching equipment was made since the technology demonstration had proven that
the injectors, when blowing under choked conditions, operated in a punchless mode. The
injectors were designed with a more conservative cooling factor of 50-55% rather than the lower
value of 30-35% used during the last demonstration test.

Figure 24. Front view of SMC – an extended (4 m by 17 m) Peirce-Smith


type converter (Photographic archives of J.P. Kapusta, 1999).

(a) (b)
Figure 25. Commissioning installation: (a) back of SMC vessel with shrouded injectors in
blowing position, and (b) schematics of injector radial assembly on SMC shell [1].
292 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

Falconbridge opted to purchase an air compressor dedicated to the SMC and mandated Air
Liquide Canada to design the gas train for mixing pure oxygen and compressed air. The nominal
flow rates for the oxygen enriched air and nitrogen streams were initially 8 240 Nm3/h and
1 537 Nm3/h (5 225 scfm and 975 scfm), respectively, while their nominal pressures were
520 kPag and 620 kPag (75 psig and 90 psig), respectively. The gas supply system was capable
of providing an oxygen enriched air pressure of up to 690 kPag (100 psig) and a nitrogen
pressure of up to 1 030 kPag (150 psig). Finally, the shrouded injectors were mounted on the
SMC shell in a radial position as schematically represented in Figure 25 (b).

SMC Commissioning – May 1999. Commissioning of the SMC vessel started on 5 May 1999
and was not without hiccups. Again, strong with the experience gained and the lessons learned
during the technology demonstration, Falconbridge and Air Liquide personnel quickly identified
the causes of the trouble and implemented corrective actions to resolve the issues at hand. Details
have been published by Kapusta, Stickling and Tai [1]. The first blow of the SMC occurred on
14 May 1999, and by 21 May 1999, a decision was made to reduce the number of active injectors
from 13 down to 10 by plugging with ram mix the injectors at both extremities of the vessel (one
directly under the feed and dip rod ports on one side, and a second under the burner port on the
other side), and one injector in the centre of the vessel directly across from the mouth and spout.
A fourth injector was plugged on May 23. The flow set points for the remaining 9 injectors were
subsequently adjusted. This action dramatically improved splashing and eliminated plugging of
the spout and ports. The shrouded injectors were continuously monitored even though they
operated under their original process design criteria. As expected – and as needed – they were
also performing in punchless mode blowing a 30% oxygen enriched air into a highly metallized
matte from the electric furnace at a constant flow rate throughout the blows. Considering that the
SMC vessel equipped with ALSI Technology was a new process for both operators and
engineers at Falconbridge, a program to inspect the injectors and auxiliary equipment was put in
place by the commissioning team. The purpose was to ensure an adequate monitoring of the
performance of the SMC, particularly in terms of refractory wear at the tuyere line. A routine
was developed and a procedure established to perform regular refractory thickness measurements
through the inner pipes of the injectors.

Shrouded Injector Design Modifications – August 1999. Within 3 months of commissioning, all
stakeholders in the SMC operations had become sufficiently comfortable with the new vessel
operation and decided to increase the oxygen enrichment level from 30% to 40%. Operating the
SMC at 40% oxygen with nine injectors (down from 13 at commissioning) required a new, larger
injector design. By 15 August 1999, injectors with a 3.2 cm O.D. (1.25 in.) were installed in the
SMC with a flow rate capacity of 1 072 Nm3/h/injector (680 scfm). Their annular space
dimension remained unchanged to maintain a cooling factor of 50-55%.

Shrouded Injector Design Modifications – January 2000. Eight months after commissioning,
operating the SMC had become part of the work routine in the Falconbridge converter aisle. As
confidence was gained by successfully operating ALSI Technology, further modifications were
brought to the injector design. Most notably, the annular gap size was reduced to lower the
cooling factor down to the 30-35% level that was achieved during the last demonstration
campaign of December to February 1999. The overall oxygen enrichment was however
maintained at 40%. This achievement marked a major milestone in the commercial
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 293

implementation of ALSI Technology considering that a full size converter vessel was capable,
within 8 months of its commissioning, of converting highly metallized matte at the highest
oxygen enrichment level ever achieved at Falconbridge with the lowest cooling factor to date
without any punching. The myth that contrary to steelmaking, high oxygen shrouded injection
was not technically feasible in a nonferrous metal converter was finally proven unfounded.

Xstrata Nickel Smelter – SMC Operation and Optimization (2000-2009)

Within the decade, the SMC vessel saw further modifications to improve and optimize its
operation. The decade also saw the acquisition of Falconbridge by Noranda (keeping the name
Falconbridge) and then Falconbridge by Xstrata. Salt and Cerilli have presented updates of the
converter aisle operation at the Xstrata Nickel smelter [29-30]. Some of the key developments
discussed were as follows:

• May 2002 – rotation of the injector mounting assembly by 34 degrees requiring a slightly
longer injector. The benefits were a 15-cm (6 in.) increase of the injector submergence,
an increase in stirring of the bath, and a further shifting away from the refractory bricks
of the contact between the oxygen enriched air and the molten matte. The goal was to
lower refractory wear immediately above the injectors. Figure 26 shows the tear-shaped
refractory wear pattern above the injectors described by Bustos et al.[27] that would be
reduced with deeper submergence;
• May 2002 – Reduction of the number of active injectors down to 6 with a lowering of the
total blast rate down to about 6 450 Nm3/h (4 100 scfm) and an increase in oxygen
enrichment to 43%;
• 2005 – Replacement of hood and back flap resulting in a large reduction of fugitive
emissions from the vessel;
• June 2006 – replacement of the Garr gun with a central feeding chute, using a modified
shrouded injector (inner pipe replaced by a blank) that injected only nitrogen to provide
stirring of the bath under the chute. This major change allowed an increase of cold charge
feeding of 45%;

Figure 26. Refractory wear pattern above shrouded injector


(Photographic archives of J.P. Kapusta, 1999).
294 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

Thai Copper Industries – Shrouded Injection Commissioning and Operation (2005-2006)

The Thai Copper Industries smelter (TCI) is located about 200 km southeast of Bangkok near the
Map Ta Phut deep sea port in one of the industrial centres of Thailand. The project began in
September 1995 with the signing of an Engineering, Procurement and Construction Management
contract (EPCM) with Sofresid S.A. and Davy International. Construction of the smelting and
refining complex started in late 1996 but was suspended in February 1998 due to the Asian
economic crisis of 1997 and the subsequent devaluation of the Thai currency, the baht. By that
time, construction was already at an advanced stage: most of the concrete, structural steel work
and buildings were completed, and major pieces of equipment were installed, the remaining
being stored on site. Construction resumed in January 2003 after TCI awarded a new EPCM
contract to Aker Kvaerner. The smelter, acid plant and electro-refinery were completed and
handed over to its owner by Aker Kvaerner in May 2004. The smelter started production in July
2004 and operated for a year until Management of TCI, following an incident at the smelter,
interrupted all production and requested a full review of all operating and safety practices. Pre-
commissioning was re-initiated in August 2006 when teams of engineers and operators from the
Philippines and from INDEC and CODELCO of Chile were contracted. Hot commissioning of
the smelter took place in December 2006. The following section presents a summary of the
Hoboken-type converters and shrouded injectors commissioning.

