The Tap-Hole - Key To Furnace Performance PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

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

net/publication/304561704

The tap-hole - key to furnace performance

Conference Paper  in  Journal of the Southern African Institute of Mining and Metallurgy · May 2014
DOI: 10.17159/2411-9717/2016/v116n5a12

CITATIONS READS

6 2,884

1 author:

Lloyd Nelson
Anglo American
26 PUBLICATIONS   124 CITATIONS   

SEE PROFILE

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

Converter processing of platinum group metals View project

http://dx.doi.org/10.17159/2411-9717/2016/v116n5a12 View project

All content following this page was uploaded by Lloyd Nelson on 04 September 2016.

The user has requested enhancement of the downloaded file.


http://dx.doi.org/10.17159/2411-9717/2016/v116n5a12

The tap-hole — key to furnace


performance
by L.R. Nelson* and R.J. Hundermark†

This is largely a consequence of differing


processing conditions (process temperature,
"1AC=<B<
superheat (T), Prandtl number, Pr = CP/k,
The critical importance of tap-hole design and management for furnace
where  = dynamic viscosity, CP = specific heat
performance and longevity is explored through examining some of the
capacity and k = thermal conductivity, and
specific matte, metal, and slag tapping requirements of non-ferrous copper
blister and matte converting and smelting, ferroalloy smelting, and resulting heat flux). But this can also be
ironmaking systems. Process conditions and productivity requirements and influenced strongly by industrial operating
their influence on tapping are reviewed for these different pyrometallurgical philosophy in terms of furnace design for
systems. Some critical aspects of the evolution of tap-hole design to meet the campaign life longevity (i.e. greater capital
diverging process and tapping duties are examined. Differences and expenditure for longer, say 20–30 years’ life)
similarities in tapping practices and tap-hole management are reviewed. versus furnace productivity (i.e. number of
Finally, core aspects of tap-hole equipment and maintenance are identified – heats/campaigns to provide the greatest
aspects that are considered important for securing improved tap-hole possible dilution of fixed costs per unit of
performance and life, so pivotal to superior furnace smelting performance.
commodity produced). And this may not even
F13C@;< be consistent within a given commodity; all
tapping, tap-hole, ironmaking, ferroalloy, non-ferrous, matte, slag, blister, ironmakers (blast furnace (BF) campaign life–
smelting. based) supply downstream steelmakers (who
use heat/campaign-based converters and/or
electric arc furnaces).
However, regardless of the specific tap-
AE@C;8>EBCA hole configuration or operating philosophy,
The sheer diversity of tapping configurations owing to the addition of dynamic (often
used on industrial pyrometallurgical periodic) and more intense process conditions
operations is at first bewildering. They range (exposure to higher temperatures leading to
from historical tilting furnaces without tap- accelerated corrosion, greater turbulence, and
holes to modern eccentric bottom tapping elevated rates of mass and heat transfer) and
(EBT) tilting and/or bottom slide-gate electric higher concomitant thermomechanical forces
arc furnaces; to classical single tap-hole (from thermal or flow shear stresses), furnace
multiphase tapping (e.g. metal/matte and performance and longevity is intimately linked
slag); to dedicated phase tap-holes (e.g. to tap-hole performance. For good reason Van
dedicated metal/matte-only and slag-only); to Laar (2001) titled his paper ‘The taphole: the
dedicated phase multiple tap-hole heart of the blast furnace’ at the 2001
configurations (up to eight metal/matte-only symposium entitled The taphole – the blast
tap-holes and six slag-only tap-holes); to more furnace lifeline (Irons, 2001), while the title of
esoteric metal/matte-only siphons and slag the 2010 Coetzee and Sylven (2010)
overflow skimming, e.g. Mitsubishi contribution ‘No taphole – no furnace’ and the
Continuous Process (Matsutani, n.d.). This can staging of the SAIMM Furnace Tapping
be further complicated by periodic batch conference in 2014 suggest continued
tapping; consecutive tapping on a given tap- criticality and relevance.
hole; alternating tap-hole tapping practice;
near-continuous slag-only tapping, with
discrete batch matte/metal tapping on higher
productivity, but low metal/matte fall (<20%
by mass feed) Co and Ni ferroalloy and
* Anglo American Platinum Ltd.
platinum group metal (PGM) matte furnaces;
† Anglo American plc.
near-continuous tapping through batch
© The Southern African Institute of Mining and
tapping of individual tap-holes that are opened
Metallurgy, 2016. ISSN 2225-6253. This paper
consecutively (Tanzil et al., 2001; Post et al., was first presented at the, Furnace Tapping
2003); to fully continuous tapping on coupled Conference 2014, 27–28 May 2014, Misty Hills
multi-furnace cascades (Matsutani, n.d.). Country Hotel, Muldersdrift, South Africa.
L

 
       VOLUME 116  465
The tap-hole — key to furnace performance
By first comparing and contrasting some of the process ® Sheer metal fall and productivity of ironmaking BFs
conditions and resulting tap-hole and tapping requirements >10 000 t/day hot metal (HM), achieved through near-
of different commodities, we make an attempt at identifying continuous tapping at more than double the rate and
key elements of tap-hole design, physical tapping practices, velocity of, but through tap-hole diameters not too
equipment, and monitoring and maintenance practices dissimilar to, other commodities
characteristic of superior tap-hole management and required ® High pressure of tapping liquids of ironmaking BFs (up
to secure increased tap-hole performance and prolonged life. to 5 bar blast pressure at tuyeres, to add to already
high hydrostatic pressure of comparatively thick slag
C55C;BE16<=F>B7B>G=@C>F<<GDA;GC=F@DEBA9 and thick and dense metal)
>CA;BEBCA< ® More limited accessibility of smaller circular blast and
To provide some context to the range of tap-hole designs, electric furnaces (EFs) (up to 22 m diameter) to
and operating and maintenance practices adopted for multiple tap-holes, than larger rectangular six-in-line
different commodities, it is instructive to compare some key (6iL) furnaces (up to 36 × 12 m)
process physicochemical and operating conditions prevailing. ® Low comparative temperatures and superheats of
Notable features include: (often near-autogenous) copper smelting

Table I
Indicative1 process limits, properties and operating conditions for specific commodities
Iron making Cr ferroalloy Mn ferroalloy Ni ferroalloy Cu blister/matte Ni Matte PGM matte

Furnace BF SAF/DC–arc BF/SAF Circ/6iL EF FF/TSL 6iL/TSL/FF 6iL/Circ/TSL


M + S tap–holes 1–4 1–3, 1–2+1–2 1–2, 2+2 2+4–6 2–8+2–6 2+2 2–3+2–3
Tmetal/matte, °C 1480–1530 1500–1650 1300–1450 1430–1550 ~1170–1320 1150–1300 1300–1500
Tmetal/matte, °C ~350 50–100 50–150 20–350 100–250 50–300 400–650
Tslag, °C 1480–1530 1600–1750 1350–1550 1550–1630 1170–1350 1200–1400 1450–1600
Tslag, °C ~200 <50 50–100 50–150 50–100 50–150 50–200
qaverage, kW/m2 25 5 5 50–100 20–100 20–50 30–100
qpeak tap–hole, kW/m2 >200 >15 >15 >200 >300 >200 >300
metal/matte, t/m3 7 ~6.7 ~5.5 ~7.5 ~5–7.5 ~4.5 ~4.2
slag, t/m3 2.8–3.1 2.7–3.2 2.7–3.3 2.8–3.2 3.5–4 2.8–3.2& 2.8–3.2
μmetal/matte, Pa.s ~0.007 ~0.007 0.005 ~0.006 0.002–0.005 0.003(0.05&) 0.0025
μslag, Pa.s 0.1 ~0.5 0.7–1.5 ~0.5 0.03–0.07 0.3 0.3
kmetal/matte,NW/m°C 50 ~20 ~14 ~30 ~5–160 17& 17&
kslag,NW/m°C ~0.5 ~0.2 ~0.2 ~0.7 ~2–8 0.8 (8&) ~0.8
Cp,metal/matte, MJ/t°C 0.8 ~0.9 ~0.9 ~0.5 ~0.5 ~0.7 ~0.8
Cp,slag, MJ/t°C ~1 ~1.7 ~1 ~1.2 ~1 1.25& ~1.3
metal/matte, /°C 8 x10–5 7 x10–5 – 8 x10–5 1 x10–5 1 x10–4 1 x10–4
slag, /°C – – – – – 3 x10–4& 3 x10–4
Prmetal/matte 0.1 0.3 0.5 0.2 0.01 0.13 (2.1&) 0.12
Prslag 50 – – – – 470 (47&) ~450
Hmetal/matte > MTH, m ~2 0.3–0.6 0.3–0.6 0.15–0.3 0.25–0.4 0.25 ~0.3
Hmetal/matte–STH+Top, m ~2 (0.3+) ~1 (0.3+) ~1 0.6–1+0.4–1 0.2–0.4+0.2–0.4 0.2–0.4+0.2–0.6 0.5+0.6–0.9
Ptop of liquid level, bar 5 >1$ >1$ >1$ ~1 ~1 >1$
dmetal/matte tap–hole, m# ~0.07 0.07–0.2 0.04–0.1 0.04–0.1 ~0.05 ~0.07 0.04–0.07
vtapping, m/s 5 (to 8) ~4 ~2–4 ~2–4 ~2–4 ~2–4 ~2–4
.
m metal/matte, t/min* 7 ~1–4 1–2.5§ ~1.5–3 1–3 ~2.5 0.5–1.5§
Metal/matte fall 60–75% 35–50% 35–60% 5–20% ~40% 30–40% 10–25%
Tap–hole repair, w 4 >12 >26 1–2/8 4 3–9/26 1–4/12
Tap–hole life, y 10 (12) 2–6 2–6 1–4 1–4 1–3 1–2
Furnace life, y 15–20 12 20 20 6–12 30 12

#Non-HM tap-holes often start ~40 mm diameter *FA and non-ferrous instantaneous batch mass tapping rate &(Sheng et al., 1998).
NAt §Higher
. $Operate
process temperature; Mn solid at 727°C value also typical m slag with significant charge burden

1Some operations may operate quite far from these generically indicative values. Mills and Keene, (1987) and Sundstrӧm et al. (2008) provide much of the slag

and matte properties data, respectively


L

466 VOLUME 116


 
      
The tap-hole — key to furnace performance
® Relatively low superheats of ferroalloys (FA) in DC arc (as defined by the prevailing coolant operating
and submerged-arc furnaces (SAFs) pressure).
® Higher viscosity (and Pr), but lower thermal Somewhat paradoxically, when the thermal conductivity
conductivity and density of slag than metal/matte of matte is accounted for (kmatte approximately 20 times that
® High thermal conductivity (k) of liquid blister Cu of kslag), estimates of hmatte remain approximately 20 times
® Extreme superheat (T ) of PGM matte (Shaw et al., that of hslag. This is despite the significantly higher Pr
2012; Hundermark et al., 2014). number of slag (Robertson and Kang, 1999; Table I) and its
positive contribution to both natural and forced convection
"?D9G7@FF!FG?BABA9G)  5DEEF4?B<EF@G>C==F@G%:BE& heat transfer Nusselt numbers through correlations:3 Nu =
=CEFAEBD?< hL/k  (GrPr)¼ and (Re½Pr1/3), respectively.
A striking industrial observation is the ease with which slag So, considering the first condition, compared to slag,
freeze linings can be formed and maintained (almost ‘self– superheated matte of potentially four times greater superheat
healing’) from even superheated slag, provided cooling is (Tmatte up to 650°C) and approximately 20 times the
adequate. It is also quite remarkable how effectively just a convective heat transfer coefficient delivers far greater
thin accretion layer of slag (a couple of millimetres thick) can incident heat flux than slag (qmatte = hmatte Tmatte = approx.
provide a sufficient thermal resistance to appreciably lower 80qslag) and so is capable of up to a couple of orders of
critical lining and copper hot-face temperatures. magnitude greater thermal ‘hit’ of the cooling system
In stark contrast, especially in PGM matte and blister Cu (condition 1 above). This higher heat flux of matte compared
processing, equivalent matte/metal accretion formation often to slag leads to higher temperatures of critical lining hot-
seems near impossible to achieve, to the extent that the faces (e.g. refractory and copper cooler – conditions 2 and 3),
operation of copper coolers on blister Cu requires which then (condition 2) all too easily exceed the unusually
‘demonstrated ability to maintain a protective accretion low Tf of matte, due to its unusually low solidus (850°C) and
coating’ (George, 2002). Or stated in another way in the PGM even liquidus (950°C) temperatures.
matte industry: the operation of copper coolers unprotected In such a situation a copper cooler unprotected by any
from direct contact with superheated liquid matte is simply alternative thermal barrier (e.g. refractory/slag) is at
not tolerated. significant risk from any superheated matte/blister Cu ‘hit’
Considering the heat transfer conditions applicable to the that can rapidly lead to hot-face temperatures rising to where
successful implementation of a water-cooled composite the cooler copper simply melts (1085°C). Yet for most slag
copper lining, four key criteria can be defined when systems these conditions are rarely violated; stable slag
considering the influence of process heat flux, q = hbT accretion freeze linings prevail, supported additionally by a
(where T = TB – Tf and hb = convective heat transfer high-viscosity slag ‘mushy zone’ adjacent to Tf (Guevara
coefficient from bulk process liquid of temperature TB, to and Irons, 2007) to protect the composite cooling system.
accretion freeze lining2 of temperature Tf ), into and out Comparing kmatte, kFA, kHM, and kblister Cu of 17, 10, 50, and
through the composite cooling system. The latter is described 160 W/m°C and resulting Prmatte, PrFA, PrHM, and Prblister Cu
for the simplest one-dimensional case by values of approximately 0.2, 0.2–0.5, 0.1, and 0.01,
respectively (Table I), one can estimate ratios of convective
qC = (Tf – TC)/(xf /kf + xR/kR + 1/hI + xC /kC + 1/hC)
heat transfer relative to PGM matte as hmatte:hFA:hHM:hblister Cu
where qC = composite cooler heat flux; Tf = effective = 1:~1.5:~2:~5, respectively. Relative to matte, convective
accretion freeze lining temperature in contact with process heat transfer coefficients of HM and blister Cu are greater.
liquid (whether matte or slag); TC = bulk temperature of Maximum superheats TPGM matte, TFA, THM, Tblister Cu of
cooling fluid; xf and kf are, respectively, thickness and 650, 150–350, 350, and 350°C, respectively, will tend
thermal conductivity of the accretion freeze lining; xR and kR somewhat to help balance the resulting process heat fluxes, q
are thickness and thermal conductivity of the residual = hT. So it would appear that it is low Tf (listed here at its
refractory; hI = convective heat transfer coefficient at the solidus lowest extreme) of Tmatte, TFA, THM, Tblister Cu of 850,
cooler hot-face; xC and kC are thickness and thermal >1250, 1130, and 1065°C that most limit the ability to form a
conductivity of residual refractory; and hC = convective heat protective accretion freeze lining, and so render copper
transfer coefficient of the cooling medium (e.g. air or water). coolers ultimately more prone to thermal ‘hit’ by (PGM)
Following the example of Robertson and Kang (1999), matte/blister Cu.
we describe some relevant limiting conditions for such a heat
transfer system:
(1) For an accretion to freeze (sustainably), q must be
less than qC
2T commonly used to describe the real freeze-lining temperature Tf.
(2) The cooling system hot-face temperature (be it liquidus
Recently, Fallah-Mehrjardi and co-authors (2014) proposed a
refractory or copper) must be less than Tf of the mechanism that supports the temperature of the interface of stationary
specific accretion in question (be it metal/matte or steady-state freeze-lining deposit (Tf ) being lower than the liquidus
slag) temperature (but no lower than Tsolidus), which potentially facilitates
(3) The copper hot-face temperature must not exceed operations with freeze linings at temperatures below the liquiduss.
copper’s melting point (or copper’s long-term service
3Grashof number, Gr = gΔTL3/(/ )2, Reynolds number, Re = vL/,
g = gravitational acceleration,  = volume expansion coefficient, T =
limit of < 461°C) surface to bulk liquid temperature difference, L = characteristic length,
(4) Usually, unless specifically designed for, the boiling  = dynamic viscosity, = density, = fluid velocity, h = convective heat
point of the cooling medium should not be exceeded transfer coefficient, and k = thermal conductivity.
L

 
       VOLUME 116  467
The tap-hole — key to furnace performance
AEF9@DEF;GED=6:C?FGDA;GED==BA9G<1<EF5G5DAD9F5FAE which phase is being withdrawn (typically < 0.625 when
lower viscosity phase is withdrawn;  0.8 when uppermost
Key aspects of tap-hole design and tapping operation,
viscous layer is withdrawn), Fr = /(dg / ) and is the
maintenance, and monitoring will be presented separately for
discharge velocity,  is the density difference between
convenience. However, it should be emphasized that all
heavier and lighter liquid, and is the density of the lighter
aspects need to be considered as part of an integral system,
liquid (Liow et al., 2001, 2003). Using assumed
which must be managed as such for success. Overly focusing
physicochemical properties and tap-hole conditions (Table I),
on one component at the expense of another (e.g. tap-hole
one can predict he of the order of 0.12 m for copper FF settler
clay optimization, without due consideration for mudgun and
and PGM EF smelting (and theoretically even ironmaking BF
drill capabilities) is unlikely to yield optimal results. A ‘chain
conditions). Not too surprisingly, therefore, the dedicated
being only as strong as its weakest link’ adequately describes
slag tap-holes located up to 1 m above the metal/matte tap-
the role of integration of all aspects of the tap-hole and
holes, coupled with tight metal/matte level control (to a
tapping into a comprehensive system for sound management.
maximum height of 0.25–0.4 m above matte tap-holes on
blister Cu and PGM matte furnaces – Table I), permit slag
'1=F<GC7GED==BA9G<1<EF5< tapping substantially free of metal/matte from the interface
Tapping systems can be conveniently categorized according with the bulk slag, and entrained specifically through tapping
to the product phases being tapped and the process (ignoring the presence by other sources of entrained and
conditions prevailing: primarily temperature, T (versus unsettled metal/matte droplets).
solidus or liquidus), k, and Pr. Similar two-phase liquid entrainment and an initial
declination of the slag interface towards the tap-hole as
#("&#'%(!!$& tapping commences followed by a switch to initial inclination
With its high Pr number and elevated melting properties and even ‘pumping’ out of the tap-hole later in the tap has
(Table I), slag – provided it is kept free of metal/matte/bullion been modelled on BFs by CFD (Shao, 2013; Shao and Saxen,
– is potentially the simplest liquid for which to design an 2011, 2013a, 2013b). However, in the modelling of BF
effective tap-hole system, comprising merely a high-intensity tapping, He and co-authors (2012) caution that the metal
water-cooled copper slag tap-block protected by an accretion should not be maintained at a depth too low above the tap-
freeze lining of product slag. A significant advantage of slag- hole, as one runs a risk of entraining process gas by ‘viscous
only tapping is that it facilitates direct downstream treatment fingering’ during tapping, especially (1) when the slag
of slag by either traditional water granulation (Atland and viscosity is high, or (2) in the presence of a permeable bed of
Grabietz, 2001; Szymkowski and Bultitude-Paull, 1992), or, solids through tapping occurs (e.g. coke bed).
increasingly, ‘dry’ air atomization (sometimes with energy The efficacy of intense copper cooling (predominantly in a
recovery) to obtain useful slag products amenable to circular slag tap-block configuration) is clear (Figure 1 and
handling and sale in ironmaking, steelmaking, and Ni and Figure 2). These coolers directly impart a thicker protective
SiMn ferroalloy applications (Ando–, 1985; Rodd et al., 2010). freeze lining than the alternatives of just top lintel copper
Dedication of the tap-hole to slag is particularly effective blocks, or ‘inverted-U’ square copper blocks and circular
for handling corrosive slags (especially acidic slags >50% block water-cooled copper pin designs (Marx et al., 2005;
SiO2 that are fundamentally incompatible with basic and Henning et al., 2010) (the latter choosing rather to try to
some other refractory oxides), because there is no chemical moderate freeze lining thickness). These latter designs all
potential for reaction with a frozen slag of essentially the avoid the presence of water below the tap-hole. It is a moot
same composition. Thus retention of a protective freeze lining point whether this is indeed universally a safer situation,
reverts to a more predictable issue of designing for thermal especially if control of furnace operating levels is adequate,
equilibrium thickness, and adoption of suitable safety factors simply because of the less desirable trade-off of imparting an
to provide some protection against deviations therefrom. inherently thinner protective freeze lining with less cooling.
On many industrial furnaces, a combination of level Concerns frequently articulated of overly cooling copper
measurement and phase separation is more than adequate to coolers (Trapani et al., 2003; Marx et al., 2005; Henning et
tap slag free of metal/matte. Nishi (2007) reports on the al., 2010) are extravagant costs, fear of preventing easy tap-
importance of designing the height of the slag tap-hole to
avoid Mn ferroalloy discharge through it. This is also a
typical requirement of more quiescent EF or slag cleaning
furnace (SCF) processes of low (< 20%) metal/matte fall
(effectively ‘slag-making’ processes, that may even be subject
to near-continuous slag tapping, such as Co and Ni ferroalloy
and base metal and PGM matte smelting). On other matte
flash furnace (FF) to TSL converting processes (e.g. blister Cu
to PGM matte, respectively), it is typically necessary to equip
them with downstream FF settling and/or SCF processes for
further recovery of pay metals from slag, especially oxidic
losses that require recovery through reductive processes.
Theoretically, the critical height for entrainment (he) of a
two-layer liquid through an orifice of diameter (d) is related /B98@FG+-??GE:@FFG<?D9GED=6:C?F<GC=FAG3BE:GAFD@6>CAEBA8C8<G<?D9
to dFr0.4, where depends on the density difference and ED==BA9GCAGDGG5DEEFG78@AD>FG CEB>FG:CC;GFE@D>EBCA
L

468 VOLUME 116


 
      
The tap-hole — key to furnace performance
oxygen) in service with carbon-unsaturated metal/matte.
Corrosion of both carbon-based and oxide refractories is
invariably accelerated by slag, even to the extent that
corrosion becomes catastrophic, e.g. if acidic slags make
contact with basic refractories (such as magnesia).
Depending on the specific slag system, amphoteric (alumina)
refractories can also be susceptible to both acidic (e.g. high-
silica) or basic (e.g. high-lime) slags.
Refractory-lined overflow launders are used in
continuous tapping of copper matte and slag from the
Mitsubishi Continuous Process smelting furnace, and certain
corrosion challenges are presented (addressed largely by
/B98@FG.-B9:6BAEFA<BE1G3DEF@6>CC?F;G<?D9GED=64?C>,G3BE:G<C?B;B7BF; fused cast magnesia-chrome). Somewhat remarkably,
<?D9G7@FF!FG?BABA9G>C@FGE:@C89:G3:B>:G<?D9GB<GED==BA9GDEG+$# unlined water-cooled copper tap-plates are routinely fitted on
to the furnace exterior for combined matte-slag tapping
elsewhere in the copper industry, such as TSL furnaces. This
hole opening, or freezing of a tapping stream. Even with the presumably is only possible owing to the comparatively low
least intense top lintel or shallow-cooled (i.e. water circuits temperature (< 1200°C, Table I) and relatively low copper
outside the furnace) copper and refractory-lined slag tap- matte superheat in combination, critically, with slag that has
blocks, problems associated with the latter two operational the potential to freeze (even if only as a thin layer a couple of
aspects can occur, and are generally coupled with undesirable millimetres thick) as a protective accretion on copper tapping
increased copper slag tap-block wear rates. Szekely and surfaces.
DiNovo (1974), in a modelling study of the critical factors for Most combined metal-slag tap-hole processes are
tap-hole blockage of a molten stream (e.g. during tapping), characterized by lower slag-metal ratios of about 0.4–1.5 t
determined that nozzle diameter was most critical, followed slag per ton metal (metal fall is approximately 35–60% in the
by metal superheat, with the extent of preheating (or in this case of Cr and Mn ferroalloys, Table I), or significantly lower
case cooling) of the nozzle walls being less significant. 0.2–0.4 t slag per ton HM in ironmaking BFs (metal fall is
Effectively, this implies that the tapping channel diameter approximately 65%, Table I), to near-slagless tapping in Si
should be enlarged if the slag tapping stream is freezing. (and Si alloy) processes. A striking feature of the ironmaking
So again it is a moot point if reduced cooling intensity, BF is its sheer productivity (>10 000 t/day) coupled with
including the removal of water circuits from beneath the complex internal process structures (‘deadman’ and tap-hole
tapping channel, indeed universally represents the safer ‘mushrooms’). Even with multiple tap-holes, these process
option, if the consequent (sometimes inadequate) protective structures would complicate attempts to control hot metal and
freeze lining thickness results in increased copper hot-face slag levels adequately and to the extent necessary to permit
temperatures that will reduce the long-term integrity of the effective dedicated metal- and slag-only tapping. Therefore,
copper block itself (i.e. requires sustained temperatures below as with the majority of older ferroalloy SAFs and BFs, deep
461°C [Robertson and Kang, 1999]). Furthermore, if the tap- cooling is generally not contemplated, with limited water-
hole is still prone to ‘slow tapping’ even with less intense cooled elements being applied more judiciously.
cooling, it may suggest that an alternative operational
tapping strategy is appropriate.
)$(%)')%(#(%%)'%(!!$&
Some of the larger ferroalloy furnaces for Mn and DC Cr
Provided that metal/matte can be tapped substantially slag-
alloy production also operate separate slag tap-holes, which
free, a configuration for dedicated metal/matte tapping is
assist greatly in separating post-tap-hole metal- and slag-
possible. Theoretically, it can be calculated that the
handling logistics. In many instances the separate slag tap-
separation of slag to at least 0.07 m above the metal/matte
holes are merely refractory graphite/microporous
tap-hole should facilitate matte tapping without slag
carbon/carbon tap-blocks (usually the former two owing to
entrainment ( drops to 0.625 for tapping of the denser, less-
improved resistance to wetting and lower corrosion by slag).
viscous phase [Liow et al., 2003]). Efficient separation of
Increasingly, deep-cooled (i.e. water-cooled copper extending
metal/matte from slag already in the furnace decidedly
inside the furnace) copper lintel, or ‘inverted-U’ blocks are
simplifies post-tap-hole handling and associated logistics.
used to promote cooling of such refractory slag tap-holes.

" $&)')%(#(%%)'(&'#('%(!!$& ) )& ($&'%(!"#)


This is decidedly the norm, but it also often presents the Some furnaces are equipped with emergency/drain tap-holes
greatest design challenge because of the different natures of (Newman and Weaver, 2002) that are used when the furnace
slag and metal and their chemical incompatibility with linings does not drain from operating tap-holes (Cassini, 2001), or to
selected as suitable for the other phase. Traditionally, effect bath drainage to a lower level than normal operating
refractory tap-blocks (refractory oxide or carbon-based) were tap-holes for safer repairs. Some operators prefer to avoid
adopted for combined metal/matte and slag tapping. With few such tap-holes for fear that they potentially increase risk by
exceptions, the refractory oxides are relatively resilient to tempting non-emergency/non-drain use, and present another
metal- and matte-only tapping. Carbon-based tap-blocks risk weakened region of furnace lining (at a higher pressure
carbon dissolution and/or oxidation (e.g. by dissolved head) for unplanned drainage.
L

 
       VOLUME 116  469
The tap-hole — key to furnace performance

'D=6:C?FG;F<B9A Several refractory types (Figure 3) are used in BF


tapholes and their environs (Stokman et al., 2004; Jameson
(!"#)'(&'%(!!$&(&&)#')(%'% (&) et al., 1999; Irons, 2001; Van Laar, 2001; Van Laar et al.,
On a large furnace crucible wall, bath heat transfer can 2003; Brunnbauer et al., 2001; Atland and Grabietz, 2001).
reasonably be approximated as one-dimensional. In the They include:
simplest configuration of a long circular tap-hole, heat ® 100% alumina (the most ‘insulating’: k = 1–5 W/m°C)
transfer from a fast-flowing hot tapped liquid is dominated ® Pitch-impregnated carbon/alumina (Black and Bobek,
by radial heat loss in the passage down the tapping channel. 2001)
Even with a reasonably fast water cooling flow rate of ® Large carbon blocks (k approx. 14 W/m°C)
6 m3/h, it can readily be estimated using q = Q/A = (mCP)T ® Hot-pressed small carbon or semi-graphite bricks (a
that for just a 1°C rise in water temperature, the equivalent lower iron content of the latter, to reduce CO
tapping channel (tap-block or faceplate) heat flux (q) exceeds disintegration [Stokman et al., 2004; Spreij et al.,
0.5 MW/m2. 1995])
In a real tapping channel, in addition to the tapping ® Microporous (potential advantages of less metal
channel heat transfer, heat transfer from the contained infiltration if the maximum pore size is less than 1 m
furnace bath also exists, which results in a three-dimensional [Stokman et al., 2004; Piel et al., 1998; Spreij et al.,
heat transfer situation that is more extreme than in almost 1995; Tomala and Basista, 2007]), large carbon or
any other region of the furnace crucible. The tap-hole semi-graphite blocks
specifically is invariably subjected to the most arduous of ® Thermally conductive graphite (k approx. 140 W/m°C,
conditions (Van Laar et al., 2003; Van Ikelen et al., 2000): frequently applied as ‘safety’ tiles glued to the steel
the highest liquid (metal/matte and slag) velocities, affected wall in the immediate tap-block vicinity [Van Laar et
by the degree of radial or peripheral flow and total flow that al., 2003; Edwards and Hutchinson, 2001; Atland and
converge on the tap-hole to achieve the productivity set- Grabietz, 2001])
point; the highest turbulence (increased by gas entrainment ® Sometimes graphite with high-alumina silicon carbide
and even blowing under pressure, and associated enhanced castable in the centre (favoured for reasons of
mass and heat transfer from both stream tapping and improved tapping stream dissolution and erosion
through the action of any tap-hole clay flash devolatilization resistance over graphite in the event of the latter’s loss
and subsequent ‘boiling’ at the back of the channel); wildly of freeze lining or protective baked tap-hole clay inner
fluctuating and periodic thermal loads (from cool, dormant annulus, somewhat improved tolerance to oxygen
conditions, heating rapidly when the tap-hole is opened with lancing over graphite, provision of some heat storage
oxygen, or hot liquid tapping, and with tap-hole clays for tap-hole clay baking, and possibly some improved
‘boiling’ and gas bubble-driven circulation upon tap-hole tolerance to microcracking induced through mudgun
and drill impact forces)
closure); and high dynamic loads (the action of opening and
® The use of higher conductivity silicon carbide (Brown
closing a tap-hole). Tap-holes are also prone to gas leakage,
and Steele, 1988) in conjunction with a carbon
especially when operated under pressure in a BF, which may
surround and alumina tapping channel hot-face bricks
result (particularly in the case of ironmaking or ferroalloy
has also been reported (Yamashita et al., 1995). In
processes adopting carbon-based refractories) in a
some instances, heat removal is further enhanced by
continuous threat of exposure to, and reaction by, CO (the
the addition of water-cooled iron or copper tap-hole
risk of carbon deposition), oxidation by injected oxygen, air,
notch channels, or even water-cooled copper
or steam (especially if water leaks), slag and maybe even
inserts/plate coolers (Irons, 2001; Van Laar, 2001).
SiO(g), and reaction with volatile gas species such as alkalis
and zinc (which leads to refractory attack) (Van Laar, 2001;
Van Laar et al., 2003; Spreij et al., 1995; Iiyama et al., 1998;
Tomala and Basista, 2007).

(!"#)' ) (%" ')$&


Clearly, to be successful, tap-hole designs need to cater not
only for average, but peak, process heat flux conditions. Van
Laar (2014) suggests that in BF tap-holes, peak heat fluxes
exceeding 1 MW/m2 have been detected, which is
considerably in excess of the normal average heat fluxes
measured (25 kW/m2, Table I). This would not be
inconsistent with a 1.4 MW/m2 event involving metal
encroaching on the lower zone of a copper waffle cooler
recorded in Co ferroalloy production (Nelson et al., 2004).
Nearly all tap-holes are designed with a length that
exceeds the adjacent sidewall thickness. Unfortunately, this
provides only short-term protection against liquid breakout in /B98@FG-'D=6:C?FGDA;GFA(B@CA<GF8B==F;G3BE:G@F7@D>EC@1GC7G(D@BC8<
the tap-hole area, because the tap-hole length will at best E:F@5D?G>CA;8>EB(BEBF<GDA;GBAEF9@DEBCAGBAECG
/G?BABA9G)D7EF@G"EC,5DAG
rapidly recede to its thermal equilibrium dimension. et al.0G.22*
L

470 VOLUME 116


 
      
The tap-hole — key to furnace performance

Table II "#)'"'%)'%(!"#)'#(' ""


Carbon-based refractory and onset of key wear and The second crucial feature, specific to ironmaking BF tap-hole
attack mechanisms (Van Laar et al., 2003; Spreij design, is the active development and continuous renewal of
et al., 1995; Tomala and Basista, 2007) a tap-hole clay (also described as mud) ‘mushroom’ to
provide some hot-face protection on the back of the tapping
Thermomechanical and Onset channel (Figure 4) (Uenaka et al., 1989; Jameson et al.,
chemical attack mechanisms temperature* °C
1999; Eden et al., 2001; Nightingale et al., 2001, 2006;
Alkali and zinc# 400 Tanzil et al., 2001; Atland and Grabietz, 2001; Cassini, 2001;
CO deposition 450 Wells, 2002; Horita and Hara, 2005; Kageyama et al., 2005,
Stress cracking 500 2007; Nakamura et al., 2007; Niiya et al., 2012; Kitamura,
Oxidation (enriched, or air)# 600 2014). The ‘mushroom’ requires tap-hole clay for its
Steam oxidation 700 development and consists additionally of incorporated slag,
CO2 oxidation 1050 iron, and coke. Tsuchiya and co-workers (1998) hypothesize
Liquid penetration, corrosion (e.g., by carbon 1150 that a necessary condition for the development of a
dissolution, or by slag) and ensuing erosion#
‘mushroom’ is that the tap-hole length can be extended only
when the holding space for the injected tap-hole clay is
*Depending on specific refractory type; oxide- or carbon-based, calcined effectively realized, so that the major part of the tap-hole clay
anthracite or graphite aggregate, or binder-derived (Spreij et al., 1995)
surface is covered by the coke column (Figure 5). Niiya and
(binder more prone to attack than aggregate) and associated trace
impurity catalysts (e.g. Fe) co-authors (2012) hypothesized further that the tap-hole clay
#Especially in tap-hole region (Piel et al., 1988) is ‘extruded in the furnace like strings’ and that these ‘strings
accumulate in the coke-free spaces by folding together with
solidified iron and/or slag’. Other conditions required for
In all instances involving the use of composite refractory
types (Figure 3), especially when water-cooled components
are included, a critical design requirement is to cater for
differential thermal expansion properties that can easily differ
by an order of magnitude, with the potential to cause gaps,
stresses, and strains, so raising the potential for liquid
infiltration (Van Laar, 2014). An experience reported
(Duncanson and Sylven, 2011) of furnace campaign life
reduced from 14 to just 3 years when switching from a
design where ‘the original furnace had forced air cooling in
the bottom, but no additional (water) cooling for the furnace
walls’ (and, by inference, attempt at freeze lining in, or at
least near, the tap-block) may well illustrate this. Moreover,
the additional requirement for effective freeze linings around
thermal equilibrium has led Singh and co-authors (2007) to
/B98@FG-%8<:@CC5&G>CAAF>EBCAGC7GED=6:C?FG:CE67D>FGDA;G%;FD;5DA&
state: ‘but in the present Indian scenario with process B(F@<BCAGC7G;F<>FA;BA9G?B8B;GD@C8A;GE:FG%58<:@CC5&GB<G;F=B>EF;GA
parameters not stable … it is difficult to maintain the ED==BA90GE:B<G>C54BAF<G3BE:G=F@B=:F@D?G?B8B;G7?C3GD@C8A;GE:FG%;FD;5DA&
conditions inside the furnace desirable for a true freeze ECGBA>@FD<FGE:FG(F?C>BE1GC7GE:FG?B8B;G7?C3G4F?C3GE:FGED=6:C?FGDA;G?BABA9
lining,’ so failing to ‘give the expected lifetime of over 25 :CE67D>F0G3BE:GE:FG=CEFAEBD?G7C@G:CEEF@G7?C3G>CA;BEBCA<GDA;GFA:DA>F;
years’. 3FD@G)D7EF@GDAG DD@0G.22+*
For the adoption of any freeze lining concept, half
measures are entirely unacceptable. The achievement of just
a partial and/or periodic freeze lining will prove unsuccessful
and present a considerably more dangerous operating
condition than a traditional insulating tap-hole design
concept.
The first technique crucial to tap-hole refractory longevity
is the ability to create and retain a protective accretion freeze
lining or skull (Eden et al., 2001), as tap-hole performance is
greatly compromised by operating in the partial or substantial
absence of a stable accretion freeze- lining, which is
described as a ‘no-skull’ condition (Stokman et al., 2004).
Accretion freeze lining thickness has already been shown to
be enhanced by placing refractories of higher conductivity in
actively cooled furnace-lining systems, with the resulting
colder refractory presenting fundamentally more resistance to
attack by a number of wear mechanisms, depending on the
temperature of onset of thermomechanical or chemical attack /B98@FG$-">:F5DEB>G;@D3BA9GC7GE:FG:C?;BA9G<=D>FG7C@GBAF>EF;G58;GBAEC
by a given mechanism (Table II). E:FG
/G)D7EF@G'<8>:B1DGet al.0G+##*
L