Thai Copper Industries Smelter. TCI was designed to process imported concentrates with a
yearly production capacity of 165 000 tonnes of cathodes. Concentrates were blended and
smelted continuously in a primary smelting reactor – El Teniente Technology – to produce high
grade matte or white metal. The white metal was then converted in batches into blister copper in
Hoboken siphon converters – two hot and one on stand-by – all equipped with ALSI Technology
for sonic injection of oxygen enriched air. Final refining was carried out in two anode furnaces
prior to anode casting and electrorefining. Complete details of the flowsheet and equipment have
previously been described by Kapusta, Wachgama and Pagador [31-32]. The three Hoboken
converters were 3.8 m in diameter by 8.5 m long. Figure 27 shows one of the Hoboken
converters installed at TCI.

(a) (b)
Figure 27. One of the 3 Hoboken converters at TCI: (a) March 2004 (before commissioning)
with siphon in forefront, and (b) December 2006 (after commissioning) with view of small-sized
converter mouth and its lid [31-32].
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 295

Injector Design. The original decision to implement shrouded injection at TCI dates back to
1996. That year, Union Minière Engineering (UME) had just been awarded a contract to supply
its Hoboken siphon converter technology to TCI. Two years earlier, Union Minière and Air
Liquide had successfully completed a plant demonstration of shrouded injection at its Hoboken
smelter in Olen, Belgium [26]. In 2003, after reviewing and modifying the converter operating
parameters as part of its ECPM contract, Aker Kvaerner requested that Air Liquide Canada
(ALC) design a new set of injectors reflecting the changes in operation of the converters and
incorporating the improvements to the design of shrouded injectors that had been undertaken by
Air Liquide Canada between 1997 and 2003. These new injectors, by then the fourth generation,
were designed and fabricated based on the operating parameters and design criteria supplied by
Aker Kvaerner and presented in Table IX. The new injectors were designed for a cooling factor
ranging between 35 and 45% with an O.D. of 48.3 mm (nominal 1.5 in. – 160S).

Table IX. Shrouded Injector Design Criteria for TCI


Nominal Maximum
Number of Injectors per Converter 28
3
25 590 Nm /h 37 050 Nm3/h
Total Blowing Rate per Converter
(16 225 scfm) (23 490 scfm)
3
915 Nm /h 1 325 Nm3/h
Total Blowing Rate per Injector
(580 scfm) (840 scfm)
Overall Oxygen Enrichment 30.4% 27.8%
Injector Submergence 0.75 m (30 in.)
Bath Temperature 1250oC (2282oF)

Setting the oxygen enrichment at 30% as a design criterion for the injectors was based on a
desire to be on the conservative side and ‘minimize’ the process risk. After all, this was to be the
first commercial implementation of shrouded injection in a copper smelter. Strong with my
experience at Falconbridge of shrouded injection for both the technology demonstration in the
old Converter No. 7 and its implementation in the SMC, I believed that 30% oxygen was too low
for a technology meant to be for intense, high oxygen injection. The wisdom of lead process
engineer Richard McClincy of Aker Kvaerner prevailed. Richard had just reminded me of the
valuable engineering lesson that when considering oxygen enrichment in pyrometallurgy,
designing for the lesser as a first step is the mark of an experienced metallurgist who knows that
once he has successfully commissioned a vessel or a plant, he can modify the design for higher
enrichment levels. Was not this the approach Brian Macnamara, lead process engineer at
Falconbridge, used for the SMC vessel too? At 30.4% oxygen enrichment and about
25 600 Nm3/h total blowing rate per converter, 28 shrouded injectors were needed. Figure 28 (a)
shows the back of a Hoboken converter in stand-by position with the injectors installed with a
close-up on the injectors in Figure 28 (b).

If the enrichment level risk was kept low, a number of questions remained unanswered. How will
the injectors behave when so closely interspaced (165 mm) in the tuyere line? Will they interfere
with each other? Would that be detrimental to the formation of protective accretions?
Commissioning promised to be eventful if not challenging.
296 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

(a) (b)
Figure 28. One of the three Hoboken converters at TCI: (a) back view with all 28 shrouded
injectors installed [31], and (b) close-up on a few injectors showing the ball valve at the back of
the injectors allowing visual inspection (Photographic archives of J.P. Kapusta).

Converter Aisle Commissioning. Details of the pre-commissioning events in November 2006


and until 10 December 2006 were previously published by Kapusta, Wachgama and
Pagador [31]. These activities included establishing a proper converter practice (DCS/PLC two-
level permission access: ‘design level control’ versus ‘operation level control’), ensuring a
proper gas delivery and flow rate control (correction of flow rate equations, optimisation of
compressed air system, and improvements of loop control responses and PID control settings),
reviewing and correcting alarms and interlocks, and providing training to engineers and
operators on ALSI Technology.

In the evening of 12 December 2006, white metal was tapped from the primary smelting reactor
(El Teniente) and transferred to Hoboken Converter No. 2, initiating the hot commissioning of
the converters with ALSI Technology. In spite of several emergency roll out due to failures of
the acid plant gas cooling system, the converting cycle proceeded without any problem.
Unforeseen requirements for additional compressed air in the primary smelting section forced the
start-up of all three compressors available and the sharing of compressed air between smelting
and converting. This temporary situation resulted in instabilities in air supply to Converter No. 2
since fine tuning of the PID control system for the compressed air supply had been established
during pre-commissioning for two compressors dedicated to the converters. Figure 29 presents
the flow rates of air, oxygen, and nitrogen during the various blows of the first cycle and shows
the slowness of the air flow rate to respond to set-point changes (step changes in response to
operators input are shown on the graph).

Photographs of accretions at the tip of the injectors were taken through the mouth after each roll
out all along the converting cycle. As seen on Figure 30, the size and shape of the accretions are
dynamic parameters that depend on the timing within a blow (chemistry of the bath) and on the
oxygen enrichment level used: larger accretions were obtained during slag blows or for low
oxygen enrichment levels while smaller accretions were observed during copper blows or for
high oxygen levels.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 297

500

Air
Gas Flow Rates, Nm3/min. 400

300

200

N2

100
O2

0
0 30 60 90 120 150 180 210 240 270
Elapsed Time, minutes

Figure 29. Oxygen, nitrogen, and compressed air flow rates through the shrouded
injectors during the first charge of the converting campaign [31].

(a) (b)

(c) (d)
Figure 30. Changes in accretion shape and size during a blowing cycle and at various oxygen
enrichment levels: (a) slag blow and high O2, (b) early copper blow and high O2, (c) mid copper
blow and low O2, and (d) end of converting cycle [31].
298 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

Learning during Commissioning and Operation. TCI operators and engineers learned very
quickly how to convert white metal in a Hoboken converter with ALSI Technology and
developed their own best practices. Unlike conventional tuyeres, the shrouded injectors operate
without the need for punching. However, early on in the first converter campaign and when
blowing under high magnetite conditions, operators developed the practice of checking the
injectors by manually rimming a steel bar through their inner pipes. Although not necessary, this
practice gave operators some additional confidence prior to the start of a new cycle. No special
care or maintenance, either between blows or between converter cycles, is normally required.
When operating with level of oxygen enrichments above 27%, maintaining control of the heat
balance in the converter, and hence of the converting reaction kinetics, is of outmost importance.
This heat balance control was achieved at TCI by cold dope additions along the converting cycle.
When the aisle crane encountered delays, regular and speedy charging of reverts or scrap was not
possible, requiring operators to reduce the oxygen enrichment level. During such instances, the
protective accretions grew much larger resulting in an increase of the injection back pressure as
the control system reacted to maintain the gas flow rates.