 
       VOLUME 116  471
The tap-hole — key to furnace performance
increasing the tap-hole length to develop the ‘mushroom’ graphite sleeve inside an insulating carbon tap-block (Figure
then include there being sufficient tap-hole clay sintering 6). This concept, intriguingly, is the converse of placing
time in the holding space and the specific characteristics of insulating refractory oxide inside graphite, reported as a
the clay during and after heating and sintering. ‘Mushroom’ preferred option for ironmaking BFs.
stability can be adversely affected by the ‘floating’ of an Hearn and co-workers (1998) describe the reasons for
ironmaking BF ‘deadman’, especially if it is physically this as follows: the end hot-face of the graphite insert is
connected to the back of the ‘mushroom’ (Van Laar, 2001). protected by a carbon tap-block, while the cold-face is
Water leaks are also reported to cause a ‘mushroom’, a frozen protected by a removable carbon ‘mickey’ block, which can be
skull, and lining damage (Van Laar et al., 2003; Van Laar, replaced if damaged by either drilling or oxygen lancing, to
2001). secure a flat mating surface against which the mudgun can
The necessary condition of a ‘holding space covered by a more effectively close without excessive tap-hole clay bypass.
coke column’ may well explain why a protective tap-hole clay During tapping the graphite absorbs the tap heat, which the
‘mushroom’ is routinely reported only for ironmaking BFs. In outer annulus carbon tap-block of lower thermal conductivity
non-ferrous processing coke is absent (or substantially cannot transmit as effectively, so ensuring a hot tap-hole
absent), so the necessary requirement of a coke column to with improved flow rates. The heat retained in the graphite
cover tap-hole clay in the holding space is missing. Moreover, sleeve after tapping and immediately following tap-hole
as we describe later, certainly in electric smelting of PGM closure by the mudgun aids tap-hole clay baking. At the next
mattes, matte superheat is so high (as much as 650°C, Table tap, a 45 mm diameter hole is drilled through the baked
I) that tap-hole clay injected into matte appears to react near- taphole clay core to create a tap-hole clay annulus inside the
instantaneously, with the release of gas and extreme graphite sleeve that affords some protection against its
turbulence, so that a tap-hole clay-based ‘mushroom’ cannot coming into direct contact with the molten tap stream.
be stabilized. Obviously, the tap-hole clay can erode with time. With the
While a coke bed is a well-reported feature of ferroalloy removal of the front ‘mickey’ carbon block, the graphite
smelting (Nelson, 2014), it remains local to the electrode tips. sleeve can be core-drilled out and both items replaced to
The extension of the coke bed to the furnace tap-hole – a effect a taphole repair. An additional tap-hole repair design
necessary condition of the proposed mechanism of feature involves splitting in two and gluing the carbon tap-
‘mushroom’ development – would almost certainly result in a block (which contains the graphite sleeve) with carbon paste
condition too conductive for effective electrical power input. A rammed to close the gap between it and the adjacent furnace
genuine ‘mushroom’, at least in the equivalent sense to that sidewall lining, a measure that allows for easier removal with
of an ironmaking BF, therefore seems improbable. At best, less peripheral lining damage during replacement in planned
some extent of tap-hole clay ‘self-lining’, but not a maintenance (Duncanson and Sylven, 2011; Coetzee and
‘mushroom’, is depicted in ferroalloy electric SAFs (Ishitobi et Sylven, 2010; Coetzee et al., 2010).
al., 2010). Some Mn (Ishitobi et al., 2010) and DC arc Cr (Sager et
al., 2010) ferroalloy furnaces make use of inserted water-
) "(##"'%(!"#)')$& cooled copper components on both metal and slag tap-blocks,
The ironmaking BF tap-hole refractory list fairly represents components that range from top lintel to ‘inverted-U’ designs,
the experience in Cr, Mn, and Si ferroalloys, one of an to cool the graphite (advantage of less wetting by slag) or
increasing general trend towards the use of materials of microporous carbon (if dissolution and erosion of graphite by
higher thermal conductivity, and to what is colloquially the metal tapping stream prove too aggressive) tap-blocks.
known in the industry as ‘freeze linings’. For traditional $$&%)&$%' (%) ""#)'%(! #"')$&
insulating (especially large) furnace designs, just 2–6 years
Quite different, though, are the more intensely cooled tap-
of furnace lining life on Cr and Mn ferroalloys are commonly
block designs on blister Cu (Henning et al., 2011; Marx et al.,
reported (De Kievit et al., 2004; Van der Walt, 1986; Coetzee
2005; George-Kennedy et al., 2005; George, 2002; Zhou and
and Sylven, 2010; Coetzee et al., 2010), with one slag tap-
Sun 2013; Newman and Weaver, 2002; pers. comm. 1999,
hole life reported to be as short as 2 months (Van der Walt,
2003) and non-autogenous processes requiring electric
1986). However, longer furnace lifetimes of 10–15 year have
been achieved on traditional insulating linings in Japan.
Generally, Cr and Mn ferroalloy SAFs have made use of only
refractory alumina tap-blocks, silicon carbide tap-blocks
surrounded by alumina, carbon, or microporous carbon
blocks.
This supports a progression from more insulating
refractories (refractory oxide castable and brick, carbon-
based ram or Söderberg paste), to carbon blocks of
intermediate thermal conductivity and even more thermally
conductive semi-graphites and graphites. The latter designs
have delivered in excess of 20 years’ lining life on some large
/B98@FG- DEF<EG7F@@CD??C1GED=6:C?FG;F<B9AGECG4FGBA>C@=C@DEF;GBAECG"/
Mn ferroalloys furnaces (Van der Walt, 1986; Hearn et al.,
7@FF!FG?BABA90GBA>C@=C@DEBA9G@F=?D>FD4?FG>D@4CAG%5B>,F1&G4@B>,GCAG>C?;
1998). 7D>F0G@F=?D>FD4?FG9@D=:BEFG<?FF(FGBA<B;FGE3C6=BF>FG>D@4CAGED=64?C>,0
An emerging trend is of an additional composite >D@4CAG@D55F;GD9DBA<EG<B;FG>D@4CAGED=64?C>,<0G3BE:G>D@4CAG4?C>,G:CE6
refractory variant involving use of a thermally conductive 7D>FGDEGE:FG4D>,GC7GE:FGED=6:C?FG)D7EF@G 8A>DA<CAGDA;G"1?(FA0G.2++*
L

472 VOLUME 116


 
      
The tap-hole — key to furnace performance
smelting, such as Ni and Co ferroalloy (Henning et al., 2010; In Pb bullion smelting (Veenstra et al., 1997; pers. comm.
Nelson et al., 2004, 2007; Walker et al., 2009; And , 1985; 1999, 2003), blister copper (Henning et al., 2011; pers.
Voermann et al., 2010; pers. comm. 1999, 2003), base metal, comm. 1999), and PGM matte ACP top submerged-lance
and PGM matte furnaces (Cameron et al., 1995; Shaw et al., converting (Nelson et al., 2006; pers. comm. 2003), circular
2012; Hundermark et al., 2014; Nolet, 2014; pers. comm. copper tap-blocks have also been used, with both annular
1999, 2003, 2010). These almost universally adopt water- graphite and silicon carbide inserts, or silicon carbide, high
cooled copper tap-blocks of rectangular shape: three-sided alumina, or graphite tapping module bricks.
(inverted U-shape, so there is no water-cooled copper below So whereas ironmaking BF superheats of 350°C may
the tapping channel), four-sided ‘dogbox’ (Figure 14; Nelson seem challenging to copper-cooled operations, they are only
et al., 2007), or high-intensity one-piece waffle cooler copper half the matte superheats experienced on the highest
tap-block designs (Figure 7 and Figure 8). Some are intensity non-ferrous operations. Consider also the
equipped with pin cooling (with inverted-U water passages significantly lower melting temperatures of many mattes
[Henning et al., 2010]—Figure 9). (<950°C, Table I) and this effectively makes it impossible to
These copper coolers are lined internally with a square develop any protective matte freeze lining, even when using
configuration of surround bricks, usually made of magnesia higher cooling water flow rates (but still short of those
(graphite was apparently also trialed successfully in nickel legislated for designation as pressure vessels).
matte smelting [Cameron et al., 1995], but was reported to Notwithstanding this limitation, Ni and PGM mattes also
have been discontinued), containing internal tapping module have a greater solubility for copper than do iron and steel,
refractory bricks through which the tapping channel runs blister copper, and copper mattes; so additionally they have a
(Figure 7, Figure 8, Figure 12 and Figure 14). The latter greater driving force for the chemical dissolution, not merely
comprises refractories that vary with commodity: almost melting, of copper.
exclusively pitch-impregnated magnesia in Ni ferroalloys As we have described earlier, in such a harsh
(Nelson et al., 2007; pers. comm. 1999, 2003), magnesia- pyrometallurgical processing environment the consequence of
chrome in blister Cu or matte (Cameron et al., 1995; Nolet, a superheated matte/blister Cu ‘hit’, or lancing a water-cooled
2014; George- Kennedy et al., 2005; pers. comm. 1999, tap-block (George-Kennedy et al., 2005) and tap-hole failure
2003), or alumina-chrome in PGM mattes (Nolet, 2014; pers. is extreme. It can occur rapidly with a near-identical sequence
comm. 1999, 2003). Both graphite and silicon carbide have
been trialed in matte smelting (Cameron et al., 1995; pers.
comm. 1999, 2003).
For Pb bullion (temperatures of 800–1100°C tapping,
with 700°C drossing) (Veenstra et al., 1997; pers. comm.
1999, 2003) and PGM matte processes (Shaw et al., 2012;
Hundermark et al., 2014; Nolet, 2014; pers. comm. 1999,
2003), process superheats are high (Table I). Specifically for
the latter, process temperatures are elevated to the extent that
the potential for corrosion of magnesia chrome refractory by
PGM matte above 1500°C has recently been investigated
(Lange et al., 2014). Good evidence of expected significant
matte penetration and signs of FeO and MgO corrosion
products have been found, but not as yet a CrS product
suggested by any proposed mechanism. This suggests a /B98@FG-">:F5DEB>GC7GDG:B9:6BAEFA<BE1G3DEF@6>CC?F;G>C==F@G5DEEF
potential for high refractory wear rates with exceptionally ED=64?C>,G<1<EF50G3BE:G3DEF@6>CC?F;G>C==F@G48??AC<FGFEFA<BCAGDA;
3DEF@6>CC?F;G>C==F@G7D>F=?DEF0G4CE:GC8E<B;FG78@AD>FGA?1GE:FG>C==F@
high matte superheats (approaching 650°C, Table I).
7D>F=?DEFG:D<G%BA(F@EF;6&G3DEF@G=D<<D9F<0G3:F@FG3DEF@GB<GD4<FAEGCA
E:FG8A;F@<B;FGC7GE:FGED==BA9G>:DAAF?G

/B98@FG-B9:6BAEFA<BE1G>C5=C<BEFG3DEF@6>CC?F;GDA;G@F7@D>EC@BF<GED=6 /B98@FG#-DEF@6>CC?F;G>C==F@G=BAGDA;G%BA(F@EF;6&G5FED?GED=64?C>,
4?C>,G7C@G4?B<EF@G>C==F@G)D7EF@GFC@9F6 FAAF;1Get al.0G.22$* ;F<B9AG)D7EF@GFAABA9 )%'(#0G.2+2*
L

 
       VOLUME 116  473
The tap-hole — key to furnace performance
of events, regardless of furnace size (Nelson et al., 2006). Modelling has similarly motivated the deepening of the
The potential for catastrophic cooler failure and/or furnace metal bath of a circular HC Mn ferroalloy SAF (but still with a
refractory breakout (Zhou and Sun, 2013; Newman and horizontal tapping channel, presumably in part because of
Weaver, 2002) within, most commonly, a few minutes of the absence of anything equivalent to a ‘sitting deadman’) by
mudgun closure, is high (Hundermark et al., 2014). A removing a full course of carbon blocks to reduce the
breakout following mudgun closure has even prompted one peripheral liquid flow velocity along the wall to a draining
PGM producer to resort to drilling and lancing, but to closing tap-hole (Ishitobi et al., 2010). The reduced peripheral flow
tap-holes with clay manually using stopper rods rather than induced by the deepening of the hearth reduced metal tapping
mudguns (Coetzee, 2006). temperatures by an average of 40°C (to 1350°C), despite the
uprating of the transformer capacity to permit a simultaneous
()!#(%)'(&' ) (%" '$&) %')$& increase of the electrode current by 25 kA to raise the
External faceplates are important for providing a ‘perfectly’ average power load at night by 2.3 MW, combined with
flat vertical mating face for the mudgun to engage the tapping operation at a higher coke loading to allow approach to metal
channel (for accuracy of tap-hole clay quantity injected into carbon saturation (so limiting wear by dissolution of the
the tapping channel, so ensuring minimal bypass), coupled carbon lining). Deepening of another Japanese HC Mn
with the refractory insert, for providing a mechanism to help ferroalloy furnace gave benefits of marginally increased
secure tight joints along the length of the tapping channel to power input, faster tapping, and increased productivity
minimize infiltration and gas leakage (Eden et al., 2001), (Nishi, 2007). On Si ferroalloy SAFs (Kadkhodabeigi et al.,
and to help prevent the entire tapping channel lining from 2011), where metal drains through a porous bed of solids to
dislodging and ‘tapping’ out of the furnace lining owing to the tap-hole, crater pressure and bed permeability
internal furnace pressure (comprising both internal operating significantly influence the rate of drainage of metals to and
pressure and any blast pressure and hydrostatic head). The through the tap-hole.
last of these incidents has apparently been experienced in the In the largest rectangular six-in-line PGM matte smelting
past on a Ni matte EF. furnace, the matte inventory can exceed 600 t, with contained
Thermal fatigue cracking or direct matte attack of water- metal value exceeding US$50 million. Furnace deepening will
cooled copper faceplates, typically associated with matte come at a greater cost. Fortunately, with a combination of
splashing during tap-hole plugging, presents a risk of water periodic and low-volume matte tapping (< 20% matte fall)
leaks. Sacrificial refractory or metallic cover plates have been through an end-wall of an inverted arch hearth design, in a
used to address this risk (Cameron et al., 1995), with the rectangular furnace configuration, tap-hole wear has recently
introduction of inverted-U water-cooled pipe arrangements to been predictable even at operations exceeding 60 MW power
secure the absence of water-cooling directly below the input (Hundermark et al., 2014). With a circular furnace
tapping channel, a measure that better mitigates the risk of
configuration more conducive to the development of
matte making contact with water.
circumferential flow along the sidewall to a draining matte
tap-hole, especially when the matte tap-hole is located almost
(!"#)'$&#$&(%$"&'(&'(%$)')( %'!')$&
on the top of the skew line of the hearth invert, it is not
Tap-holes are normally designed with a horizontal or vertical inconceivable that conditions for accelerated matte tap-hole
(e.g. EBT) orientation. The notable exception is the near-
wear could develop, even at far lower inputs of power.
universal implementation of inclined tap-holes (approx. 10°)
on ironmaking BFs. Modelling has shown that inclined tap-
holes, coupled with longer tapping channels and deeper
'D=G<F8FA>BA9
hearth sumps (the minimum sump depth is 20% of the A variety of strategies are adopted, depending largely on
hearth diameter [Jameson et al., 1999; Gudenau et al., productivity requirements, number and layout of tap-holes,
1988]) that drain liquid deeper in the furnace (further from and process conditions. For single tap-holes processing dual
the sidewalls), lower liquid velocities (and resultant wall metal-slag mixtures, total reliance is placed on the availability
shear stress and wear) both below the tap-hole and at the of the sole tap-hole. Such tapping systems are especially
wall periphery (that otherwise lead to undercutting and so- common in Cr and Mn ferroalloy SAFs, which may emphasize
called ‘elephant’s foot’ wear) (Stokman et al., 2004; Eden et the importance of the tapping stream superheat (average-to-
al., 2001; Smith et al., 2005; Dash et al., 2004; Jameson et maximum heat flux 1–10 kW/m2 [De Kievit et al., 2004;
al., 1999; Post et al., 2003). The localized higher velocities Table I]) over absolute temperature in describing an onerous
below the tap-hole are attributed to the draining of liquid process condition.
down past the ‘mushroom’ (Figure 4, Van Laar, 2001). The That said, a still impressive 5 700 t/d HM in a campaign
higher peripheral velocities at the wall periphery are more a life of 13 years at the time of reporting was achieved from a
function of draining through and around a ‘deadman’ (Dash single taphole BF operation (Ballewski et al., 2001). Similarly
et al., 2004; Jameson et al., 1999; Tanzil et al., 2001). the Mitsubishi Continuous Process for copper relies on
Optimum tap-hole inclination was modelled as 15° (Dash et continuous liquid flow down heated launders from smelting,
al., 2004). Tapping conditions are further noted to distort to slag cleaning, to converting, and to anode refining
fluid flow to the extent that, towards the end of tapping, the furnaces, this being effected through a combination of
slag is lowest in the vicinity of the draining tap-hole, inclined furnace overflow, skimming, and siphon tapping
to its highest at the opposite side of the BF (Post et al., 2003; arrangements, at overall availabilities exceeding 92%
Tanzil et al., 2001). We are aware of at least one high-carbon (Matsutani, n.d.). These examples illustrate what is possible
(HC) Cr ferroalloy furnace equipped with a declined tap-hole. with superior tap-hole management and tapping practices.
L

474 VOLUME 116


 
      
The tap-hole — key to furnace performance
"&)%$)'$&$$(#'%(!!$&'! (%$) gas (Nightingale et al., 2001; Tanzil et al., 2001). In contrast,
the requirement on the multiple tap-hole, lower metal/matte
Consecutive tapping on an individual tap-hole is a common
fall (<20%) Ni ferroalloy and matte furnaces is primarily to
traditional practice on several ironmaking BFs (Rüther, 1988;
secure maximum tap-hole and furnace reliability. This is
Cassini, 2001), ferroalloy, and matte-smelting operations.
especially true of high-intensity PGM matte furnaces, with
Even on two-tap-hole BFs, tapping campaigns of 4 days to 3
their onerous matte superheat, T approx. 650°C, that
weeks are reported (Rüther, 1988). Matte tap-hole
imposes integrity challenges on even the most intensely
temperature trends in Ni matte smelting clearly demonstrate
water-cooled, refractory-lined copper tap-hole.
the accumulation of heat in the tap-hole refractory when taps
On the highest intensity of these operations, even with
are in close succession (Cameron et al., 1995; Figure 10).
less frequent matte tapping events, the practice generally is to
Similar rising temperature trends with tapping have been
alternate tapping between the available tap-holes in order to
observed in PGM matte smelting (Gerritsen et al., 2009;
give the tap-holes maximum ‘recovery’ time to lower tap-hole
Figure 11). With an ironmaking BF interpretation this could
temperatures between taps. This is reported (Eden et al.,
possibly be considered desirable for promoting tap-hole clay
2001; Mitsui et al., 1988; Entwistle, 2001; Cameron et al.,
baking and sintering. However, in the more intensely
1995; Gerritsen et al., 2009) and has been modelled in the
superheated matte-only tap-hole environment this is rather
BF (Ko et al., 2008). The merits of such an approach,
interpreted to imply that a resting or recovery period of no
originally diagnosed from scrutinizing well-instrumented
tapping is called for, to help lower refractory temperatures
copper tap-block and cooling water temperature tapping
and re-establish improved accretion, as evidently occurred on
trends, are presented using the latest fibre-optic temperature
the tap-hole on the furnace in Figure 10.
measurement trends available in PGM matte smelting (see
#%) &(%$&'%(!"#)'! (%$) section on Advanced tap-hole monitoring).
At first glance an alternating tap-hole practice would
This variant, also described as ‘side-to-side’ casting
appear to complicate the timing of minor routine, monthly
(Petruccelli et al., 2003), is certainly the norm for achieving
planned tap-hole maintenance activities (Nolet, 2014).
the highest of productivities through optimal tap-hole
condition, consistent operability, and reliable availabilities; it However, it should be appreciated that, despite such diligent
also best supports preventative tap-hole maintenance. This is monthly repairs and essentially slag-free tapping, process
true of two tap-holes (Petruccelli et al., 2003) and tap-hole conditions remain so onerous that all but the hot-face matte
pairs on four-tap-hole ironmaking BFs (Rüther, 1988; tapping module bricks have to be replaced roughly every
Steigauf and Storm, 2001); 2–8 metal-only and 2–6 slag-only quarter to secure incident-free tapping, good tap-hole
tap-holes on blister Cu and ferroalloy furnaces (George,
2002; Zhou and Sun, 2003; Newman and Weaver, 2002;
George- Kennedy et al., 2005; Nelson et al., 2004, 2007;
Walker et al., 2009; pers. comm. 1999, 2003); and up to
three matte- and three slag-only tap-holes on base metal and
PGM matte EFs (Nolet, 2014; Nelson et al., 2006; pers.
comm. 1999, 2003). It includes ironmaking BF variants
described as ‘back-to-back’ or ‘mother-daughter’ tapping
(Irons, 2001; Cassini, 2001), where a pair of taps is made
before alternating tap-holes. In the case of the ironmaking
BF, this practice of a pair of taps is usually in response to
suboptimal conditions, such as inadequate draining or
persistent taps of short duration.
A detrimental feature reported for alternating tapping on /B98@FG+2-'1=B>D?G:FDEG?CD;GDA;GD>>858?DEBCAGC7G:FDEG3BE:GED=<GBA
BFs, where a zone of low permeability exists between tap- >?C<FG<8>>F<<BCAG)D7EF@GD5F@CAGet al.0G+##$*
holes, is the potential for the slag level to rise due to
excessive pressure loss, which disrupts bosh gas flow (Iida et
al., 2009; Shao, 2013; Shao and Saxen, 2011, 2013a,
2013b). Slag levels could conceivably fluctuate on SAFs
similarly, owing to the presence of less permeable zones. Iida
and co-workers (2009) recommend enlarging the tap-hole
diameter (by approx. 10%) as the best remedy to alleviating
this issue.
While operating at a still impressive HM superheat, T
approx. 350°C, the focus on the BF is largely HM
productivity-driven, with up to 75% metal fall and daily
targets exceeding 10 000 t HM, thus demanding the most
effective and efficient tapping with reliable operability. Most
operators appear to seek to operate somewhere close to a
‘dry’ hearth condition (De Pagter and Molenaar, 2001), in
which hot metal and slag levels in the hearth are kept as low /B98@FG++-'@FA;GC7GED=6:C?FGEF5=F@DE8@FG@B<FG3BE:G>CA<F>8EB(FGED==BA9
as possible (Van Laar et al., 2003), but without escape of hot CAGDGED=6:C?FG)D7EF@GF@@BE<FAGet al.0G.22#*
L

 
       VOLUME 116  475
The tap-hole — key to furnace performance
condition, and ultimately furnace integrity and longevity. To available is limited (e.g. owing to planned maintenance), an
undertake such a deep tap-hole repair, tap-hole temperatures effective solution involves closing on lazy-flowing slag with
and safety dictate that the furnace power needs to be lowered the mudgun, and shortly thereafter re-drilling the slag tap-
for the duration of the repair. So in fact a simultaneous repair hole open again (exposure of drill bits to slag only is far less
of all matte tap-holes by a team of masons on a furnace at aggressive than exposure to metal or matte). This can easily
lowered furnace power actually minimizes the impact on double the initial tapping rate on a ‘slow’ slag tap-hole.
overall furnace utilization. Closure on flowing slag is crucial to this operation,
Also, it should be clarified that in high-intensity PGM because it ensures easy re-drilling of tap-hole clay only to
matte smelting the ‘as-low-as-possible’ liquid matte and slag open the slag-tapping channel. In the event where the flow
levels of the BF ‘dry hearth’ operation are definitely not from a slag tap-hole has been allowed to stop, even with an
sought, nor considered desirable. Considering first the overall attempted mudgun closure, an adequate plug of tap-hole clay
liquid level, one finds that generally too high a pressure head to the inner hot-face cannot be secured. When re-drilling is
is not sought, because it promotes an increased rate of attempted, solidified slag is quickly encountered, which
tapping and increases the potential for matte infiltration of impedes the drill and can cause skew drilling – potentially
the furnace lining. Specifically, one also does not seek too towards a water-cooled copper cooler! So somewhat
high a matte level, for fear of exposing the effective slag-line, paradoxically, to be safer, oxygen lancing with its ability to
water-cooled copper waffle coolers to a greater risk of making ‘cut’ open, and so straighten, the solidified slag tapping
contact with superheated matte. While the waffle cooler channel then becomes necessary to re-open the slag tap-hole.
design reportedly (Trapani et al., 2002; Merry et al., 2000)
caters for metal contact of copper waffle coolers in Ni (Nelson 'D=6:C?FGC=FABA9
et al., 2007) and Co (Nelson et al., 2004) ferroalloy It is essential to be able to ‘quickly and certainly open the
processing, contact by matte, especially superheated matte, tap-hole whenever required’ (Tanzil et al., 2001).
can rapidly lead to catastrophic failure. Discounting the most primitive past practices of ‘pricking’
However, this still does not warrant seeking the lowest or ‘excavating’ the tap-hole open, a wide range of tap-hole
possible matte level. This is because the matte-only tap-hole opening methods are adopted (Ballewski et al., 2001),
is especially configured to be refractory oxide-lined, with including:
generally good corrosion resistance to matte, but with
® Manual oxygen lancing, suggested near universally to
decidedly poor corrosion resistance to acidic slags (> 50%
be minimized to < 1% of taps (Jameson et al., 1999), or
SiO2 content). Indiscriminate lowering of the matte level
for ‘emergency only’ on ironmaking BFs (Ballewski et
would therefore not only expose the tap-hole to the risk of
al., 2001). This practice has led directly to a reported
‘slagging’ by the hotter slag, but would accelerate corrosion,
blister tap-hole failure and resulting explosion on at
and ultimately wear, of the refractory lining. A target
least one site (George-Kennedy et al., 2005), and yet is
minimum matte level is therefore simultaneously sought with
still adopted as the primary means of tap-hole opening
matte operated below the maximum matte level permissible.
on 36% of PGM matte furnaces (Nolet, 2014)
In respect of the slag level, the absolute minimum furnace
® Automated or robotic oxygen lancing (pers. comm.,
slag level is controlled by its interface with matte. Operation
2010)4
around the slag tap-hole, located typically approximately 1 m
® A soaking bar technique5
above the matte tap-hole (Table I), represents the lowest
® Conventional pneumatic drilling (air)
overall pressure head condition on the matte, which is
® Improved pneumatic drilling (nitrogen and/or water-
beneficial. However, at the highest smelting rates with < 20%
mist-bit cooling)
matte fall, slag make becomes significant, which requires
® Hydraulic drilling (nitrogen and/or water-mist-bit
near-continuous tapping in contrast to periodic batch matte
cooling)
tapping. With the slag level only at the level of the slag tap-
hole, the pressure head is simply inadequate for slag tapping
rates to be acceptable. So a practical minimum operating slag
level exists, above which slag tapping rates are adequate for
achieving an efficient rate of slag drainage (even if multiple
slag tap-holes are open). 4See also http://www.mirs.cl/img/video/punzado_descarga_escoria_
hornos.wmv
Finally, the maximum permissible top of slag level is 5The soaking bar practice found favour in iron BF tapping as an emerging
designed relative to the slag tap-hole. This measure primarily development to replace tap-hole drilling in the 1980s. It involved
ensures that superheated slag does not rise above the zone of pushing/hammering a 50 mm bar through the mud in the tapping channel.
The bar promised to provide improved thermal conductivity from the inner
sound crucible containment below the top of the copper hearth up the tapping channel, which helped bake and sinter the tap-hole
coolers, but also limits excess pressure head at both the slag clay better. To open the tap-hole, the bar was reverse-hammered out of
the tapping channel, now of well-defined dimension, and with the promise
tap-hole and the underlying matte tap-hole.
of no risk of skew drilling or oxygen lancing damage. This practice,
however, had fallen out of favour by the 1990s, because it required
#('%(!!$& (1) time-consuming predrilling to assist with the soaking-bar insertion and
(2) an assessment of the all-critical drill depth. Furthermore, matching this
Where consecutive tapping practice has indeed found depth to an optimal tap-hole-clay addition was difficult, shorter tap-hole-
nonferrous application is during ‘slow’ slag tapping on both clay curing times increased the risk of a tap-hole re-opening, and
Ni ferroalloy and PGM matte smelters. The slag tap-hole has hammering in and removing the bar damaged the tap-hole and
‘mushroom’ in other ways (Jameson et al., 1999; Van Ikelen et al., 2000;
a tendency to open fast and then the tapping rate declines Steigauf and Storm, 2001; Ballewski et al., 2001; Entwistle, 2001;
with time. In situations where the number of slag tap-holes Östlund, 2001)
L

476 VOLUME 116


 
      
The tap-hole — key to furnace performance
® Combination pneumatic drilling (without opening) and Once a liquid has penetrated a refractory, corrosion by the
deliberate lancing of the last remaining metal/matte infiltrating liquid becomes possible. Campbell and co-workers
plug. (2002) describe corrosion as a ‘cooking time’ to illustrate that
It is worth noting that to avoid contamination by iron or its rate relates to how long a penetrated refractory has been
other elements, metallurgical-grade silicon tapping requires a at a temperature that supports reaction. Furthermore, as
variety of alternative tools to open a tap-hole and maintain corrosion rate conforms to Arrhenius’s Law, an exponential
the flow of metal. These alternatives include an electric (as opposed to linear) scale of temperature is required to
stinger (connected to a busbar system from the furnace predict the increase in the rate of corrosion with temperature.
transformers), a kiln gun (Guthrie, 1992)6, steel and graphite Once a refractory has been penetrated and further
lances, wooden poles, and graphite bott tools (Szymkowski weakened by corrosion, erosion becomes possible if the shear
and Bultitude-Paull, 1992). stress,  = (dv/dy) induced by the liquid flow through the
tap-hole is sufficient to remove refractory. Once again,
(!!$&' (%) temperature affects liquid viscosity, whereas the rate of
A primary requirement of tapping is to reliably secure the tapping affects the velocity gradient (dv/dy). Estimated
desired rate of furnace products. Thus, it is important to tapping velocities of 1–5 m/s suggest that the applied shear
establish the factors influencing tapping rate. Guthrie (1992), force is a few orders of magnitude lower than the hot
applying Bernoulli’s equation, provides a useful estimate of modulus of rupture of most refractories. So it is well-argued
· that tap-hole refractory erosion cannot occur until the
tapping rate, m = CD(d2/4)(2gH)½, through a tap-hole of
diameter d, where, CD is a discharge coefficient (approx. 0.9), refractory structure has somehow first been weakened by
g is the gravitational acceleration constant, and H is the liquid penetration and corrosion (Campbell et al., 2002).
effective liquid head of the phase being tapped, with a phase In PGM matte tap-holes an annulus of tap-hole clay does
of density . not appear to persist in lining the tapping module refractories
Mitsui and co-workers (1988), combining Bernoulli’s and (Figure 12). However, the same (low) velocities may possibly
Darcy- Weisbach’s equations, estimated the iron BF tapping provide a shear force that is in excess of the hot modulus of
· rupture of poorly baked/sintered tap-hole clay. So in
rates as m = (d2/4)(2[P/ + gH]/[1 + l/d])½, thereby
including a correction for the tapping-channel length (l). operations that critically depend on a ‘maintainable’ baked
This yields typical iron BF tapping rates of 7 t/min (approx. and sintered annulus of tap-hole clay to line the tapping
10 000 t/day on a near-continuous tapping basis) and liquid channel to protect the tap-hole refractory (e.g. especially
tapping velocities of 5 m/s in tap-holes of 70 mm diameter by when combined tapping of more corrosive slag, as in
3.5 m length. Both approaches show that tap-hole geometry ironmaking BFs), far more attention should be paid to the
strongly influences tapping rate (with velocities of up to issue of tap-hole clay sintering and erosion-resistance
8 m/s recorded [He et al., 2001; Atland and Grabietz, 2001]), properties (Mitsui et al., 1988).
primarily through the tap-hole diameter. The second equation
suggests tap-hole length as the next most significant
influence.
In the case of Si ferroalloy SAFs (Kadkhodabeigi et al.,
2011), where metal must drain through a permeable bed of
solids to the tap-hole, the height of liquid metal influences
the onset of gas breakthrough to the tap-hole and the
concomitant sudden drop in tapping rate, but exerts less
influence than crater pressure and bed permeability on the
initial tapping flow rate.

(!"#)' )( ')(&$
Given a dominant influence of tap-hole dimensions on
/B98@FG+.-DEEFGED==BA9G5C;8?FG4@B>,G3BE:G5DEEFG>C@F GACGF(B;FA>FGC7
tapping rate, it is instructive to consider factors contributing DGED=6:C?FG>?D1GDAA8?8< G=FAFE@DEF;G;FA<FG4@BEE?F6!CAFGDAA8?8<GBA<B;F
to tap-hole wear (Figure 12), which are elegantly 4@B>,
summarized by three sequential steps: penetration, corrosion,
and erosion (Figure 13; Campbell et al., 2002).
The first step in refractory wear involves the penetration
of refractory, the rate of which, upen, can be described by a
capillary-force-driven flow according to r
cos/4lp, where r
is the capillary (pore) radius,
is surface tension,  is the
contact angle, lp is penetration depth, and  is liquid
viscosity. The last property (viscosity) is related inversely to
process temperature.

/B98@FG+-F=@F<FAEDEBCAGC7G@F7@D>EC@1G3FD@G5F>:DAB<5<G)D7EF@
6See also http://www.youtube.com/watch?v=u_4cEWTzQnI D5=4F??G)%'(#0G.22.*
L

 
       VOLUME 116  477
The tap-hole — key to furnace performance
Generally, the potential adverse influences of suboptimal (!"#)' $##$&' ) $ ))&%
tapping velocities are:
The requirements to control and optimize the rate of drainage
® Too slow tapping—limits tapped production; delays to the tap-hole (to reduce liquid velocities and wear of the
liquid drainage, which may potentially be unsafe if furnace lining) and the associated tapping rate through it (a
critical furnace levels are threatened (e.g. matte controlled liquid tap with stable post-tap-hole conditions)
encroachment to near the vicinity of copper coolers, or impose a need to maintain a constant and optimal tap-hole
slag overflow over the design maximum crucible length and smooth shape (Van Ikelen et al., 2000). The
containment height) length is usually as long as is practicably achievable, while
® Too fast tapping—induces loss of control, thereby one maintains a near-cylindrical channel shape of defined
creating unsafe tapping and post-tap-hole conditions; diameter. In reality, some extent of fluting towards the hot-
in the extreme, and only then, promotes tapping face (conveniently modelled as a cone [Van Ikelen et al.,
channel and furnace lining erosion. 2000; Nightingale et al., 2001]) with erosion at the hot-face
These influences may have more adverse consequences (conveniently modelled as a paraboloid to represent a zone
than erosion does. for ‘mushroom’ development [Van Ikelen et al., 2000;
Nightingale et al., 2001]) has been inferred from tapping

$##$&'! (%$) channel temperatures, drill depths, and their distributions


Owing to the potential for oxygen-induced lancing damage to (Mitsui et al. 1988; Van Ikelen et al., 2000; Nightingale et al.
tap-holes, the vast majority of operations seek to practise 2001).
drilling the tap-hole open. This typically includes sacrificing In ironmaking operations with lower metal fall (a high
the drill bit and, potentially, the drill rod. In at least one slag ratio of lower density) it is argued that ‘the decision for
Japanese Mn ferroalloy operation, to conserve costly drill bits, diameter and tapping practice must be focused on slag’
the operator withdraws the drill as soon as metal is expected (Brunnbauer et al., 2001). This highlights the role of reliable
to be encountered, places a sacrificial crimped steel pipe over drilling, as it represents the primary means for controlling
the drill bit, and then drills the hole open. This protects the tap-hole diameter.
drill bit enough to permit re-use.
(!"#)' $##$&') $!)&%'(&'"&% "#
" $&(%$"&' $##$&'(&'!#'")&'#(&$& Owing to the excessive risk of skew drilling (directly
! (%$) contributing to similarly skew oxygen lancing in combination
drilling and ‘plug’-lancing practice), especially to operations
On most alloy-only and matte-only tap-holes operated in the
with water-cooled copper tap-blocks, practice typically
substantial absence of any tap-hole hot-face ‘mushroom’, a
requires that the accurate alignment (to surveyed tap-hole
combination of deep drilling followed by ‘plug’ oxygen
centre/s [Estrabillo, 2001]) of mudgun/s and drill/s be
lancing is practised deliberately. The aim is to drill through
checked and, if necessary, recalibrated at the start of each
the tap-hole clay as (consistently) deep as possible (700–
shift (Irons, 2001). Tap-hole-centering notches are also
1200 mm, depending on tap-hole design length), until the reported; they locate and indent the tap-hole clay to help keep
drill encounters resistance from a ‘plug’ of metal/matte/ the drill from ‘walking off’ from the centre of the tap-hole
residual entrained slag. Experience indicates that attempts to (Estrabillo, 2001).
drill further through this ‘plug’ often lead to unintended skew In addition, guided and stiff drill rods are essential to
drilling. This measure is particularly hazardous in a water- reducing excessive drill flex and securing a straight, centred
cooled copper tap-block configuration, and often results in tap-hole. Guide systems include automatic travel to within
the drill simply getting stuck in the tapping channel. Even limits, followed by a hydraulic pin, sometimes colloquially
with reverse percussion hammering (Bell et al., 2004), it may called ‘antlers’ (Black and Bobek, 2001), being physically
become impossible to free a stuck drill bit and rod, an positioned down into latch hooks. For drilling 4 m long
outcome that requires the tapper to resort to oxygen lancing ironmaking BF tap-holes (requiring 6 m drill rods), additional
to remove the obstruction. hydraulic rod devices are fixed to the drills to prevent
In combination practice, the drill is then withdrawn, and bending of the drill rods and drilling off the tap-hole axis
the drill length measured accurately (but manually) with a (Ballewski et al., 2001). The undesirable consequence of
gradated drill-T, which simultaneously verifies that the using a less precise suspended rock drill for tap-hole drilling
drilling was not off-centre. Once the drill-hole is confirmed as has been reported previously in a four-piece, water-cooled
being straight, oxygen lancing of the short remaining tapping copper Ni ferroalloy tap-block operation (Nelson et al., 2007;
channel ‘plug’ is then undertaken to open the tap-hole. This Figure 14 and Figure 15).
usually requires a minimum of lancing (less than one lance An encoder that measures the drill position can be
pipe). In this way there is also a lower risk of tappers losing correlated with drill torque (in hydraulic systems – Jameson
the skill of using oxygen lances safely owing to infrequent et al., 1999; Atland and Grabietz, 2001) or drill air-pressure
practice. forward drive (in pneumatic systems – Van Ikelen et al.,
The rationale behind this practice is driven by a decided 2000) and drill speed to determine automatically the start
requirement not to overfill tap-hole clay, through the addition and end of the tapping channel and hence the all-important
of a metered amount of tap-hole clay, which permits tap-hole length (Jameson et al., 1999; Van Ikelen et al.,
operation with a consistent short (as possible) tapping- 2000; Eden et al., 2001; Tanzil et al., 2001; Edwards and
channel ‘plug’ to lance. Hutchinson, 2001; Smith et al., 2005). Drill-time sigma
L