In the short operating span of the TCI smelter, important operational changes were made from
the original installation. Splashing through the mouth during roll in was systematically observed.
The remedy to this operational nuisance was to remove the injectors and close the tuyere sites
directly opposite the converter mouth. Five injectors were first removed (see Figure 31) and one
on each side of the tuyere line thereby reducing the total number of injectors from 28 down
to 21. The injector design specifications, most particularly the injector blowing rate, were left
unchanged. These simple modifications drastically reduced the splashing through the mouth,
considerably lowering the potential for mouth erosion. Another benefit of that change: with
only 21 injectors installed, the converters achieved better operational results as the injectors
could be operated at higher oxygen levels giving a better control of the accretion size. TCI was
taking the same path Falconbridge did years earlier when commissioning its SMC vessel:
starting conservatively with a higher number of injectors at a lower oxygen enrichment level and
moving to lower number of injectors at higher oxygen enrichment levels.

Closed tuyere sites

Figure 31. Back view of Hoboken converter with injectors installed


and five tuyere sites closed in front of the mouth [32].
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 299

In the months following commissioning, TCI personnel became quickly acquainted with the
operation of shrouded injection. In particular, they learned to ‘read’ the meaning of variations in
back pressures during the slag blows and finishing blows and to adjust the oxygen enrichment as
needed. They also experimented with oxygen enrichment as high as 38% to test the capacity of
the converters to treat reverts. Figure 32 shows the total amounts of reverts consumed by each of
the three reactors at TCI. PSR stands for primary smelting reactor (El Teniente), ESCF for
electric slag cleaning furnace, and HSC for Hoboken siphon converters. Within a few months of
operation, the Hoboken converters demonstrated a high capacity for reverts recycling by fully
taking advantage of the shrouded injector capabilities.

3 500
3 164
PSR
3 000 ESCF
HSC
2 500 2 350
Metric Tonnes

2 126
2 000 1 893

1 500
1 100
1 000 862 813 791

500
278

-
Dec-06 Jan-07 Feb-07

Figure 32. Consumption of reverts per vessel at TCI between


December 2006 and February 2007 showing high reprocessing rates for
the Hoboken converters within two months of commissioning [32].

Another advantage of the unique combination of the Hoboken siphon converter and ALSI
Technology is the prevention of accretion buildup, whether on the mouth, the endplate or the
areas below the siphon. In typical Hoboken converter operations, accretion builds up
significantly, in particular below the siphon, and requires long downtime and man-hours to
unclog the buildup that would otherwise restrict the exhaust gas flow. This cumbersome problem
has however not been experienced with the Hoboken siphon converters at TCI. Figure 33
presents a series of photographs of the mouth, endplates and siphon from one of the converters
after cool-down. The mouth is still in good shape with the tuyere line visible inside. The
endplates are also clear of any build up or accretion. The siphon, dismantled during a routine
maintenance repair, also presents minimal build up around the refractory lining.

Conclusions. The Thai Copper Industries smelter accomplished a world premiere in


pyrometallurgy in December 2006 when its three Hoboken siphon converters were
commissioned to operate solely with shrouded injectors discharging high oxygen enriched air at
sonic velocity, consistently, and without the need for punching machines. With no history of
copper production in Thailand, commissioning a converter aisle equipped with a new injection
technology is an achievement well worth commending, especially since this was accomplished
all the while the commissioning and start-up of the smelter as a whole, including the El Teniente
300 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

primary smelting reactor, electric slag cleaning furnace, converters, anode furnaces and acid
plants, was taking place. This success certainly demonstrated the capabilities of the Thai Copper
converter aisle crew as well as the leaderships of Messrs. Noparut Wachgama, the converter aisle
manager, and Romeo Pagador, the smelter manager. Thanks to Noparut and Romeo I recall my
different trips to TCI very fondly, especially the five weeks I spent during the late 2006
commissioning. The human and technical skills of Noparut impressed me profoundly.

TCI had hired a hundred-strong team from Chile to ensure a successful commissioning of the
El Teniente reactor and to provide operational support to the anode furnaces, casting wheels and
acid plant sections. TCI had also hired a crew from the Philippines experienced in copper making
who would support the Thai operators, especially in the converter aisle. I recall being questioned
repeatedly by the Chilean engineers about the viability of the shrouded injectors. I recall some
were certain punchless operation was not feasible. Noparut believed and trusted we could
accomplish what I was claiming about shrouded injection, even at times when the tasks were
overwhelming, the day shifts were regularly extending into night shifts, and I started doubting of
our ability to have the converters ready when the Teniente reactor would be. I had experienced
leadership in Canada from people more senior than me, during the Falconbridge trials and
implementation for instance, I was now humbled and pumped up by a local leader younger than
me. Romeo was exemplary in supporting his Thai crew in his plant manager role. The success of
the commissioning of the converter at Thai copper is Noparut’s and Romeo’s success. Being part
of that human and technical experience is one of the most memorable moments of my career. I
am still very saddened that TCI had to mothball operations in the spring of 2007 under economic
duress and limited copper concentrate availability. The opportunity to work with the Thai Copper
crew to optimize converter operation vanished and I never returned to the smelter … so far.

Figure 33. Photographs of a Hoboken converter mouth (top left), endplates (top right) and siphon
(bottom) after cool down showing the lack of buildup typical to Hoboken converters not
equipped with ALSI Technology [32].
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 301

Closing the Loop – Full Scale Single-Pipe Sonic Tuyere Injection at Lonmin Platinum

Although sonic injection trials were conducted in the 80s, as presented above, these ‘only’
proved the concept of punchless injection under sonic conditions and provided some facts about
sonic tuyere operation. Without running at least a full converting campaign, quantifying the pros
and cons of sonic injection was left to the realms of speculations. Pyrometallurgists, contrary to
investors, tend to be conservative speculators and to assume often undue constraints to new
technologies. After all, preserving the integrity of their high temperature vessels has served them
well throughout the decades. A converting campaign with a converter solely equipped with sonic
tuyeres would at last provide the necessary metrics to truly evaluate the technology, and
hopefully, dispel some if not all the myths about it.

As part of a study to improve process efficiency and increase capacity, Lonmin Platinum Ltd
(Lonmin) assessed various options for its converting operation. After reviewing these options in
2008 and 2009, Lonmin determined that sonic injection had the most promising potential to
address its future converting capacity needs. By 2010, Lonmin decided to conduct a full scale
industrial evaluation of sonic injection, using air only, to establish if punchless, high efficiency
sonic injection was feasible and beneficial in its small converters. Plant trials were planned in the
first half of 2011 and run for three weeks in the fall of 2011 in one of Lonmin’s three Peirce-
Smith type converters. Converter No. 2 was selected for the trials, solely equipped with sonic
tuyeres, and operated by injecting compressed air from a series of rented compressors. The
following section presents the highlights of the plant trials.

Lonmin Smelting and Converting Processes

Lonmin is the world’s third largest primary platinum producer with the bulk of its operations
concentrated within the companies of Western Platinum Ltd and Eastern Platinum Ltd, both
located close to Marikana in the North West province of South Africa. The Precious Metals
Refinery (PMR) is situated in the Springs district of Johannesburg.

Lonmin operates eight small concentrators that are optimized for the ore type close to the shaft
head and that produce concentrate slurries delivered to the smelter complex by tanker. The
smelter receives concentrate into a blending section that aims to keep both the Cu + Ni and Cr
levels within a narrow band. Blended slurry is dewatered in a filter press to produce a filter cake
of about 15% moisture, which is then dried in a flash dryer down to about 0.5% moisture. Dried
material is stored in silos from where it is fed to the five electric (AC) smelting furnaces. Furnace
slag is granulated directly from the tapping launder and sent to the slag mill-and-float circuit.
Matte is tapped into refractory lined ladles and transported to the converter aisle operated with
three Peirce-Smith converters with shell dimensions of 3.05 m by 4.57 m (10 ft. by 15 ft.). Two
are typically operating (hot) while the third one is being re-lined or on standby (cold). Furnace
matte is added in batches (ladles) to the converters and air is blown to oxidize iron sulphides.
Once an end-point with an iron content of approximately 0.5% to 1.5% is reached, the matte is
tapped, granulated and sent to the Base Metals Refinery (BMR) while the granulated converter
slag is sent to the mill-and-float circuit. Off-gases from the furnaces pass through a three-field
electrostatic precipitator (ESP) to remove particulates. The dust is returned to the furnace via
pneumatic conveying. The off-gases from the converters and furnaces are mixed after the ESP
302 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

and pass through a variable throat scrubber for additional particulate removal and gas quenching.
Recovered particulates are returned to the blending section in the form of slurry. The gas from
the variable throat scrubber passes through a concentrated mode Dual Alkali scrubber where
calcium sulphite is produced and thereafter sent to land-fill. More details on the Lonmin smelter
and operations have previously been published by Eksteen, Van Beek, and Bezuidenhout [33]
and by Kapusta, Davis, Bezuidenhout, Lefume and Chibwe [34].

Lonmin Rationale for Evaluating Sonic Injection

The Lonmin smelting and refining circuits were developed to process high ratios of UG2
concentrate rich in chrome. This feed characteristic has two major negative impacts on the
converting operations that have to cope with 1) high levels of dissolved Cr reporting to the
furnace matte (between 2% and 2.5%, for chrome expressed as Cr2O3), and 2) high furnace matte
temperatures (between 1450 and 1580oC).

High chrome levels in the furnace matte lead to severe chromite (FeCr2O4) precipitation during
the oxidising converting process. The precipitates accumulate as accretions around tuyeres and
can lead to more frequent plugging of the tuyeres. Consequently, more-than-usual punching of
the tuyeres is required, contributing to a more rapid deterioration of the tuyere line. Blocked
tuyeres can also rapidly lead to a misdistribution of injected air, leading to local over-oxidation
in some parts of the bath whilst other parts are poorly converted. The resulting localised blocking
of tuyeres further exacerbates the poor longitudinal mixing and changes the wave formation in
the converter, which typically results in increased amounts of slopping and splashing
experienced from the converter mouth. The high matte temperature is another challenge from the
stand-point of a converter refractory campaign life; unless the converter is pre-heated to quite
high temperatures, the refractory will spall fairly rapidly from thermal cycling when it is loaded
with furnace matte at temperatures around 1 500°C. Already weakened by frequent punching, the
tuyere line area is further compromised by the thermal cycling. Converter campaign life at
Lonmin can be as low as 20 cycles (aka “blows” in the Lonmin terminology) and is seldom more
than 40 cycles.

Lonmin decided to investigate sonic injection as an option to circumvent some of the problems
experienced by the current operation. The goal was to achieve a stable and predictable air flow
into the converter through sonic velocity injection and generate protective accretions that do not
grow to close off the tuyere exit channels. By eliminating the need for punching, refractory life
would increase, operational stability could be achieved (predictable air input and mixing in the
converter) and the cost of punching (punch bars, punch cars and punch car operators) could be
avoided. In 2010, the decision was then taken to run a full scale sonic injection test on a
converter for a period of two to three weeks to establish the operational advantages that could be
achieved with sonic injection and to complete the commercial viability of the project. Only a test
with a converter solely equipped with sonic tuyeres replacing all conventional tuyeres was seen
as a conclusive approach to evaluate blowing characteristics, including slopping, splashing,
refractory wear, accretion formation, and air flow stability. Due to careful planning, safe design
of equipment, and availability of rented compressors, the trials did not take place until late 2011.
A summary of the results of the full scale plant trials are described in the following sections.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 303

Converter No. 2 was selected and relined for the sonic injection plant trials. The Gaspé type
punching machine was removed and the sonic tuyeres installed. Figure 34 shows
Converter No. 2 during normal operation (with its punching car) and ready for the sonic injection
trials (with eight tuyeres and no punch car). Lonmin rented five compressors and manufactured a
multi-inlet pressure rated receiver (see Figure 35). During the cold and preliminary hot
commissioning period between 7 November and 23 November 2011, a number of issues with
compressed air supply and flow measurement were encountered leading to variations in flow rate
and pressure in the sonic tuyeres. These issues were addressed by letting the sonic tuyere control
the air flow and pressure into the converter.

(a) (b)
Figure 34. Converter No. 2 at Lonmin: (a) equipped with Gaspé punching machine during
conventional, low pressure operation, and (b) equipped with 8 sonic tuyeres (replacing 18
conventional tuyeres) and without punching machine [34].

(a) (b)
Figure 35. Key equipment for the trials: (a) five rented compressors connected with flex hoses,
and (b) high pressure air receiver [34].

By 24 November 2011, Converter No. 2 was blown with seven sonic tuyeres, all five
compressors in cascade mode and a fully open flow control valve (FCV). The next day an
additional sonic tuyere was installed. Figure 36 shows the recorded injection flow rate and
pressure during the cycle. The flow rate curve shows a much improved stability compared with
blows with conventional low pressure tuyeres (see Figure 37). Even more significant is the
stability of the sonic injection pressure. The stability of both the flow rate and pressure of the
304 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

compressed air demonstrated that the new operating strategy was successful. Blowing the five
cascaded compressors with a fully open FCV proved that the sonic tuyeres were self-regulating.
This was one of the major milestones achieved in the trials. The injection pressure averaged
about 310 kPag for a stable average flow rate of 9 100 Nm3/h. By 28 November 2011, six
charges (cycles) in Converter No. 2 proved to be successful in achieving the objectives set for
the technology demonstration – punchless, highly efficient converting process with acceptable
splashing in all successive blows of a complete charge.

Blow 110 (Sonic) - November 25, 2011


12,000 600
Compressed Air Flow Rate, Nm3/h

Flow rate

Tuyere Back Pressure, kPag


10,000 500

8,000 400

6,000 300

Pressure
4,000 200

2,000 100

0 0
11:16:48 12:28:48 13:40:48 14:52:48 16:04:48 17:16:48 18:28:48 19:40:48 20:52:48 22:04:48
Flow Pressure Flux Cold dope

Figure 36. Stable flow rate and pressure of 8 sonic tuyeres during Blow 110 on
25 November 2011 in Converter No. 2 [34].

Conventional (Subsonic) Blow 175 - December 20, 2011


16,000 160
Compressed Air Flow Rate, Nm3/h

Flow rate
14,000 140

Tuyere Back Pressure, kPag


12,000 120

10,000 100

8,000 80

6,000 60

4,000 40
Pressure
2,000 20

0 0
11/12/20 17:16:48 11/12/20 19:40:48 11/12/20 22:04:48 11/12/21 00:28:48 11/12/21 02:52:48 11/12/21 05:16:48

Flow Pressure Flux Cold dope

Figure 37. Unstable flow rate and pressure of 18 subsonic (conventional) low pressure tuyeres
during Blow 175 on 20 December 2011 in Converter No. 2 [34].

Starting on 28 November 2011, Converter No. 2 was operated in production mode with two
shifts per day (i.e. two converting charges per day) in sonic injection regime until the end of the
compressor rental agreement expiring on 4 December 2011. The goal of this one week intense
production period, noticeably without my assistance, was to demonstrate the ease of operation of
the converter under sonic conditions, to allow foremen and skimmers to become better
familiarized with sonic injection, and also to provide additional information on converter
operations such as refractory wear rate, reverts reprocessing capacity, blow length reduction.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 305

Conclusions – Lonmin Full Scale Trials (November-December 2011)

The sonic injection trials in one of the relatively small converters at Lonmin were successful in
achieving their objectives within the combined 2½ weeks of hot commissioning and full
production. Some highlights are summarized below.