478 VOLUME 116


 
      
The tap-hole — key to furnace performance
2001), where water-mist cooling rates are in the range of 2–5
L/min and typically 4 L/min (Tanzil et al., 2001). Water-mist
cooling systems are reported to have undergone still further
development to overcome disadvantages of increased risk of
drill equipment corrosion (Van Ikelen et al., 2000).
In ferroalloy and matte operations, especially those
equipped with any potentially hydratable magnesia-based
refractory, use of any water would be taboo (in fact even to
the extent that dew-point condensation associated with
liquid-nitrogen cooling to accelerate tapping channel repair is
/B98@FG+-DEF@6>CC?F;G%;C94C&G3BE:G5BAC@G>@D>,<GBAG<8@@C8A;G4@B>,< sometimes a concern). The short drill-bit life is largely
D@C8A;GED==BA9G5C;8?FG4@B>,G3BE:G%6<:D=F;&G>@D>,GDA;GC776>FAE@FG;@B??6
overcome when drilling only tap-hole clay (i.e. deliberately
:C?FG) F?<CAGet al.0G.22*G
not drilling metal/matte/slag) in both metal/matte-only
combination drilling and slag-only drilling open tapping
practices.
Two opposing effects of drilling on the control of tapping
channel diameter are reported. With premature bit wear,
negative fluting of the tapping channel (diameter decreasing
evenly down to the drill rod diameter towards the hot-face)
has been reported (Van Ikelen et al., 2000). Side-cutting
designs capable of cutting during both forward and reverse
drilling have been developed to limit the influence of drill-bit
wear on the resulting drilled diameter (Van Ikelen et al.,
2000). More frequently, though, a bit that fails to retain its
/B98@FG+$-'D==BA9G5C;8?FG4@B>,GGD7EF@G.+#GED=<G8?EB=?FG;@B??BA9<GC7 cutting edge tends to wander, which causes positive fluting to
E:FGED=6:C?FGD@FGC776>FAE@F0GDA;GD@FG>C8=?F;G3BE:G<,F3G?DA>BA9G) F?<CA
the hot-face (Nightingale et al., 2001; Mitsui et al., 1988;
et al.0G.22*G
Tanzil et al., 2001), or a ‘mushrooming’ effect (Estrabillo,
2001; Edwards and Hutchinson, 2001). Traditional rock drill-
(Black and Bobek, 2001) and tap-hole length (Jameson et al., bit designs provide some increased resistance to this, and are
1999) are regarded as benchmark statistics and, with the often preferred (Estrabillo, 2001), despite still requiring drill-
application of statistical process control (SPC), measures with bit replacement every tap on an ironmaking BF. This
which to quantify and effect tap-hole improvements. warrants further clarification: on ironmaking BF tap-holes
the ability to open with ‘one drill-bit for every attempt’ is
   regarded as an achievement (Estrabillo, 2001), with only a
Drill-bit shape and material – carbide (Black and Bobek, 50% success rate reported at one site (Nakamura et al.,
2001; Tanzil et al., 2001; Entwistle, 2001) or heat-resistant 2007), or an average of 1.2 drill bits per tap reported (Atland
Cr-Ni alloy (Atland and Grabietz, 2001) tips are preferred – and Grabietz, 2001). Progression from threaded to bayonet
has been the subject of intense investigation, especially in the drill-rod couplings is reported (Estrabillo, 2001) to limit the
ironmaking BF application (Van Ikelen et al., 2000; Ballewski incidence of drill rods jammed tightly in couplings.
et al., 2001; Black and Bobek, 2001; Brunnbauer et al., The direct consequence of a smooth, straight tapping
2001; Estrabillo, 2001; Entwistle, 2001; Atland and Grabietz, channel is a consistent smooth tapping stream and controlled
2001). The ability to retain a sharp cutting edge so as to cut, post-tap-hole logistics. In contrast, a tapping channel that
rather than hammer, through the tap-hole clay ‘plug’, with has an inner corkscrew shape is reported to induce a rotating
the bit cutting face presented to a debris- and dust-free face and spraying tapping stream (Van Ikelen et al., 2000), an
to drill, is essential (Estrabillo, 2001). Drill-bit diameter is outcome exacerbated by any gas-tracking on a pressurized
controlled usually within the range of 33 mm (Tanzil et al., BF operation. ‘Softer drilling’ (feed-forward pressure < 3 bar)
2001) to 45–65 mm (Steigauf and Storm, 2001; Atland and together with instructions to the operator to ‘let the drill do
Grabietz, 2001). Where hammering is considered important, the work’ and so not try to force the tap-hole open using
an inside bit face that is totally flat (to maximize maximum force, which can bend the drill rod and promote a
transmission of impact energy) is reported (Tanzil et al., corkscrew channel, is reported to lower the incidence of
2001), coupled with transition from spherical to semi- rotating and spraying tapping streams (Van Ikelen et al.,
spherical carbide shapes. 2000).
Air scavenging is typically used to clear the hole, This is remarkably akin to the requirements of successful
providing additionally some cooling of the drill bit to help oxygen lancing: a good tapper tends to use the hot burning
prolong its life (Van Ikelen et al., 2000). Further lance tip (> 2000°C) to progressively cut the tap-hole open in
improvement has involved progressively improving drill-bit a series of small precessing actions to guide the lance ever
cooling (from air, to nitrogen, to water mist) on ironmaking deeper to make a straight tapping channel. An inexperienced
BFs (Eden et al., 2001; Petruccelli et al., 2003; Van Ikelen et tapper, on the other hand, tends to try to force-burn the tap-
al., 2000; Smith et al., 2005; Irons, 2001; Steigauf and hole open by pushing hard on the thin, long and flexible
Storm, 2001; Ballewski et al., 2001; De Pagter and Molenaar, lance pipe, which readily causes it to deflect off-course and
2001; Black and Bobek, 2001; Edwards and Hutchinson, cause damage.
L

 
       VOLUME 116  479
The tap-hole — key to furnace performance
Finally, it is said that ‘a rotating drilling method for (Cassini, 2001) to release CO, CO2, H2, and/or H2O gases. In
opening the tap-hole, without hammering … is expected to high-duty applications, tap-hole clay of low gassing potential
give an improvement of the tapping process’ (Van Ikelen et is therefore a prerequisite, and almost all operators seek an
al., 2000). Similarly, many local ferroalloy and PGM matte anhydrous clay (Abramowitz et al., 1983) or ‘water-free
tap-holes are indeed opened by drill rotating action alone plastic mass’ (Smith et al., 2005).
without hammer action, despite the latter’s usual availability. A perfectly cylindrical 1 m long tapping channel 50 mm in
Even on ironmaking BFs it is suggested that ‘future diameter requires theoretically only 2 L of tap-hole clay to
advancements will be directed toward drilling the tap-hole completely fill it. This increases to 5 L if the tap-hole is worn
without the need for hammering’ (Estrabillo, 2001). on average to 80 mm diameter, by either positive fluting
(exacerbated by any oxygen lancing and/or enlargement by
'D=6:C?FG>?C<8@F bath wear of the tap-hole hot-face) or negative fluting down
It is essential to be able to ‘close the tap-hole with a high the tapping channel. Iida and co-authors (2009) even suggest
degree of certainty that the desired volume of tap-hole clay that tap-hole enlargement occurs typically at a rate of 5.6 ×
has in fact been installed’ (Tanzil et al., 2001), and 10–4 mm/s during tapping (1 × 10–3 mm/s when using ‘poorer
additionally ensure that mudgun retraction does not result in durability tap-hole mix’ [Iida et al. 2009], a practice also
an unplanned tap-hole re-opening. Total elimination of modelled by others [Shao, 2013; Shao and Saxen, 2013b]).
reopening events remains important, even given reported It is quite staggering to compare this addition with the range
improvement from 10 to just one such event per annum by reported for ironmaking BFs – admittedly with tap-hole
2000 on one site (Black and Bobek, 2001). lengths of 1.8–2 m (Edwards and Hutchinson, 2001; Atland
Especially on slag-only closure, stopper bars, water- and Grabietz, 2001), or more usually 2.5–4 m (Irons, 2001) –
cooled ‘rosebuds’, and manual stopper tap-hole clay ‘plugs’ from as little as 10–20 L (Irons, 2001) to 50–120 L (Irons,
remain common in the ferroalloy and non-ferrous industry. 2001; Atland and Grabietz, 2001; Van Laar, 2014;
Slightly more sophisticated variants are used on some of the Nightingale et al., 2001; Jameson et al., 1999; Cassini, 2001)
lower temperature and lower superheat mattes and blister Cu or even 200–300 L of tap-hole clay per closure when trying to
operations, e.g. ‘Polish plug’, comprising ceramic stabilize a ‘mushroom’ (Eden et al., 2001; Irons, 2001).
surrounding a cone-shaped tap-hole clay ‘plug’ (George- In an ironmaking BF, where tap-hole clay ‘mushroom’
Kennedy et al., 2005). Over 25% of PGM and local Ni matte operation is feasible, several operators report stable
operations still practise manual plugging of tap-holes (Nolet, (consistently deep) tap-hole length and reduced tap-hole clay
2014; Coetzee, 2006). consumption, i.e. ‘not excessive addition’ (Nightingale et al.,
However, by far the majority of ferroalloy furnaces, 70% 2001; Tanzil et al., 2001; Cassini, 2001), and reduction by as
of PGM and local Ni matte operations (Nolet, 2014), and all much as 50% to 100–120 L on a 3 m tap-hole length
ironmaking BFs have increasingly adopted sophisticated and (Nightingale et al., 2001), which led to generally improved
powerful mudguns to effect tap-hole closure. Again, the overall practice (Smith et al., 2005; Jameson et al., 1999;
importance of considering mudgun, tap-hole clay, and tap- Black and Bobek, 2001; Tanzil et al., 2001; Estrabillo, 2001;
hole operating practice holistically as a fully integrated Nightingale and Rooney, 2001; Bell et al., 2004; Cassini,
system cannot be understated – coupling a hard new- 2001). This is particularly the case when the tap-hole-clay
generation tap-hole clay with an old weak mudgun incapable injection rates – rapid to assist with clean plugging of tap-
of properly delivering the clay into the tap-hole is bound to hole clay down the tapping channel, yet with sufficient time
fail. Smith, Franklin, and Fonseca (2005) describe this well: for densification and crack sealing of the protective annular
the ‘design of tap-hole clay is usually a compromise between tap-hole-clay tapping-channel core (Andou et al., 1989;
“equipment capability” and “process” requirements.’ Smith et al., 2005) – and quantities added are controlled
predictively, based on prior tapping and drilling metrics.
&') $!)&%'(&'"!) (%$"& Again, this can involve SPC to control tap-hole length
Manual plugging may at first glance seem extremely (e.g. to 3.1 m; Jameson et al., 1999) by varying the tap-hole
simplistic, requiring a direct interface of the operator with a clay volume (around a 100 L setpoint; Jameson et al., 1999);
hot tapping stream. However, if the operation is not correctly or by advising the operator of the recommended tap-hole clay
controlled, excessive tap-hole clay addition – which is volume after 1.5 hours of tapping, basing the advice on
possible with the use of automated mudguns – can potentially automatically measured tap-hole lengths and tap-hole
have a destructive, but often hidden, action on a tap-hole and diameter (the latter automatically inferred from measured
lining environs. It was not that long ago that one of the blast pressure, liquid level, and mass tapping rates
authors witnessed a large furnace, about 30 m in length, (Nightingale et al., 2001; Tanzil et al., 2001). Continuous
‘disappear from view’ due to excessive gas release and a weighing using load cells and microwave radar level detection
concentrate blowback when a tap-hole was closed with a full are used to determine hot metal torpedo and/or slag-ladle
25 L mudgun load of wet clay recently ‘dug from the veld’. filling rates, and thus related mass tapping rates (Tanzil et
Other observations include both metal and matte ‘boils’ at the al., 2001; Cassini, 2001; Shao, 2013). Operation usually
back of tap-holes, tap-hole ‘blows’, and even gas eruption involves increased tap-hole clay injection when the tap-hole
from tar binder (Mitsui et al., 1988) caused by mudgun length decreases, and decreased clay injection when the
closure involving use of excessive tap-hole clay with high length increases. In consecutive individual tapping practice in
loss-on-ignition content. Water flashes with a 1500-times particular, a common additional practice advocated on the
volume increase at bath temperatures, and hydroxides, other resting tap-holes is for occasional tap-hole clay
carbonates, and hydrocarbons can react almost injection to maintain the ‘mushroom’ condition on those tap-
instantaneously and decompose, devolatilize, and crack holes, which otherwise are subject to progressive dissolution
L

480 VOLUME 116


 
      
The tap-hole — key to furnace performance
(if metal is marginally carbon-unsaturated) and wear in al., 1999). Slew pressure is usually set slightly higher than
contact with hearth liquid (Jameson et al., 1999; Nightingale the mudgun barrel pressure (200–315 bar tap-hole clay
and Rooney, 2001). pressure, which results in a pushing force of > 60 t on the
Ironmaking BF experience suggests that less than one- tap-hole face/faceplate, particularly to push higher-strength
third of tap-hole clay purchased is pushed through the gun. tap-hole clays [Van Ikelen et al., 2000; Smith et al., 2005;
This wastage is ascribed to combinations of (1) incorrect Black and Bobek, 2001; Atland and Grabietz, 2001; Cassini,
storage under uncontrolled conditions of temperature; (2) the 2001]) – a measure that tends to limit the potential for
tap-hole clay getting wet; or (3) situations where the tap-hole bypass of clay between the nozzle and tap-hole face/faceplate
clay is allowed to go beyond its useful shelf life. Of the (Eden et al., 2001; Cámpora et al., 1998; Jameson et al.,
remaining tap-hole clay, only 24% is estimated to be 1999; Entwistle, 2001). Automatic control of the mudgun
delivered into the tapping channel (Smith, Franklin, and contact force is also preferred in order to limit the risk of
Fonseca, 2005) (Figure 16). Nozzle cleaning, push-out waste undue mechanical damage to the tap-hole refractory, a
(used to ensure that tap-hole clay is compressed in the control that one site achieved by a variable-machine,
mudgun barrel), clay leakage between the nozzle and tap- minimum-pressure setpoint of 150 bar plus a variable
hole face (Figure 17 and Figure 18), mudgun clean-out, and proportion of 0.3 times the plugging pressure (Ballewski et
20% for ‘mushroom’ replacement constitute the remaining al., 2001). In the absence of rigid faceplates, tap-hole face
portion of tap-hole clay usage. wear can be estimated from a relationship to cylinder stroke
Sacrificial wooden or ceramic nozzle covers – known measured by LVDT (Black and Bobek, 2001; Entwistle,
locally as ‘dinner plates’ (Ndlovu et al., 2005; Figure 19) – 2001).
are commonly used to limit tap-hole clay losses associated In the extreme practice of combination drill and ‘plug’
with mudgun push-out waste (full nozzle cover) and nozzle- oxygen lance, which aims to avoid excessive tap-hole clay
face/faceplate leakage (full or annular nozzle cover [Ndlovu delivery beyond the tapping channel hot-face (for fear
et al., 2005; Eden et al. 2001; Jameson et al., 1999; De
Pagter and Molenaar, 2001; Brunnbauer et al., 2001;
Estrabillo, 2001; Bell et al., 2004]). A 25% reduction in
mudgun-nozzle tap-hole clay leakage events, from a
somewhat poor norm of 50%, has been reported for this
practice (Estrabillo, 2001).
Well-designed faceplates normally further improve
mating with a flat nozzle face – common on Co and Ni
ferroalloy and matte-smelting operations. However, where
faceplates are absent, some ironmaking BF operations have
adopted tapered nozzle tips, for which better sealing against
the tap-hole socket is claimed (Steigauf and Storm, 2001).
Upgrading to high-nitride mudgun barrels is also cited as a /B98@FG+->F<<B(FGED=6:C?FG>?D1G41=D<<G) ;?C(8Get al.0G.22$*
factor preventing wear (Petruccelli et al., 2003; Bell et al.,
2004).
On modern mudguns, rapid and automated pressure-
regulated mudgun slew is applied to minimize damage to the
mudgun nozzle, and to lower the risk of heavy impact on the
tapping channel face and/or channel, a risk that might
otherwise crack or even dislodge tap-hole refractory and the
ironmaking BF ‘mushroom’ (Smith et al., 2005; Jameson et

/B98@FG+-CAE@C??F;G;C<D9FGC7GED=6:C?FG>?D1G) ;?C(8Get al.0G.22$*

/B98@FG+-<EB5DEBCAGC7GED=6:C?FG>?D1G>CA<85=EBCAGDA;G3D<EFG)D7EF@
/B98@FG+#-"D>@B7B>BD?G58;98AGAC!!?FG%;BAAF@G=?DEF&G) ;?C(8Get al.0G.22$*
"5BE:G)%'(#0G.22$*
L