Accretion Formation and Refractory Wear. The formation of protective accretions at the tip of
the sonic tuyeres was continuously monitored visually and photographically. Their dynamic
nature is illustrated on Figure 38 by their changing shape and size at the tip of four of the eight
tuyeres after each blow of Charge 110 on 25 November 2011. These photographs demonstrate
the impact of bath chemistry, temperature and motion on the protective accretion growth process.

Ladles 1 and 2 (end of first blow) Ladle 3 (end of second blow)

Ladle 4 (end of third blow) Ladle 5 (end of fourth blow)

Ladle 6 (end of fifth blow) Finishing Blow (end of sixth blow)


Figure 38. Dynamic accretion formation and changes in shape and size during the blows
of charge 110 (25 November 2011) in Converter No. 2 [34].
306 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

The positive impact of protective accretions formed under sonic flow was measured during the
trials. At Lonmin, a converter campaign starts with a refractory thickness at the tuyere line of
660 mm (full reline) and typically ends when the refractory thickness is down to about 250 mm.
Over the years, conventional converting campaigns have averaged 28 cycles between relines,
corresponding to an average refractory wear per cycle of 14.8 mm. For the seventeen sonic
charges of the plant trials, the maximum refractory wear rate ranged between 10.3 and 11.1 mm
per cycle, corresponding to 37-40 cycles per campaign, or a 34% reduction in refractory wear
with sonic injection compared with conventional subsonic conventional injection. The above
measurements of refractory wear, even if done over a short period of time or a short number of
blows, still provides an industrial validation of the theory that the accretions formed during sonic
injection are indeed protective rather than disruptive.

Slopping and Splashing. One of the key challenges of the plant trials was to demonstrate that
slopping and splashing could be controlled under sonic injection, or better yet reduced, compared
to bubbling injection, in spite of the smaller size and larger mouth characteristic of the converters
in operation at Lonmin. The early hot commissioning blows produced excessive splashing out of
the converter mouth, that even if sporadic was fairly large in volume, generating a lot of concern
with the commissioning team. However, once the compressed air supply issues were addressed
and the sonic tuyeres were permitted to control the flow and pressure into the converter,
splashing reduced considerably, first to levels comparable, and then to lower levels than
Converters No. 1 and No. 3 operating under bubbling regime. Controlled splashing was also
accompanied by a stable flow rate and pressure of compressed air (Figure 37).

Converter Efficiency and Productivity. The trials achieved another milestone in quantifying the
increase in converter efficiency and productivity under sonic injection mode compared with
conventional bubbling mode. An initial analysis of the first week of hot commissioning with 7 or
8 sonic tuyeres indicated that Converter No. 2 took an average of 388 minutes of in-stack time to
complete a cycle with an average of 8 850 Nm3/h of compressed air compared to 469 minutes for
Converters No. 1 and No. 3 for the month of November 2011 with an average blowing rate of
about 11 000 Nm3/h. These values corresponded to an average of 7.7 minutes blown per tonne of
matte charged for sonic injection compared to an average of 9.6 minutes blown per tonne for
conventional blowing, equivalent to a reduction of about 20% in blowing time for the sonic
injection cycles even at a lower blowing rate. Converter No. 2 also consumed an average of
8.8 tonnes of reverts per cycle compared to 3.0 tonnes for Converters No. 1 and No. 3 for the
month of November 2011, an increase of close to 200%. Those comparative results are
summarized in Table X.

Table X. Conventional (Bubbling) Versus Sonic (Jetting) Injection Regime


Conventional (No. 1 & 3) Sonic Trials (No. 2)
Number of tuyeres 18 7 or 8
Average in-stack time 469 minutes 388 minutes
Average air blowing rate 11 000 Nm3/h 8 850 Nm3/h
Average blowing time 9.6 min./tonne of matte 7.7 min./tonne of matte
Average reverts consumption 3.0 tonnes/cycle 8.8 tonnes/cycle
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 307

The Next Milestone. After analysing the results of the full-scale plant trials and completing a
commercial evaluation for changing its converting operation over to sonic injection on all three
converters, the Lonmin smelter personnel applied for the capital of a permanent installation that
would be justified based on operational cost difference alone (in particular reverts consumption,
refractory wear, cost of punching, etc.). The last milestone – commercial implementation – to
complete the research and development loop initiated by Brimacombe and his group in the 70s
for sonic single-pipe tuyere injection in converting might finally be attained in the coming years.

At this point, mention should be made that Prévost, Létourneau, Perez, Lind, and Lavoie [35]
reported the use of 3 high pressure tuyeres in the Noranda Converter (NCv) in addition to the
9 conventional low pressure tuyeres. Unfortunately, no details have been published so far about
the dimensions, operating pressure, enrichment level, and implementation date of the high
pressure tuyeres. An update on the operation of the NCv may be presented during the Ralph L.
Harris Memorial Symposium.

Shrouded Injection in Non-Ferrous Pyrometallurgy – Is The Future Finally Here?

The aim of this paper was to provide a historical perspective of the Canadian contributions to the
field of submerged gas injection: from the pioneering work of Savard and Lee at Canadian
Liquid Air in steelmaking (1948-1968) and of Themelis, Szekely and Tarasoff at Noranda in
non-ferrous metals processing (1960s), to the ground breaking research of the Brimacombe
group at UBC in gas injection phenomena in non-ferrous metals converting (1971-1986), and
finally to the push to put sonic injection to practice by Bustos at Air Liquide Canada (1991-2001)
whom I had the privilege to join along the way in 1996.

Although high oxygen shrouded injection was developed and promoted by Air Liquide for bath
smelting and converting, the sole successes occurred in converting (Falconbridge and Thai
Copper Industries). Of the two, only the Shrouded Tuyere Slag-Make Converter process at
Glencore-Xstrata, which won the 2009 MetSoc Innovation Award, is still in operation today with
ALSI Technology. The SMC has become for me a sort of Canadian beacon for the potential of
high oxygen shrouded injection in nonferrous pyrometallurgy.

From a technical point, shrouded injection would be simpler to test or implement in smelting,
compared to converting, since bath smelting is characterized by continuous operation within a
narrow range of bath composition and temperature and is not prone to thermal chock the way
batch converting is. With typically one single primary smelting vessel in a smelter, running trials
of any sort in the smelting reactor is a proposition that carries a high level of risk; any failure
could compromise the entire smelter production.

In comparison, a typical converting aisle has two or more converters, so with careful production
and maintenance planning, the risks of a trial can be somewhat mitigated. All risks considered, as
I mentioned in my introduction, the time may have come for bath smelting and converting of
sulfide mattes to re-evaluate the potential of high oxygen sonic injection and consider the
technology as part of the drive to boost productivity, improve energy efficiency, or achieve a
lower environmental footprint by reducing fossil fuel usage and off-gas volumes.
308 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

Executives and engineers at Chinese copper smelters have understood the opportunity that high
oxygen shrouded injection affords to copper bath smelting. The drive for the Chinese copper
industry was highly challenging, as expressed by K. Jiang, L. Li, Y. Feng, H. Wang, and
B. Wei [1]: “Under the background of the high expense for introducing foreign technology, the
complication of ingredients of copper concentrates and requirements of environmental
protection and energy conservation in China, it is an important task for our metallurgists to find
an effective way to solve SO2 pollution in PS converters and to provide a new copper smelting
process that is more advanced than the existing Mitsubishi process, flash smelting and other
continuous copper smelting processes with shorter process times, less investment, lower costs
and higher recovery rates and better comprehensive utilization”.

The early research on high intensity bath smelting in China was conducted by Shuikoushan
Mining Corporation in the 80s for lead smelting with pilot tests completed in 1991 on a
3 000 tonnes per day pilot furnace [37-38]. The technology was however not commercially
developed at the time. Shuikoushan Nonferrous Metal Co. Ltd. and China Nonferrous
Engineering and Research Institute (ENFI) continued research, developed the SKS bottom
blowing copper process in 1994 [1] with a first adoption in 2001 and implementation in 2008 at
the Tang Loong Smelter of the Sin Quyen Copper Complex in Vietnam. Jiang Jimu, one of the
developers of the SKS copper process, was reported to have considered the successful
implementation at Sin Quyen Copper Complex as “an important step of SKS copper smelting
process leaping from experiment to industrialized production, and is also an important step for
China's independent intellectual property right to enter the international market.” [39].

The second (and larger) commercial implementation of the SKS bottom blowing copper smelting
furnace took place in 2008 at the Dongying Fangyuan Nonferrous Metals Co., Ltd. when
production was expanded from scrap re-melting to copper concentrate smelting at a nominal
capacity of 100 kt/a [1, 38]. K. Jiang et al. [1] described the SKS copper furnace as similar to the
Noranda Reactor though contrasting by the fact that the blast is blown via high oxygen ‘guns’,
which are Savard-Lee type shrouded injectors using air as a shrouding medium.

Considering the above description of the process, none of its individual elements seems
innovative in itself. The name of the process, ‘Bottom Blown Oxygen furnace’ is also highly
reminiscent of the leap forward in the steel industry in the 60s with Oxygen Bottom Blown
Metallurgy, aka the OBM process, patented by Maxhütte and Air Liquide [7]. Bringing together
in one vessel all the global wealth of knowledge in copper making and submerged gas injection
phenomena is in my view a true achievement and a major step towards my introductory
comments from Hoefele and Brimacombe: “It is unlikely that current blowing practice in
converters can be tolerated indefinitely […] It may well be that the nonferrous converter of
tomorrow bears a striking resemblance to the steelmaking furnace of today”. This achievement
may be described as a ‘return to the future’ with the SKS process using shrouded injection being
at last the ‘OBM-type’ furnace of the nonferrous metal industry that Hoefele and Brimacombe
alluded to.

Even if the SKS process does not seem to be selected for large scale copper projects above
100 kt/a, as mentioned by K. Jiang et al. [1] – flashing smelting remaining the technology of
choice – the SKS process appears to be fairly successful and attractive for small-scale copper
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 309

smelters in China (100 kt/a or less). What will it take to change that trend? The SKS copper
process is new and plant sizes adopting the technology are already gradually increasing. If the
shrouded tuyeres prove to be robust, there seems to be no reason why smelting capacity over
150 000 to 200 000 kt/a would not be reached.

New research collaboration between the University of Queensland (UQ) in Australia and the
Chinese metals company Shandong Fangyuan Nonferrous Metals Group (Fangyuan) has been
announced by Kathy Grube in UQ’s online newsletter on 19 July 2011 [40]. Grube reported that
the partnership between UQ and Fangyuan “[…] aims to develop new copper smelting
technologies with lower fuel consumption, lower CO2 emissions and higher copper recovery.
[…] Fangyuan developed and industrialised the world’s first commercial oxygen enriched
bottom-blowing technology (referred to SKS Technology) in 2008 under the visionary leadership
of Chairman and President, Mr. Zhi-Xiang Cui.”

Grube further reported Mr. Cui as saying “[…] the Chinese Government had recommended that
SKS Technology be rolled out to replace old copper-making technologies in China. Fangyuan’s
installation is the first modern copper smelting technology owned by the Chinese industry with
strong backing from the Chinese Government. Oxygen enriched bottom-blowing technology is
one of [the] most advanced copper smelting technologies, which has lower fuel consumption,
lower CO2 emissions and higher copper recovery. We are looking forward to working with The
University of Queensland to further improve this technology.”

I am also eagerly looking forward to hearing and reading more about the future developments
from China and Australia on shrouded injection in nonferrous matte smelting. I also cannot help
but relate results of high oxygen shrouded injection in copper smelting achieved by SKS
Technology and those in copper and nickel converting achieved by ALSI Technology. I have
regularly been asked over the years of my involvement with ALSI Technology whether nitrogen
shrouding could be substituted by air shrouding. My answer was systematically the same. Sonic
conditions in both the inner pipe and annular space are the keys to successful shrouded injection.
The choice of shrouding medium is solely based on process specifics.

In the case of converting, especially under the highly corrosive conditions of slag making
conversion of the Glencore-Xstrata SMC vessel, where both bath composition and temperature
change drastically during batch cycles, nitrogen shrouding appeared the preferred route to
ensuring optimum protection of the refractory lining. I do believe that in less aggressive
converting environments, air shrouding should still provide substantial benefits compared with
conventional low pressure bubbling injection. However, without the benefits of air shrouding
trials – which had been planned but not yet conducted – I could never prove my beliefs with
facts. I have not yet found a technical paper describing the successful use of the SKS process in
converting; maybe success has already been achieved but not yet published. Until then, a simple
comparison of accretions formed during air shrouding versus nitrogen shrouding offers some
qualitative insights.

Close-up photographs of accretions formed during shrouded injection using ALSI Technology in
nickel and copper converting are shown in Figure 39. The accretions exhibit a dual shape
accretion formed of a mushroom base protecting the periphery of the injector tips (sonic nitrogen
310 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

shrouding effect) combined with a pipe-type accretion pushing the gas-bath interface further
away from the refractory (sonic oxygen enriched air effect). Those combined mushroom and
pipe-type accretions have been deemed to be one of the key reasons of the success of ALSI
Technology in the SMC vessel. As for air shrouding, the only photograph of accretions I could
find for the SKS Technology in published literature is shown on Figure 40. In comparison, the
accretions only exhibit the pipe-type shape typical of sonic injection. Those have been sufficient
to ensure great success in copper smelting for the SKS furnace. Could air shrouding accretions
provide sufficient protection in highly corrosive converter environment? I am sure plant trials
will be conducted sooner than later, providing the necessary facts in lieu of an answer.

(a) (b)
Figure 39. Combined mushroom-type (from N2) and pipe-type (from Air-O2) accretions at the
tips of shrouded injectors (ALSI Technology) in (a) nickel, and (b) copper converting.

Figure 40. Pipe-type accretion growth at the tips of bottom blowing oxygen injectors
(SKS Technology) [38].

Designing shrouded injectors must ensure the right balance of cooling: enough to ensure the
formation of protective accretions though not in excess to avoid the formation of oversized,
lower porosity accretions. Higher oxygen enrichment per injector and a lower number of
injectors per vessel is the optimum approach with ALSI Technology. One of the most persistent
myths about high oxygen shrouded injection is the belief that high enrichment per injector, say
above 40% O2, is tantamount to overblowing of the converter charge and out-of-control heat
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 311

generation. A detailed heat and mass balance would correct that perception. In fact, if properly
designed, a small number of shrouded injectors could inject less oxygen than a conventional
converter with 48 to 60 tuyeres blowing air, and yet be much more efficient as the heat losses
through the off-gases would be significantly reduced. A controlled excess of heat in the bath also
becomes an opportunity to reprocess scrap or reverts. Retaining full control of the heat balance,
even with larger excess heat, can be achieved when the cold dope is delivered to the converters
without cranes, typically via a chute. The Glencore Xstrata nickel smelter in Sudbury has
mastered the practice, even using submerged sonic injection of nitrogen under the chute to
provide appropriate stirring of the solid additions [29-30]. For bath smelting, a higher overall
oxygen enrichment could translate into a higher smelting rate, further intensifying the smelting
rate, as achieved with the SKS furnace. Once again, I cannot wait to witness the upcoming
developments of shrouded injection in bath smelting and converting. I also hope to soon have the
opportunity to visit one of the new high intensity vessels.