 
       VOLUME 116  481
The tap-hole — key to furnace performance
otherwise of the tap-hole clay boiling and ensuing damage to ® After curing, it should attain the required strength
the tap-hole hot-face), precise control of tap-hole clay input is (often described as ’sinterability’ [Abramowitz et al.,
imperative. This often involves measurement and automated 1983]) without shrinkage to ensure a tight seal within
control of the injected tap-hole clay volume. Indeed, in the tap-hole (and not prematurely in the mudgun), do
several instances when tap-hole clay addition has been so in the required mudgun dwell time, and plug the
excessive (Hundermark et al., 2014; Ndlovu et al., 2005) it hole until the next tapping time
has been demonstrated that controlled reduction of tap-hole ® It should effect safe tap-hole closure (i.e. without
clay additions (closer to the volume predicted theoretically for subsequent re-opening) without damage to the tap-
‘normal’ tap-hole dimensions) has even resulted in increased hole and furnace lining (e.g. through limited gas
drilling depths, further enhanced by improved furnace evolution and associated turbulence), yet with the
operating control of allowable upper matte temperature ‘mushroom’ remaining stable where required, e.g. in an
(Figure 20). ironmaking BF. This requires consideration of both
On ironmaking BF operations (Smith et al, 2005; effective tap-hole clay displacement in the injection
Ballewski et al., 2001; Tanzil et al., 2001; Bell et al., 2004), direction (Uenaka et al., 1989; Nakamura et al., 2007;
staggered, multi-stage mudgun injection at different speeds Kitamura, 2014) and a ‘good spreading ability in the
can be practised to achieve optimal tap-hole conditions. This direction perpendicular to the injection direction’ to
may involve (Bell et al., 2004) (1) a first fast push of 45 kg maintain a stable ‘sedimentary deposit that is gradually
tap-hole clay to displace any other material from the tapping and stably grown’ (Kitamura, 2014) and exhibiting
channel, followed by a slower push of another 45 kg clay to good high-temperature adhesion to the constituents
build the ‘mushroom’, and a final very slow push of variable already present in the tapping channel (Niiya et al.,
clay mass to build the ‘mushroom’ still further and compact 2012)
the tap-hole clay in the tap-hole, and (2) a second very slow ® It should be soft enough to be readily drilled straight
push 5 minutes after the first push, with < 5 kg tap-hole clay down the middle of the tapping channel without
added to compact the tap-hole clay still further and close deviation and in an acceptable time (this is especially
voids. To diminish the risk of tap-hole breakout, the mudgun important where productivity constraints exist, as in an
then remains in position for 5 minutes to allow adequate tap- ironmaking BF)
hole clay curing before the mudgun is removed from the tap- ® It should allow a stable, controlled tapping stream flow
hole face. On another operation, with a constant ram without surging or splash (often associated in
hydraulic pressure of 275 bar, a rate of tap-hole clay injection ironmaking with blast gas tracking [He et al., 2001;
of 14 kg/s was sought (Black and Bobek, 2001). Pan and Shao, 2009] and gas entrained with ‘viscous
fingering’ to above the critical value that induces a
(!"#)'#( deleterious splashing casting stream [He et al., 2002;
 2012], even to the extent of slug flows [Shao, 2013;
Stevenson and He, 2005; Shao and Saxen 2011,
Typical requirements cited for tap-hole clay include the
2013b])
following (Abramowitz et al., 1983; Andou et al., 1989;
® It should be ideally ‘hard’ and durable (Abramowitz et
Uenaka et al., 1989; Hubert et al., 1995; Ballewski et al.,
al., 1983) enough to withstand penetration, corrosion,
2001; Cassini, 2001; Wells, 2002; Smith et al., 2005; Horita
and erosion by the tapped metal/matte and/or slag and
and Hara, 2005; Kageyama et al., 2005, 2007; Nightingale et
so preserve a protective annulus between the tap
al., 2006; Nakamura et al., 2007; Pan and Shao, 2009; Niiya
stream and tap-block refractory (without additional
et al., 2012; Kitamura, 2014):
corrosive reaction to the tap-hole refractory), thereby
® It should be soft and plastic enough to inject when extending the useful life of the tapping channel with an
pushed by the mudgun, but ‘hard’ enough to displace acceptable, controlled diameter, shape (i.e. minimal
tapping liquid effectively and deliver a ‘plug’ of tap- long-term fluting), and length.
hole clay only to the required depth in the tapping
To ensure optimal tap-hole clay quality, additional
channel
measures for tap-hole clay preparation are recommended
(Black and Bobek, 2001; Delabre et al., 1991; Hubert et al.,
1991). These measures may include a stand-alone tap-hole
clay storage building, maintaining a 10-day supply of tap-
hole clay, and controlling the temperature in the building and
the in-process temperature for storing tap-hole clay at the tap
floor to 25–30°C (Abramowitz et al., 1983). Maturation of the
clay as a function of the binder quantity and type for two
months is reported (Delabre et al., 1991) to permit one clay
to attain desired plasticity properties, these properties also
being dependent on storage conditions. On the other hand,
prolonged storage of resin-bonded tap-hole clay, especially at
temperatures exceeding 40°C, is reported specifically as being
/B98@FG.2-F;8>F;GED=6:C?FG>?D1GBAF>EBCAG)41G2*G?FD;BA9GEC detrimental to its performance (Wells, 2002). However,
BA>@FD<F;G;@B??BA9G;F=E:0GF(FAG4F7C@FGEB9:EF@G5DEEF6EF5=F@DE8@FG>CAE@C? especially for tar-bonded tap-hole clays, a minimum of 15
) ;?C(8 )%'(#0G.22$* days’ ageing is reported as essential to secure adequate tap-
L

482 VOLUME 116


 
      
The tap-hole — key to furnace performance
hole clay loss in plasticity and increased hardness (Hubert et et al., 2004) had reverted to tar-bonded tap-hole clay, while
al., 1995). A tap-hole clay producer even reports forced in Europe tar, resin, and resin-tar binder combinations all
cooling of tap-hole clay to avoid any risk of continued undue continued to find favour (Irons, 2001). By 2005 one supplier
temperature rise before final packaging of product of tap-hole clay reported that only two ironmaking plants in
(Nakamura et al., 2007). Japan were using resin-bonded tap-hole clay (Horita and
Hara, 2005).



  Tar-bonded tap-hole clays are generally thermoplastic,
Most technical developments of tap-hole clay originate from hard (often requiring pre-heating of the tap-hole clay in the
the ironmaking BF industry, where the high productivity (10 mudgun barrel by gas heaters, hot water, or steam to become
000 t/d HM), combined metal and slag duty, high pressure pliable [Ballewski et al., 2001], especially for operation in
(approx. 10 bar at tap-hole - Van Laar, 2014), and long tap- colder climates), and slower curing (a cast time of 2 hours is
hole length (2.5–4 m – Table I) make high demands on tap- deemed insufficient for full curing and sintering [Black and
hole clay quality. Mitsui and co-workers (1998) use lowering Bobek, 2001], although only 20–30 minutes is frequently
of specific tap-hole clay consumption (kg/t HM) to outline encountered as being available for curing in practice [Uenaka
early developments from the 1970s to 1988. These include a et al., 1989; Shao and Saxen, 2011, 2013a; Shao, 2013]).
progression from coke, to alumina, to silica, and back to Slow curing necessitates the mudgun remaining in position
pitch-impregnated alumina (Niiya et al., 2012) and high- for an extended time after plugging to avoid the tap-hole re-
alumina clays comprising a fine matrix (< 45 m and > 50% opening unintentionally. Unlike resin binders, tar-bonded
by mass [Kageyama et al., 2007; Horita and Hara, 2005]) tap-hole clay is reported to have the advantage of forming a
and/or coarser aggregates (1–3 mm and approx. 20% by transition-free union with a carbon-based refractory, which
mass [Horita and Hara, 2005]). These clays include variously results in a monolithic tap-hole lining (Ballewski et al., 2001)
additions of zirconia, kyanite (Andou et al., 1989), SiC, and and improved adhesiveness at high temperatures (Niiya et
metals or nitrides of silicon, aluminium, and ferrosilicon. al., 2012). Radiant heating from the tapping launder may
These elements and compounds are added as fine powders to necessitate protection of the barrel by metal or ceramic
the matrix to lower porosity, reduce shrinkage, decrease insulating shields (Bell et al., 2004), or even water-cooling.
volatiles, increase antioxidant action, lower wettability by Mudguns with a partial or full circumferential water jacket
slag, and improve extrudability, sintering, and resistance to and dual heating/cooling systems are quite common in
corrosion and erosion (Abramowitz et al., 1983; Andou et ironmaking BFs and some Cr and Mn ferroalloy operations.
al., 1989; Uenaka et al., 1989; Black and Bobek, 2001; These systems are often automated to operate at a fixed
Wells, 2002; Smith et al., 2005; Kageyama et al., 2007; temperature setpoint, e.g., a constant 50–65°C (De Pagter and
Nakamura et al., 2007; Pan and Shao, 2009; Niiya et al., Molenaar, 2001; Ballewski et al., 2001; Bell et al., 2004;
2012; Kitamura, 2014). Mention is also made of a trend to Atland and Grabietz, 2001); or for maximum flexibility an
smaller particle size for improved compaction (Black and adjustable, controlled temperature range, e.g., 25–90°C
Bobek, 2001) and better sealing of the tapping channel (Black and Bobek, 2001) is provided for and tailored
against gas egress (Pan and Shao, 2009). Some sources even specifically to a given tap-hole clay type in use (Cámpora et
claim that ultrafines (< 10 m [Kageyama et al., 2007]) al., 1998). A reduction in tap-hole clay consumption by
improve strength, resistance to corrosion and abrasion, and wastage of between 10 and 30% is reported in uses of water-
an ‘ability to go straight during gun-up instead of extending cooled mudguns (Ballewski et al., 2001; Atland and Grabietz,
transversely inside the furnace’ (Nakamura et al., 2007). 2001).
Improved corrosion resistance and higher positive residual Resin-bonded tap-hole clays are faster curing (Uenaka et
expansion coefficients of pure silica and pure alumina al., 1989; Wells, 2002; Kageyama et al., 2005), a property
sources compared with aluminosilicates are also reported promoting shorter mudgun dwell time and quicker tap-hole
(Mitsui et al., 1998). Such a ‘swelling’ characteristic (Mitsui turnaround (of importance in a high-productivity operation
et al., 1998; Cassini, 2001; Nightingale et al., 2006) is such as an ironmaking BF). Occasionally, though, the tap-
important for helping to seal a tap-hole subject to hole clays can cure too quickly, which leads, in hotter tap-
temperature fluctuation from the extreme of superheated holes, to the clay curing before injection is complete (Jameson
tapping temperatures to cold closure conditions in water- et al., 1999; Nakamura et al., 2007); or, in the extreme, to its
cooled tap-blocks. Additives are also beneficial in instances blocking prematurely in an excessively hot mudgun barrel
where the clay has not fully baked before the next tap, and and possibly delaying an effective tap-hole closure. Resin-
provide strength at lower clay temperatures (Delabre et al., bonded tap-hole clay can be prone to greater volatility upon
1991). A somewhat more empirical approach has similarly heating (Kageyama et al., 2005), to more undesirable gas
led to convergence on the use of tap-hole clays of high evolution (observed in local industry), and is generally softer
alumina content for high-intensity operations in the local (to the extent of being found incapable of effecting tap-hole
pyrometallurgical industry. closure on some high-temperature and superheated Cr metal-
only and PGM matte-only tap-holes). Some resin-bonded tap-
  hole clays have also been reported to cure too hard for
Traditionally, coal tar pitch was used as a binder (approx. acceptable drill times (< 15 minutes), a development
20% by mass [Kageyama et al., 2005]) in tap-hole clay. This requiring binder reformulation (Nakamura et al., 2007). In
was followed by a period in the 1990s where phenolic resin high-intensity PGM matte operations, the risk of failing to
found favour. By 2001 it was reported that 90% of Japanese close timeously a ‘vicious’ superheated tap is considered so
ironmaking BFs (Irons, 2001) and a Canadian producer (Bell extreme that procedures further dictate that no matte tap-hole
L

 
       VOLUME 116  483
The tap-hole — key to furnace performance
be opened without the availability of two fully prepared phase binder A plus binder B (made of several mixtures)
mudguns loaded with tar-bonded tap-hole clay to close the manifested comparable plasticity, high-temperature
matte tap-hole. adhesivity, high thermal expansion, and low erosion
Ballewski and co-workers (2001) observe that generally properties to existing tap-hole clays.
‘the lower the temperature, the more difficult the correct Industry’s adoption of non-polluting tap-hole clays has
choice of a binder system for mud becomes . . . otherwise the not been universal. This is possibly owing to concerns
front tap-hole area would extend negatively on the cold side’. regarding some perceived deficiencies in their performance in
Ostensibly for other reasons, just such a tap-hole extension tap-hole duty for certain commodities compared to more
outside the furnace (colloquially described as a bullnose, traditional tar-based clay products.
Figure 8) is precisely what has tended to happen with
intensely cooled copper tap-blocks. 
  
Abramowitz and co-workers (1983) reported that ‘small Tap-hole opening, the act of tapping metal/matte/slag, and
changes (< 5%)’ in either light oil loss (260°C for 6 hours) or tap-hole closure all lead to increased environmental
loss on ignition (defined by them at a temperature of 1204°C, emissions around the furnace. That is, emissions associated
rather than a more common 900°C, or 1000°C [Hubert et al., with drilling uncured tap-hole clay, or fumes released in
1995]) can ‘change many dimensional and strength oxygen lancing; release of process gases such as CO or H2
properties (as high as 119%)’ of tap-hole clays. This under pressure, especially in ironmaking BFs, or SO2,
emphasizes the need for close control of conditions in the possibly even H2S, by release or reaction, especially in matte
manufacture of all tap-hole clays, if one is to yield a product smelting, but also other trace gases, e.g., Cl and F, or
of consistent quality. Using cold crushing strength (> 7.6 contained volatile heavy metal impurities, e.g. Pb, As, Cd,
MPa) and workability (18–28%) as quality criteria in the and Zn, depending on specific composition; and volatile
early 1980s, tap-hole clay manufacturers found that tap-hole emissions from injecting tap-hole clay. Extraction systems on
clay rejection rates of up to 40%, sometimes more, were not tap-hole, launder, and ladle hoods (Figure 1), and even on
uncommon. Rejection rates below 15% were suggested as entire tapping aisles, are increasingly required to achieve the
acceptable. necessary and acceptable workplace hygiene and
In the local industry, variable supply and quality of coal environmental abatement.
tar pitch has at times led to suboptimal ‘cutting’ additions of
oils to overly viscous pitch, with additions of resin to try to 'D=6:C?FG5DBAEFADA>FGDA;G?B7F
restore curing times. With binder additions of typically 20%
by mass (Kageyama et al., 2007), this has often led to tap-  ))&%(%$)'($&%)&(&)
hole clays being prone to excessive gas evolution and having Ironmaking BFs incorporate robust designs that usually last
suboptimal handling and plugging characteristics, properties for more than 10 years with little maintenance reported
that make the clay possibly suitable for less onerous slag tap- (Steigauf and Storm, 2001) of the castable at the front cold-
hole closure, but unsuitable for high-duty, superheated Cr face (and without ‘mickey’ bricks). An original four-tap-hole
metal-only and PGM matte-only applications. construction that lasted 12 years is also reported. Other BF
sites report a 28-day cycle of casting tap-hole pairs (Steigauf
   and Storm, 2001), or recasting of tap-hole faces in planned
While imposing a minimum 45-minute tap-hole clay curing maintenance scheduled every 18 weeks (Tanzil et al., 2001).
time before re-drilling and tapping is reported to result in less Tap-block graphite block inspection every 4 years is also
emission of fumes (Estrabillo, 2001), tar binders pose health reported (Black and Bobek, 2001).
risks through the release of polycyclic aromatic compounds Longer time-frames and operation to tap-hole breakout
such as benzopyrene, which are carcinogenic (Perez et al., (usually within 3–4 years) are also practised on many local
2001; Hershey et al., 2013; Irons, 2001). The release of ferroalloy furnaces, but usually with the consequence of far
similarly undesirable formaldehydes and phenols is more severe furnace lining damage and shortened cycle times
associated with resin binders. Molenaar (Irons, 2001) argues to the next breakout. A notable exception is a campaign life
that benzopyrene particles in the air condense on dust, and of 9–12 years before a first small tap-hole repair, reported on
hence some protection is afforded by wearing a mask, which combined metal-slag tap-holes of freeze lining design on a Cr
is ineffectual for protection against formaldehyde and phenol ferroalloy furnace (Duncanson and Sylven, 2011). On Cr
gases. ferroalloy furnaces with water-cooled copper tap-block
Non-polluting tap-hole clay is therefore desirable, elements, other periodic planned maintenance may present
provided that it can adequately meet the arduous duty and ideal opportunities to effect annual slag tap-block and/or
requirements of tap-hole clay without introducing further risk biannual metal tap-block repairs.
(e.g. tap-hole liquid breakout). Tarless tap-hole clays have Most typically in Mn ferroalloys, total furnace (and by
been available since the 1970s (Hubert et al., 1991), as well inference tap-hole) life is reported as being only 6–10 years
as tap-hole clay utilizing commercial tar binder of ‘one (De Kievit et al., 2004; Van der Walt, 1986; Hearn et al.,
thirtieth of the benzopyrene’ content of ordinary coal tar for 1998), with some early freeze lining furnace designs giving
binders (Kitamura, 2014). More recently clays which do not over 20 years of life being the exception (Van der Walt, 1986;
contain any polycyclic aromatic hydrocarbons have also Hearn et al., 1998). On many Mn ferroalloy furnaces,
become available (Lungmuß Feuerfest, 2014). However, it periodic ‘mickey’ block replacement may be planned and
was reported (Perez et al., 2001) that non-polluting single- performed as often as every 6 months, with a tap-block
phase binders have proved unsuccessful, but that a single- campaign life of three years being typical. With the freeze
L

484 VOLUME 116


 
      