Epilogue

I do hope that my paper has convinced some of the younger engineers that even if some ground
breaking technologies were developed 40 or 50 years ago, there are still great opportunities to
push those developments yet a few steps further, especially as energy conservation and
efficiency, lower concentrate quality, and environmental footprint are taking more of the
forefront of the industry concerns and preoccupations. Professor Ralph L. Harris was never shy
about stirring discussions, sometimes with unconventional or even controversial and thought-
provoking comments, a sure way to trigger a reaction from his interlocutors and some depth in
their answers. I hope I have found a balance in my challenging of myths, stating of facts, and
expressing of a few dreams. I also hope my statements and opinions will be challenged by my
peers, particularly those strengthened by the results of their pilot testing or plant trials, or better
still, by their commercial implementation successes.

Acknowledgements

I am very grateful to Air Liquide Canada for granting me access to their archives while preparing
my manuscript. This access allowed me to better research some of the historical facts while also
discovering old photographs of the golden years of R&D. The Managements of Falconbridge at
the time (now Glencore Xstrata), Thai Copper Industries, and Lonmin are gratefully
acknowledged for allowing me to document photographically the technology trials and
implementation at their respective sites. I also wish to thank BBA for the permission and support
to prepare this manuscript. My sincere appreciation also goes to Dr. Phillip Mackey for
reviewing my manuscript and providing valuable comments and suggestions. Above all, I am
greatly appreciative of the opportunity to contribute to the Ralph L. Harris Memorial Symposium
that Cameron Harris, Matthew Kreuh and Sina-Kashani-Nejad, symposium co-organizers, have
offered me. Their patience in waiting for my manuscript, and then editing it, is commendable.

References

1. E.O. Hoefele and J.K. Brimacombe, “Flow Regimes in Submerged Gas Injection,”
Metallurgical Transactions B, 10B (1979), 631–648.
312 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

2. J.K. Brimacombe, K. Nakanishi, P.E. Anagbo, and G.G. Richards, “Process Dynamics: Gas-
Liquid,” Proceedings of the Elliott Symposium on Chemical Process Metallurgy,
(Cambridge, MA: Massachusetts Institute of Technology, 1990), 343–412.

3. P.J. Mackey and J.K. Brimacombe, “Savard and Lee – Transforming the Metallurgical
Landscape,” Proceedings of the Savard/Lee International Symposium on Bath Smelting, ed.
J.K. Brimacombe, P.J. Mackey, G.J.W. Kor, C. Bickert, and M.G. Ranade (Warrendale, PA:
The Minerals, Metals and Materials Society of AIME, 1992), 3–28.

4. H. Bessemer, “Manufacture of Iron,” British Patent No. 2 768, 7 December 1855.

5. G. Savard and R. Lee, “Method and Apparatus for Treating Molten Metal with Oxygen,”
US Patent No. 2 855 293, 7 October 1958.

6. G. Savard and R. Lee, “Improvements in Metallurgical Process,” French Patent


No. 1 450 718, 18 July 1966.

7. H. Knuppel, K. Brotzmann, H.G. Fassbieder, G. Savard and R. Lee, “Method for Refining
Pig-Iron into Steel”, U.S. Patent No. 3 706 549, 19 December 1972.

8. G. Savard and R.G.H. Lee, “Submerged Oxygen Injection for Pyrometallurgy,” Proceedings
of the Savard/Lee International Symposium on Bath Smelting, ed. J.K. Brimacombe,
P.J. Mackey, G.J.W. Kor, C. Bickert, and M.G. Ranade (Warrendale, PA: The Minerals,
Metals and Materials Society of AIME, 1992), 645–660.

9. R. Lee, “Innovations in Ferrous Pyrometallurgy – A Canadian Perspective,” CIM Bulletin, 84


(950) (1991), 125–131.

10. J.K. Brimacombe, P.J. Mackey, G.J.W. Kor, C. Bickert, and M.G. Ranade, ed., Proceedings
of the Savard/Lee International Symposium on Bath Smelting (Warrendale, PA: The
Minerals, Metals and Materials Society of AIME, 1992).

11. N.J. Themelis, P. Tarassoff, and J. Szekely, “Gas-Liquid Momentum Transfer in a Copper
Converter,” Transaction of the Metallurgical Society of AIME, 245 (1969), 2425–2433.

12. G.C. McKerrow and D.G. Pannell, “Gaseous Deoxidation of Anode Copper at the Noranda
Smelter,” Canadian Metallurgical Quarterly, 11 (4) (1972), 629–633.

13. P.J. Mackey, Private Communication, 21 June 2013.

14. G.N. Oryall and J.K. Brimacombe, “The Physical Behavior of a Gas Jet Injected Horizontally
into Liquid Metal,” Metallurgical Transactions B, 7B (1976), 391–403.

15. J.K. Brimacombe and E.O. Hoefele, “Non-Ferrous Metal Treatment,” U.S. Patent
No. 4 238 228, 9 December 1980.

16. J.K. Brimacombe, S.E. Meredith and R.G.H. Lee, “High Pressure Injection of Air into a
Peirce-Smith Converter,” Metallurgical Transactions B, 15B (1984), 243–250.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 313

17. A.A. Bustos, G.G. Richards, N.B. Gray, and J.K. Brimacombe, “Injection Phenomena in
Nonferrous Processes,” Metallurgical Transactions B, 15B (1984), 77–89.

18. G.G. Richards, K.J. Legeard, A.A. Bustos, J.K. Brimacombe, and D. Jorgensen, “Bath
slopping and splashing in the copper converter,” Proceedings of the Reinhardt Schuhmann
International Symposium on Innovative Technology and Reactor Design in Extraction
Metallurgy, ed. D.R. Gaskel, J.P. Hager, J.E. Hoffmann, P.J. Mackey (Warrendale, PA: The
Minerals, Metals and Materials Society of AIME, 1986), 385–402.

19. J.K. Brimacombe, A.A. Bustos, D. Jorgensen, and G.G. Richards, “Towards a Basic
Understanding of Injection Phenomena in the Copper Converters,” Proceedings of the
Physical Chemistry of Extractive Metallurgy Symposium, (New York, NY: American
Institute of Mining Engineers, 1985), 327–351.

20. A.A. Bustos, “Injection Phenomena and Heat Transfer in Copper Converters” (PhD Thesis,
University of British Columbia, Vancouver, BC, 1984).

21. A.A. Bustos, J.K. Brimacombe, G.G. Richards, A. Vahed, and A. Pelletier, “Development of
Punchless Operation of Peirce-Smith Converters,” Proceedings of the Copper 87 – Cobre 87
International Conference, Vol. IV – Pyrometallurgy of Copper, ed. C. Diaz, C. Landolt and
A. Luraschi, (Santiago, Chile; Universidad de Chile, 1987), 347–373.