The tap-hole — key to furnace performance
lining tap-hole design, life in excess of 18 months for the thermal contact and reduce gas tracking (Edwards, and
annular replaceable carbon block and graphite sleeve design Hutchinson, 2001). Details of several grouting and zoned-
has been reported in SiMn production (Hearn et al., 1998), plug (blind) repairs and basic procedures in ironmaking BFs
and 400 mm (out of 870 mm) wear of tap-block hot-face in are described by Cámpora and co-workers (1998), Yamashita
just over three years of service is reported in HC FeMn and co-workers (1995), and Ballewski and co-workers
production (O’Shaughnessy et al., 2013). Apparently this (2001). A caution is sounded: great care should be exercised
wear is not attributed to erosion by tapping practices alone in grouting, using a sufficient number of open grouting
(which involve drilling and minimal use of oxygen lancing); points in the repair vicinity, to avoid the risk of grouting
rather, it is suspected that standard furnace thermal leading to excessive build-up of pressure and so leading
equilibrium conditions do not always permit the tap-holes to inadvertently to refractory movement and even to lining
remain at their design length (O’Shaughnessy et al., 2013). failure.
On Mn ferroalloy furnaces with water-cooled copper tap-block A comprehensive mudgun and drill inspection
elements, planned maintenance activities are understandably programme, with weekly, monthly, quarterly, and annual
more aggressive, with ‘mickey’ repairs being carried out as activities to ensure equipment reliability, and early detection
frequently as every 4 months. and prevention of possible failures, is described by Petruccelli
Blister tap-holes were reported to operate for 8 000 t (2003). Reliability of air supply on pneumatic drills is quoted
between inner change, while flash converting furnaces were fairly frequently as a cause of poor drilling, with air
projected to deliver more than 4 years of life (George, 2002). accumulators and new compressors being installed to address
The latest furnace life estimate is now in excess of 5 years the problem (Petruccelli, 2003).
(George-Kennedy et al., 2005).
On Ni ferroalloy and Ni and PGM matte furnaces, 'D=6:C?FG5CABEC@BA9
preventative maintenance may be time-based on lower
intensity furnaces, but is more usually based on number of %(&( '%(!"#)'"&$%" $&
taps (Nolet, 2014; Jastrzebski et al., 2012) rather than on Three general levels of tap-hole monitoring are identified:
mass tapped, with the assumption that a ‘tapping event’
® Limited use of single thermocouples inserted into the
(comprising tap-hole opening, tapping, and tap-hole closing)
lining, some around the tap-hole, often associated with
is a more significant determinant of tap-hole wear than the
a furnace campaign (let alone tap-hole) life of under 6
mere act of tapping. As previously intimated, especially on
years on both (historically) ironmaking BFs (Eden et
higher-intensity superheated PGM matte operations,
al., 2001; Jameson et al., 1999) and ferroalloy furnaces
excessive tapping rates can also be used to trigger tap-hole
(Van der Walt, 1986; Hearn et al., 1998; De Kievit et
maintenance.
al., 2004; Coetzee and Sylven, 2010; Coetzee et al.,
Typical tap-hole maintenance cycles that result are 1–4
2010; Duncanson, and Sylven, 2011)
weeks between faceplate refractory insert and shallow
® Progression to more thermocouples (15–50),
tapping module brick replacement; quarterly for deep tapping
predominantly duplex in configuration, to permit heat
module and/or surround brick replacement; 1–2 years for full
flux calculations and monitoring on ironmaking BFs
tapping channel repair and potentially water-cooled copper
(Stokman et al., 2004; Jameson et al., 1999; Irons,
tap-block replacement (Nolet, 2014).
2001; Entwistle, 2001) and ferroalloy furnaces (De
In addition, condition-based maintenance can be
triggered immediately by any of the following: suspicion of Kievit et al., 2004). Furnace campaign life now ranges
any water leak; overly skew drilling; overly skew lancing more typically between 10 and 20 years on both
(less easy to diagnose); excessive oxygen lance consumption; ironmaking BFs (Van Laar et al., 2003; Eden et al.,
undue difficulty in tap-hole closure by tap-hole clay; 2001; Jameson et al., 1999) and ferroalloy furnaces
damaged faceplate refractory insert; damaged faceplate (the (Van der Walt, 1986; Hearn et al., 1998; De Kievit et
flat, vertical, mating surface presented to the mudgun nozzle al., 2004; Duncanson and Sylven, 2011)
is compromised); the insert tapping channel diameter is ® Dedicated multiple thermocouples for in-tap-hole
greater than a prescribed limit (practice requires the oversize temperature measurement (Estrabillo, 2001), heat flux
diameter to be followed down the tapping channel, replacing probes (equipped with thermocouples) (Atland and
adjacent tapping module bricks until the diameter is deemed Grabietz, 2001), and, in the extreme, up to 30 copper
within a prescribed limit); and tap-hole temperature spikes thermocouples and water circuit RTDs (to determine
reaching above alarm limits. temperatures, and water temperature rises and
associated local heat fluxes) to monitor water-cooled
!)$(#'($&%)&(&) copper tap-blocks (let alone the adjacent furnace
Online repair techniques to improve the tap-hole condition on lining), practices that are adopted on some high-
ironmaking BFs include (Yamashita et al., 1995; Jameson et intensity non-ferrous operations.
al., 1999; Ballewski et al., 2001) the following: use of higher
(&)'%(!"#)'"&$%" $&
plasticity tap-hole clays to help seal gaps and reconstruct
‘mushrooms’; use of an emergency ‘nozzle can’ (Estrabillo, Conditions inside the tap-hole during tap-hole clay curing can
2001); injection of resin down a partially drilled (blind) tap- be determined by drilling a pilot hole and
hole to seal cracks and reduce gas tracking; and grouting ® Inserting temporary thermocouples down the tapping
through injection under pressure of tar-bonded carbon channel to determine tap-hole temperature profiles with
mortars to fill voids more generally and so re-establish depth, e.g. on ironmaking BFs temperatures rise on
L

 
       VOLUME 116  485
The tap-hole — key to furnace performance
average from 200 to 800°C (maximum 550 to 1200°C) Although conventional duplex thermocouples are capable
from 0.25 to 1.75 m down the tap-hole (Abramowitz et of detecting the accumulation of thermal energy in tap-hole
al., 1983; Delabre et al., 1991; Cámpora et al., 1998; refractories when tapping in close succession on one matte
Ballewski et al. 2001; Nightingale et al., 2006; Niiya et tap-hole (Figure 10 and Figure 21), fibre optics provide more
al., 2012), thereafter reaching a plateau down the 3 m detailed local mapping of the distributions and rises in
long tap-hole (Ballewski et al. 2001) and with time temperature associated with consecutive tapping.
(700– 900°C within 30 minutes of injection (Entwistle, Alternatively, the beneficial effects of resting a tap-hole
2001), and applicable to establishing tap-hole-clay set- to lower temperatures, as practised in alternating tapping
up times (Abramowitz et al., 1983; Mitsui et al., 1988; procedures, may require less exhaustive monitoring
Delabre et al., 1991; Cámpora et al., 1998; Ballewski et (Figure 11).
al., 2001; Bell et al., 2004; Niiya et al., 2012) Preliminary results seem to confirm a distinct temperature
® Measuring with a contact thermocouple at 20 cm rise following tap-hole closure, consistent with the previously
intervals, even reported apparently in normal opening observed heat load rise following tap-hole closure (Cameron
of the tap-hole (Ballewski et al. 2001) et al., 1995). Cooling during tapping was also reported
® Inserting nitrogen-cooled (Entwistle, 2001) or water- (Figure 22). The former was plausibly ascribed to the
cooled (Ballewski et al., 2001) fibre-optic cameras to significantly increased heat flux associated with tap-hole clay
view internal tapping channel conditions coming into contact with superheated matte and the
® Core drilling the curing/cured tap-hole clay samples for associated turbulence of gas bubbles and concomitant
chemical and mineralogical/petrographic analysis and enhanced heat transfer. The cause of the latter is unknown.
physical testing (Andou et al., 1989; Van Laar et al., Depending on whether temperatures are measured at the
2003; Shao, 2013), sometimes on a two-year planned chamfer or in the tapping channel during tapping on another
maintenance cycle (Bell et al., 2004). matte tap-hole (Figure 21), recent fibre-optic measurements
seem to show some trends of rising temperatures (and only
In some BF stacks ceramic rods are integrated into the
sometimes falling) already during tapping. Temperatures
lining to permit wear to be determined by the ultrasonic
continue at best on a similar trajectory, rather than with any
measurement of rod length (here one assumes ceramic wear
distinct rise as may be expected for tap-hole clay closure
is coincident with lining wear [Stokman et al., 2004]). What
(tapping events being determined from pyrometer
is uncertain is the status of the application and the efficacy of
temperature data). Moreover, while it may be tempting to
this or any alternative, external, nondestructive testing ascribe apparent minor temperature dips around some tap-
(NDT), acoustic emission (AE) technique (Sadri et al., 2008; hole closure events to a theoretically plausible effect of tap-
pers. comm. 2010, 2011) in more intensely cooled tap-hole
regions, frequently with the presence of composite refractory
materials and/or water-cooled copper coolers.
A more recent development has been the use of electrical
resistance-based sensors (continuous along the length of the
sensor, but for peak temperature only [Hopf and Rossouw,
2006]) and fibre-optic temperature sensors (either
continuous sensor length maximum temperature, or discrete
[about 25 per sensor] temperature measurements [Gerritsen
et al., 2009; Hopf, 2014; pers. comm. 2010]) to record more
accurately and better map matte (and slag) tap-block
temperatures in Ni and PGM matte smelting.
This development is an attempt to avoid a ‘porcupine’
copper tap-block containing more conventional copper
thermocouples and water RTDs than available for labelling by
the alphabet! This recognizes the limited range (akin to /B98@FG.+-F>FAEGED==BA9G=F@BC;G>?C<F68=GC7G7B4@F6C=EB>GEF5=F@DE8@F<
‘fishing with a rod’ [Wasmund, 2003], so raising the chance CAGDACE:F@G5DEEFGED=64?C>,G)C=FA6>?C<FG<:C3AGD<G?BAF<*
of missing vital information) of only local temperature
detection by thermocouples in an intensely water-cooled
copper tap-block environment. Tap-blocks are often
equipped with yet a third redundant cooling circuit recessed
towards the cold-face. This circuit has been successfully used
on at least one occasion as a backup water circuit that
permitted the safe shutdown of the furnace under controlled
conditions following a cooler ‘hit’ and the loss of the primary
hot-face cooling water circuit and/or furnace refractory
breakout. This further recognizes the more global monitoring
capability, but poorer temperature resolution, of a rise in
cooling water temperature on intensely cooled copper tap-
blocks (akin to ‘fishing with a net’ [Wasmund, 2003], so /B98@FG..-?C<F68=GC7GED==BA9G=F@BC;G<:C3BA9G7B4@F6C=EB>GEF5=F@DE8@F
being better at capturing key thermal events). ;@C=G;8@BA9GED=G)D7EF@GF@@BE<FAGet al.0G.22#*
L

486 VOLUME 116


 
      
The tap-hole — key to furnace performance
hole cooling by the injection of tap-hole clay (modelled in Systems that require closed-circuit water cooling need to
SiMn ferroalloy tapping [Muller and Steenkamp, 2013]), take the following into account: monitoring frequency, the
closer analysis suggests that the apparent dips are more rate of change of make-up water, and the standpipe level to
likely to be an artefact of cooling induced periodically by the detect leaks (Jameson et al., 1999; pers. comm. 2010);
coarse temperature control of the copper tap-block water- differential flow (MacRosty et al., 2007; pers. comm. 2010)
cooling heat-exchanger circuit. and multi-tier sensors that involve the monitoring of copper
Clearly, still more work is required to understand and and water temperatures, water flow, and (periodically) water
explain fully the tap-block fibre-optic temperature trends. pressure (Valentas and Thierney, 2010; Shaw et al., 2013;
This is a fundamental requirement prior to attempting more Bussell et al., 2013). Such automated pressure testing of
complex projections of possible tap-hole brick wear trends by individual cooler-water circuits (at operator-selected scan
thermal modelling – projections in support of advanced rates) has proved capable of detecting even the smallest of
condition monitoring. ‘drip’ leaks on commercial furnaces (Shaw et al., 2013).
In response to difficulties and risks in the timeous These systems are most direct and effective, but more
detection of a significant tap-block temperature rise (see expensive and, for safety, dependent on coolers equipped
Dedicated metal/matte tapping section), other more advanced with redundant water-cooling circuits and/or process and
monitoring and diagnostic systems have been pursued, cooler design conditions in which the termination of a cooler
including principal component analysis (PCA), to try to water circuit supply for a brief period of pressure testing
provide some advanced view of the development of some carries no risk of converting water so entrapped to steam.
abnormal tap-hole conditions (Gerritsen et al., 2009; Plikas
et al., 2005; pers. comm., 2010), and tap-hole acoustic CA>?8<BCA<
monitoring (TAM), which has the potential to identify the
The critical importance of tap-hole design and management
development of off-centre lancing (Sadri et al., 2008;
to furnace performance and longevity on a variety of ferrous
Wasmund, 2003; pers. comm., 2010, 2011).
and non-ferrous smelting processes has been demonstrated.
Process conditions and productivity requirements dictate
'D==BA9G<1<EF5G3DEF@G:D!D@;< specific differences and similarities in tapping equipment and
Given the sheer rapidity, often with very limited warning, and practices, and in managing tap-hole operations and
consequences of a matte/blister Cu ‘hit’ of a water-cooled maintenance. Operators are challenged to benchmark
furnace component, it is quite simply deemed that refractory continually against other established best tapping practices
or accretion freeze lining must always persist to prevent such and tap-hole management systems in order to seek further
direct matte/blister Cu contact with the copper tap-blocks, a incremental improvements in safety and performance.
condition somewhat analogous to protection against the Molenaar’s vision (Irons, 2001) of the tap-hole of the
physical cause of loss of the space shuttle Columbia, being a future, now well over a decade old, was of a fully automated
‘breach in the Thermal Protection system’ (Gehman et al., and remote-controlled environment. This effectively describes
2003). This disaster involved the loss of thermal protection operating with personnel safely removed to the maximum
tiles; the analogy to protective refractory layers in a extent possible from direct interface with hot liquids, their
composite copper cooling system is patent. A final critical containment systems (hot lining and environmental), and
warning served on all furnace operating and maintenance tapping systems. While progress has indeed been made in
personnel is a learning outcome from the Columbia disaster: this direction, further effort is required to realize such an
avoid falling into the trap of complacency by analogously ideal, consistent with still further improvements in tap-hole
‘deeming damage to the Thermal Protection System an performance and life, improvements that are pivotal to
“accepted flight risk”’ (Gehman et al., 2003). Any decision ensuring the safest and highest productivity furnace tapping
not to investigate thoroughly a suspected matte/blister Cu operation possible.
‘hit’, or breach of protective refractory and/or accretion freeze
lining, should always be challenged with vigour! >,AC3?F;95FAE<
Alternative coolants are suggested as a means to mitigate
Permission by Anglo American Platinum Ltd. to publish,
some of the risks associated with linings that use water
technical discussions and contributions by Reinoud van Laar,
cooling in high-temperature molten-bath systems (Kennedy
Paul den Hoed, Bart Pieterse, Alpheus Moshokwa, Phillimon
et al., 2013). Certainly until such cooling media achieve
Mukumbe, Joalet Steenkamp, Paul van Manen, and Whitey
common commercial application, effective water leak
Seyanund, and assistance with manuscript preparation by
detection is a vital safety requirement of designs that
Emerelda Kieser are gratefully acknowledged.
incorporate water-cooled linings. Monitoring of abnormal
drops in temperature in tap-holes or linings through cooling
by water (Jameson et al., 1999; Nelson et al., 2004) or of F7F@FA>F<
abnormal rises in temperature, either through conversion to ABRAMOWITZ, H., GOFFNEY, L.J., and ZIEGERT, W.L. 1983. Taphole mix properties
steam and its subsequent transport and heating effect in the and performance for the first year of operation on Inland’s No. 7 blast
furnace. Proceedings of the 42nd Ironmaking Conference, Atlanta, GA,
nearby environs, or through the loss of the freeze lining skull 17–20 April. Iron and Steel Society of AIME, Warrendale, PA.
(Entwistle, 2001), is another procedure adopted to identify pp. 681–694.
water leaks. Other methods involve off-gas analysis for
ANDŌ, J. 1985. Development of slag blast granulating plant characterised by
increased hydrogen content (in reducing ironmaking BF and innovation of the slag treatment method, heat recovery and recovery of
ferroalloy processes) or, directly, for water vapour using slag as resources. Mitsubishi Heavy Industries, Ltd., Technical Review,
hygrometers. June. p. 7.
L

 
       VOLUME 116  487
The tap-hole — key to furnace performance
ANDOU, T., NAGAHARA, M., TSUTSUI, N., IKEDA, M., NOMURA, M., and SUETAKI, T. ENTWISTLE, J.W. 2001. The tap-hole: A unique system. Proceedings of the
1989. Wear mechanism of taphole mix. Nippon Steel Technical Report, McMaster Symposium on Iron and Steelmaking. pp. 243–258.
no. 41. pp. 7–11. ESTRABILLO, L.S. 2001.Tap-hole management practices at Lake Erie steel
ATLAND, R. and GRABIETZ, G. 2001. Tap-hole design and experience. Proceedings company. Proceedings of the McMaster Symposium on Iron and
of the McMaster Symposium on Iron and Steelmaking. pp. 211–230. Steelmaking. pp. 233–241