22. G.A. Casley, J. Middlin, and D. White, “Recent Developments in Reverberatory Furnace and
Converter Practice at the Mount Isa Mines Copper Smelter,” Extractive Metallurgy of
Copper, ed. J.C Yannopoulos and J.C. Agarwal (New York, NY: The Metallurgical Society
of AIME, 1976), 117–138.

23. A.A. Bustos, J.K. Brimacombe, and G.G. Richards, “Development of Punchless Operation of
Peirce-Smith Converters” (Internal Report, The Centre for Metallurgical Process
Engineering, University of British Columbia, 1986).

24. A.A. Bustos, “Process to Convert Non-Ferrous Metal such as Copper or Nickel by Oxygen
Enrichment,” US Patent No. 5 435 833, 25 July 1995.

25. A.A. Bustos and J.P. Kapusta, “High Oxygen Shrouded Injection in Copper and Nickel
Converters,” Proceedings of the Brimacombe Memorial Symposium, ed. G.A. Irons and
A.W. Cramb (Montreal, QC: The Metallurgical Society of CIM, 2000), 107–124.

26. A.A. Bustos, M. Cardoen, and B. Janssens, “High Oxygen Enrichment at UM-Hoboken
Converters,” Proceedings of the Copper 95 – Cobre 95 International Conference, Vol. IV –
Pyrometallurgy of Copper, ed. W.J. Chen, C. Diaz, A. Luraschi and P.J. Mackey (Montreal,
QC: The Metallurgical Society of CIM, 1995), 255–269.

27. A.A. Bustos, J.P. Kapusta, B.R. Macnamara and M.R. Coffin, “High Oxygen Shrouded
Injection at Falconbridge,” Proceedings of the Copper 99 – Cobre 99 International
Conference, Vol. VI – Smelting, Technology Development, Process Modeling and
Fundamentals, ed. C. Diaz, C. Landolt and T. Utigard (Warrendale, PA: The Minerals,
Metals and Materials Society of AIME, 1999), 93–107.
314 RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM

28. J.P. Kapusta, H. Stickling and W. Tai, “High Oxygen Shrouded Injection at Falconbridge:
Five Years of Operation,” Proceedings of Converter and Fire Refining Practices, ed.
A. Ross, T. Warner and K. Scholey (Warrendale, PA: The Minerals, Metals and Materials
Society of AIME, 2005), 47–60.

29. B. Salt, and E. Cerilli, “Evolution of the Converter Aisle at Xstrata Nickel’s Sudbury
Smelter,” Proceedings of the International Peirce-Smith Converting Centennial Symposium,
ed. J.P.T. Kapusta and A.E.M. Warner (Warrendale, PA: The Minerals, Metals and Materials
Society of AIME, 2009), 135–149.

30. B. Salt, and E. Cerilli, “Converter Aisle Improvements at Xstrata Nickel’s Sudbury Smelter,”
Proceedings of the Pyrometallurgy of Nickel and Cobalt Symposium, 48th Conference of
Metallurgists, ed. J. Liu, J. Peacey, M. Barati, S. Kashani-Nejad, and B. Davis (Montreal,
QC: The Metallurgical Society of CIM, 2009), 333–349.

31. J.P.T. Kapusta, N. Wachgama, and R.U. Pagador, “Implementation of the Air Liquide
Shrouded Injector (ALSI) Technology at the Thai Copper Industries Smelter,” Proceedings
of Cu2007, Sixth International Copper-Cobre Conference, Vol. III (Book 1) – The Carlos
Diaz Symposium on Pyrometallurgy, ed. A.E.M. Warner, C.J. Newman, A. Vahed,
D.B. George, P.J. Mackey, and A. Warczok (Montreal, QC: The Metallurgical Society of
CIM, 2007), 483–500.

32. R. Pagador, N. Wachgama, C. Khuankla, and J.P. Kapusta, “Operation of the Air Liquide
Shrouded Injector (ALSI) Technology in a Hoboken Siphon Converter,” Proceedings of the
International Peirce-Smith Converting Centennial Symposium, ed. J.P.T Kapusta and
A.E.M. Warner (Warrendale, PA: The Minerals, Metals and Materials Society of AIME,
2009), 367–381.

33. J.J. Eksteen, B. Van Beek, and G.A. Bezuidenhout, “Cracking a Hard Nut: An Overview of
Lonmin’s Operations Directed at Smelting UG2-Rich Concentrate Blends,” Proceedings of
the Southern African Pyrometallurgy 2011 Conference, ed. R.T. Jones and P. den Hoed
(Johannesburg, SA: The Southern African Institute of Mining and Metallurgy (SAIMM),
2011), 231–251.

34. J.P.T Kapusta, J. Davis, G.A. Bezuidenhout, S. Lefume, and D.K. Chibwe, “Industrial
Evaluation of Sonic Injection in a Peirce-Smith Converter at the Lonmin Platinum Smelter,”
Proceedings of the 51st Conference of Metallurgists – Towards Clean Metallurgical
Processing for Profit, Social and Environmental Stewardship Symposium, ed. R.
Schonewille, D. Rioux, S. Kashani-Nejad, M. Kreuh, and M.E.S. Muinonen (Montreal, QC:
The Metallurgical Society of CIM, 2011), 43–58.

35. Y. Prévost, P. Létourneau, H. Perez, P. Lind, and A. Lavoie, “Improving Flexibility: Recent
Developments at The Horne Smelter,” Proceedings of Cu2007, Sixth International Copper-
Cobre Conference, Vol. III (Book 1) – The Carlos Diaz Symposium on Pyrometallurgy, ed.
A.E.M. Warner, C.J. Newman, A. Vahed, D.B. George, P.J. Mackey, and A. Warczok
(Montreal, QC: The Metallurgical Society of CIM, 2007), 167–179.
RALPH LLOYD HARRIS MEMORIAL SYMPOSIUM 315

36. Kaixi Jiang, Lan Li, Yaping Feng, Haibei Wang, Bang Wei, “The Development of China’s
Primary Copper Smelting Technologies,” Proceedings of the T.T. Chen Honorary
Symposium on Hydrometallurgy, Electrometallurgy and Materials Characterization, ed.
S. Wang, J.E. Dutrizac, M.L. Free, J.Y. Hwang, and D. Kim (Warrendale, PA: The Minerals,
Metals and Materials Society of AIME, 2012), 167–176.

37. Li Cheng, Wang Jianming, Wang Zhogshi, Jiang Jimu, and Huang Qixing, “The SKS Copper
Smelting Process in China,” Proceedings of Cu99, Fourth International Copper-Cobre
Conference, Vol. VI – Smelting, Technology Development, Process Modeling and
Fundamentals, ed. C. Diaz, C. Landolt and T. Utigard (Warrendale, PA: The Minerals,
Metals and Materials Society of AIME, 1999), 83–91.

38. Baojun Zhao, Zhixiang Cui, Zhi Wang, “A New Copper Smelting Technology – Bottom
Blown Oxygen Furnace Developed at Dongying Fangyuan Nonferrous Metals,” Fourth
International Symposium on High-Temperature Metallurgical Processing, ed. T. Jiang,
J.Y. Hwang, P.J. Mackey, O. Yucel, and G. Zhou (Warrendale, PA: The Minerals, Metals
and Materials Society of AIME, 2013), 3–10.

39. Anonymous, “Patent Series of Oxygen Bottom Blow Smelting Process,” MCC Times, 16
(2008), 19.

40. K. Grube, “Fangyuan Fellowship Drives Innovations in Copper Smelting Technology,”


University of Queensland Online Newsletter (http://www.uq.edu.au/news/?article=23528),
19 July 2011.

View publication stats

You might also like