BALLEWSKI, T., PETERS, M., RUTHER, P., and SCHMOLE, P. 2001. New improvements FALLAH-MEHRJARDI, A., HAYES, P., and JAK, E. 2014. From phase equilibrium and
of the tap-hole area at Thyssen Krupp Stahl AG. Proceedings of the thermodynamic modeling to freeze linings – the development of
techniques for the analysis of complex slag systems. Celebrating the
McMaster Symposium on Iron and Steelmaking. pp. 48–56.
Megascale: Proceedings of the EPD Symposium on Pyrometallurgy in
BELL, C., BOETTCHER, B.D., HRIBLJAN, F., KINSMAN, B.M., and VAN LAAR, F. 2004. Honor of David G.C. Robertson. TMS, Warrendale, PA. pp. 259–266.
Tap-hole maintenance improvements at Dofasco. AISTech, vol. 1.
GEHMAN, H.W., BARRY, J.L., DEL, D.W., HALLOCK, J.N., HESS, K.W., HUBBARD, G.S.,
pp. 303–321.
LOGSDON, J.M., OSHEROFF, D.D., RIDE, S.K., TETRAULT, R.E., TURCOTTE, S.A.,
BLACK, B. and BOBEK, J. 2001. Improvement in tap-hole life and casting WALLACE, S.B., and WIDNALL, S.E. 2003. Report of Columbia Accident
performance at Ispat Inland’s no. 7 BF. Proceedings of the McMaster Investigation Board. National Aeronautics and Space Administration,
Symposium on Iron and Steelmaking. pp. 156–176. Washington, DC. http://s3.amazonaws.com/akamai.netstorage/
anon.nasaglobal/ CAIB/CAIB_lowres_full.pdf
BROWN, R.W. and STEELE, D.F. 1988. Bond developments in silicon carbide blast
furnace refractories during the 1980’s. Aachen Proceedings. pp. 41–46. GEORGE, D.B. 2002. Continuous copper converting – A perspective and view of
the future. Sulfide Smelting 2002, Proceedings of a Symposium held
BRUNNBAUER, G., MAUHART, J., FERSTL, A., NOGRATNIG, N., and RUMMER, B. 2001. during the TMS Annual Meeting, 17–21 February. pp. 3–13. TMS,
Tapping practice at Voest Alpine Stahl Linz. Proceedings of the McMaster Warrendale, PA.
Symposium on Iron and Steelmaking. pp. 178–182.
GEORGE-KENNEDY, D., WALTON R., GEORGE, D.B., and NEXHIP, C. 2005. Flash
BUSSELL, B., JANZEN, J., ST. AMANT, M., MIRON, R., EMOND, M., BRAUN, W., and converting after 10 years. Proceedings of the 11th International Flash
GERRITSEN, T. 2013. Improved furnace cooling water pressure leak Smelting Congress, Bulgaria/Spain. pp. 79–97.
detection system at VALE. Proceedings of INFACON 13, Almaty,
GERRITSEN, T., SHADLYN, P., MACROSTY, R., ZHANG, J., and VAN BEEK, B. 2009.
Kazakhstan. pp. 377–384.
Tapblock fibre-optic temperature system. Proceedings of Pyrometallurgy of
CAMERON, I.A., SRIRAM, R., and HAM, F. 1995. Electric furnace matte tap hole Nickel & Cobalt 2009. pp. 627–639.
developments at Falconbridge Limited. CIM Bulletin, vol. 88, no. 991. GUDENAU, H.W., KRÖNERT, W., and FUSENIG, R. 1988. Wear profiles in blast
pp. 102–108. furnace hearth, giving particular consideration to the formation of blast
CAMPBELL, A.P., PERICLEOUS, K.A., and CROSS, M. 2002. Modeling of freeze layers furnace salamander. Aachen Proceedings. pp. 28–31.
and refractory wear in direct smelting processes. Iron and Steelmaker, GUEVARA, F.J. and IRONS, G.A. 2007. Slag freeze layer formation in an electric
vol. 29, no. 9. pp. 41–45. smelting furnace. Proceedings of Cu2007. Carlos Díaz Symposium on
CÁMPORA, S., DORO, E., GIANDOMENICO, F., and GONZALEZ, J.M. 1998. Pyrometallurgy, Toronto. vol. 3 (book 2). pp. 481–493.
Improvements on the cast house operation at Siderar’s blast furnace #2 GUTHRIE, R.I.L. 1992. Engineering in Process Metallurgy. Oxford Science
since the blow-in. Proceedings of the 57th Ironmaking Conference, Publications.
Toronto, 22–25 March 1998. pp. 203–217. HE, Q., ZULLI, P., TANZIL, F., LEE, B., and EVANS, G. 2001. An insight into flow
CASSINI, C. 2001. Tap-holes operating and face repair practices at Sollac- behaviour of blast furnace taphole streams. Proceedings of the McMaster
Méditerranée FOS Plant. Proceedings of the McMaster Symposium on Iron Symposium on Iron and Steelmaking, no. 29. pp. 149–153.
and Steelmaking. pp. 68–80. HE, Q., ZULLI, P., TANZIL, F., LEE, B., DUNNING, J., and EVANS, G. 2002. Flow
COETZEE, C. and SYLVEN, P. 2010. No tap-hole – no furnace. Proceedings of characteristics of a blast furnace taphole stream and its effects on trough
Refractories 2010. Southern African Institute of Mining and Metallurgy, refractory wear. ISIJ International, vol. 42, no. 3. pp. 235–242.
Johannesburg. pp. 55–66. HE, Q., EVANS, G., ZULLI, P., and TANZIL, F. 2012. Cold model study of blast gas
discharge from the tap-hole during the blast furnace hearth drainage. ISIJ
COETZEE, C., DUNCANSON, P., and SYLVEN, P. 2010. Campaign extension for
International, vol. 52, no. 5. pp. 774–778.
ferroalloy furnaces with improved tap hole repair system. Proceedings of
INFACON XII, Helsinki. pp. 857–865. HEARN, A.M., DZERMEJKO, A.J., and LAMONT, P.H. 1998. ‘Freeze’ lining concepts
for improving submerged arc furnace lining life and performance.
COETZEE, V. 2006. Common-sense improvements to electric smelting at Impala Proceedings of INFACON VIII, Beijing. pp. 401–426.
Platinum. Journal of the Southern African Institute of Mining and
HENNING, B., SHAPIRO, M., MARX, F., PIENAAR, D., and NEL, H. 2010. Evaluating
Metallurgy, vol. 106, no. 3. pp. 155–164.
AC and DC furnace water-cooling systems using CFD analysis.
DASH, S.K., JHA, D.N.S., AJMANI, K., and UPADHYAYA, A. 2004. Optimisation of Proceedings of INFACON XII, Helsinki. pp. 849–856.
tap-hole angle to minimise flow induced wall shear stress on the hearth.
HENNING, B., MARX, F., HARTZENBERG, D., FOWLER, N., and SHAPIRO, M. 2011.
Iron and Steelmaker, vol. 31, no.3. p. 207–215.
Simulating the blister tap hole concept design using conjugate heat
DE KIEVIT, A., GANGULY, S., and DENNIS, P. 2004. Monitoring and control of transfer capabilities in STAR-CCM+ V5.02.009. Proceedings of the STAR
furnace 1 freeze lining at Tasmanian electro-metallurgical company. European Conference, Noordwijk, The Netherlands, 22–23 March.
Proceedings of INFACON X, Cape Town. pp. 477–487. HERSHEY, R.A., STENDERA, J.W., and BIEVER, G.G. 2013. Sorting out the hazards
DE PAGTER, J. and MOLENAAR, R. 2001. Tap-hole experience at BF6 and BF7 of of anhydrous taphole clays. AISTech Conference Proceedings, Pittsburgh,
Corus strip products Ijmuiden. Proceedings of the McMaster Symposium May 2013. pp. 169–174.
on Iron and Steelmaking. pp. 107–125. HOPF, M. 2014. Monitoring the wear of water-cooled tap-hole blocks by the
fibre-optic method Optisave. Proceedings of the Furnace Tapping
DELABRE, PH., DUFOUR, A., GUENARD, C., HITIER, B., HUBERT, P., LE RUNIGO, I., and
Conference 2014, Misty Hills, Cradle of Humankind, Johannesburg, South
VENTURINI, M.J. 1991. Taphole mud: a refractory key control of blast
Africa. South African Institute of Mining and Metallurgy, Johannesburg.
furnace casting conditions. Proceedings of the Unite CR Congress.
pp. 33–49.
pp. 299–305.
HOPF, M. and ROSSOUW, E. 2006. New opportunities – Exhaustive monitored
DUNCANSON, P.L and SYLVEN, P. 2011. New refractory lining direction at Jindal copper coolers for submerged arc furnaces. Proceedings of Southern
stainless – first Indian FeCr producer to install UCAR Chillkote® linings. African Pyrometallurgy 2006. Jones, R.T. (ed.). Southern African Institute
GrafTech International. pp. 2–10. of Mining and Metallurgy, Johannesburg. pp. 89–100.
EDEN, M.G., FERSTL, A., and DOEL, P. 2001. The tap-hole zone of a blast furnace HORITA, S. and HARA, S. 2005. Excellent performance taphole mix. Shinagawa
– the critical region for a long campaign life. Proceedings of the Third IAS Technical Report, no. 48. pp. 47–50.
Ironmaking Seminar, Buenos Aires, Argentina. pp. 123–130.
HUBERT, P., FARDA, H., CADILHON, P., and POIRSON, G. 1991. The total quality
EDWARDS, B. and HUTCHINSON, S. 2001. Emergency tap-hole repair No. 4 Blast control installation: Improvement factor for tap hole clays. Proceedings of
Furnace at STELCO Hilton Works. Proceedings of the McMaster the Unified International Technical Conference on Refractories, 2 edn.
Symposium on Iron and Steelmaking. pp. 273–290. Aachen, FDR, 23–26 September. pp. 397–402.
L

488 VOLUME 116


 
      
The tap-hole — key to furnace performance
HUBERT, P., PHILIPPON, B., RUER, C., and LAMBERT, F. 1995. Evolution of taphole MERRY, J., SARVINIS, J., and VOERMANN, N. 2000. Modern furnace cooling design.
clays properties during ageing. Proceedings of the Unified International JOM, vol. 52, no. 2. pp. 62–64.
Technical Conference on Refractories, Kyoto, Japan, 19–22 November.
MILLS, K.C. and KEENE, B.J. 1987. Physical properties of BOS slags.
pp. 129–136.
International Materials Reviews, vol. 32, no. 1–2. pp. 1–44.
HUNDERMARK, R.J., NELSON, L.R., DE VILLIERS, B., NDLOVU, J., MOKWENA, D.,
MITSUI, H., TORITANI, T., YAMANE, S., OGUCHI, Y., and KAWAKAMI, T. 1988. Recent
MUKUMBE, P., PIETERSE, B., SEYANUND, W., and VAN MANEN, P. 2014.
developments in tap hole mud for blast furnaces. Aachen Proceedings.
Redoubling platinum group metal smelting intensity – operational
pp. 98–107.
challenges and solutions. Proceedings of the EPD Symposium on
Pyrometallurgy in Honor of David G.C. Robertson. TMS, Warrendale, PA. MULLER, J. and STEENKAMP, J.D. 2013. Evaluation of HCFeMN and SiMn slag
pp. 189–196. taphole performance using CFD and thermochemical property modelling
for CaO-MnO-SiO2-Al2O3-MgO slag. Proceedings of INFACON XIII,
IIDA, M., OGURU, K., and HAKONE, T. 2009. Numerical study on metal/slag
Almaty, Kazakhstan. pp. 385–392.
drainage rate deviation during blast furnace tapping. ISIJ International,
vol. 49, no. 8. pp. 1123–1132. NAKAMURA, R., SUMIMURA, H., and KITAMURA, M. 2007. Development of a new
resin-bonded taphole mix for high productivity blast furnaces. Shinagawa
IIYAMA, M., NUMATA, N., IMABEPPU, M., and MIWA, T. 1998. Microcracks caused
Technical Report, vol. 50. pp. 71–76.
by thermal stress in blast furnace carbon blocks. Aachen Proceedings.
pp. 19–21. NDLOVU, J., AMADI-ECHENDU, J.E., NELSON, L.R., and STOBER, F. 2005. Operational
readiness – a value proposition. Proceedings of Nickel and Cobalt 2005.
IRONS, G. (ed.). 2001. The tap-hole – the blast furnace lifeline: design /
Challenges in Extraction and Production. 44th Annual Conference of
maintenance / operating practices. Proceedings of the McMaster
Metallurgists of CIM, Calgary, Canada. pp. 389–404.
Symposium on Iron and Steelmaking. pp. 1–352.
NELSON, L.R. 2014. Evolution of the mega-scale in ferroalloy electric furnace
ISHITOBI, T., ICHIHARA, K., and HOMMA, T. 2010. Operational improvements of
smelting. Celebrating the Megascale: Proceedings of the EPD Symposium
submerged arc furnace in Kashima works KF-1 relined in 2006.
on Pyrometallurgy in Honor of David G.C. Robertson. TMS, Warrendale,
Proceedings of INFACON XII, Helsinki. pp. 509–515.
PA. pp. 39–68.
JAMESON, D., EDEN, M.G., and GORDON, R. 1999. The tap-hole zone – The critical
NELSON, L.R., GELDENHUIS, J.M.A., EMERY, B., DE VRIES, M., JOINER, K., and MA, T.
factor in long campaign life. Proceedings of the 58th Ironmaking
2006. Hatch development in furnace design in conjunction with smelting
Conference. Iron and Steel Society of AIME, Warrendale, PA. pp. 625–631.
plants in Africa. Proceedings of Southern African Pyrometallurgy 2006.
JASTRZEBSKI, M., KOEHLER, T., WALLACE, K., NOVIKOV, N.V., NOVIKOV, N., Jones, R.T. (ed.). Southern African Institute of Mining and Metallurgy,
ZAPOROZHETS, B., SHEVCHENKO, D., KRYZHANOVSKAYA, N., and KAPRAN, I. 2012. Johannesburg. pp. 417–435.
Redesign of furnace no. 1 at the Pobuzhsky Ferronickel Combine.
NELSON, L.R., GELDENHUIS, J.M.A., MIRAZA, T., BADRUJAMAN, T., TAOFIC HIDYAT, A.,
Proceedings of COM 2012, Niagara Falls. pp. 29-41.
JAUHARI, I., STOBER, F.A., VOERMANN, N., WASMUND, B.O., and JAHNSEN, E.J.M.
KADKHODABEIGI, M., TVEIT, H., and JOHANSEN, S.T. 2011. Modelling the tapping 2007. Role of operational support in ramp-up of the FeNi-II furnace at PT
process in submerged arc furnaces used in high silicon alloys production. Antam in Pomalaa. Proceedings of INFACON XI, New Delhi. pp. 798–813.
ISIJ International, vol. 51, no. 2. pp. 193–202.
NELSON, L.R., SULLIVAN, R., JACOBS, P., MUNNIK, E., LEWARNE, P., ROOS, E., UYS,
KAGEYAMA, T., KITAMURA, M., and TANAKA, D. 2005. Eco-friendly high M.J.N., SALT, B., DE VRIES, M., MCKENNA, K., VOERMANN, N., and WASMUND,
performance taphole mix. Shinagawa Technical Report, no. 48. pp. 41–46. B.O. 2004. Application of a high intensity cooling system to dc-arc
KAGEYAMA, T., KITAMURA, M., and TANAKA, D. 2007. Effects of ultra fine powder furnace production of ferrocobalt at Chambishi. Journal of the South
additions on taphole mix. Shinagawa Technical Report, vol. 50. African Institute of Mining and Metallurgy, vol. 104. pp. 551–561.
pp. 41–48. NEWMAN, C.J. and WEAVER, M.M. 2002. Kennecott flash converting furnace
KENNEDY, M.W., NOS, P., BRATT, M., and WEAVER, M. 2013. Alternative coolants design improvements – 2001. Proceedings of Sulfide Smelting 2002. TMS,
and cooling system designs for safer freeze lined furnace operation. Warrenadale, PA. pp. 317–328.
Proceedings of Ni-Co 2013, San Antonio, TX, 3–7 March 2013. NIGHTINGALE, R.J. and ROONEY, B.J. 2001. Damage to tap-hole and sidewalls of
pp. 299–314. Port Kembla no. 5 blast furnace – causes and remedies. No.3 tap-hole
KITAMURA, M. 2014. Optimising taphole clay technology. Shinagawa Technical spool breakout. Proceedings of the McMaster Symposium on Iron and
Report, vol. 57. pp. 1–6. Steelmaking. pp. 302–307.
KO, Y.C., HO, C.K., and KUO, H.T. 2008. The thermal behaviour analysis in NIGHTINGALE, R.J., TANZIL, F.W.B.U., BECK, A.J.G., and PRICE, K. 2001. Blast
taphole area. China Steel Technical Report. no. 21. pp. 13-20. furnace hearth condition monitoring and tap-hole management
techniques. La Revue de Mètallurgie. pp. 533–540.
LANGE, M., GARBERS-CRAIG, A.M., and CROMARTY, R. 2014. Wear of magnesia
chrome bricks as a function of matte temperature. Journal of the Southern NIGHTINGALE, S.A., WELLS, L., TANZIL, F., CUMMINS, J., MONAGHAN, B.J., and PRICE,
African Institute of Mining and Metallurgy, vol. 114, no. 4. pp. 341–346. K. 2006. Assessment of the structural development of resin bonded
taphole clay. ICSTI 06, International Conference on the Science and
LIOW, J-L., JUUSELA, M., and GRAY, N.B. 2001. Viscosity effects in the discharge
Technology of Ironmaking. Osaka, Japan. ISIJ. pp. 251–255.
of a two-layer liquid through an orifice. 14th Australasian Fluid
Mechanics Conference, Adelaide University, Adelaide, Australia, 10–14 NIIYA, Y., KITAMURA, M., and KAJITANI, A. 2012. Study on the adhesion
December 2001. pp. 853–856 properties of taphole mixes. Shinagawa Technical Report, no. 55, pp.
1–10.
LIOW, J.L., JUUSELA, M., GRAY, N.B., and ŠUTALO, I.D. 2003. Entrainment of a
two-layer liquid through a tap-hole. Metallurgical and Materials. NISHI, T. 2007. The buildup of additional HC FeMn production capacity by the
Transactions, vol. 34B. pp. 821–832. deepening of a furnace. Proceedings of INFACON XI, New Delhi.
pp. 183–190.
LUNGMUSS FEUERFEST. 2014. Lungmuß taphole clays for submerged arc & shaft
furnaces (company brochure). http://www.envicomab.com/wp- NOLET, I. 2014. Tapping of PGM-Ni mattes: an industry survey. Proceedings of
content/uploads/2014/08/Lungmuss-Taphole-Clays-for-SAF-May- the Furnace Tapping Conference, Misty Hills, Cradle of Humankind,
2014.pdf Johannesburg, South Africa. Southern African Institute of Mining and
Metallurgy, Johannesburg. pp. 12.
MACROSTY, R., NITSCHKE, S., GERRITSEN, T., and KARGES, A. 2007. Advances in
furnace monitoring: instrumentation. COM 2007. Proceedings of the 6th O’SHAUGHNESSY, P., WEI, D., GUOXHU, K., and SYLVEN, O. 2013. Improved furnace
International Copper-Cobre Conference, Toronto, Ontario, 25–30 August lining performance at Yiwang Ferroalloys. Proceedings of INFACON XIII,
2007. Vol. VII. pp. 203–216. Almaty, Kazakhstan. pp. 401–406.
MARX, F., SHAPIRO, M., and HENNING, B. 2005. Application of high intensity ÖSTLUND, P. 2001. Developments in tap-hole design, tapping methods and
refractory cooling systems in pyrometallurgical vessel design. Proceedings refractory repair over the last twenty years at SSÅB Oxselösund.
of INFACON XII, Helsinki. pp. 769–778. Proceedings of the McMaster Symposium on Iron and Steelmaking.
MATSUTANI, T. Not dated. The Mitsubishi Process – Copper smelting for the 21st pp. 322–322.
century. http://www.scribd.com/doc/127397992/The-Mitsubishi-Process- PAN, C.-N. and SHAO, C.-H. 2009. Development of anti-splashing taphole mud.
Copper-Smelting-for-the-21st-Century-Manuscript China Steel Technical Report, no. 22. pp. 48–52.
L

 
       VOLUME 116  489
The tap-hole — key to furnace performance
PEREZ, J.L., TAKEDA, K., SHIRAISHI, K., INOUE, F., and HONDA, N. 2001. STEIGAUF, C. and STORM, L. 2001. Optimization of tap-hole construction,
Development of non-polluting tap-hole mixture. Proceedings of the Unified maintenance, and casting practice for high productivity, high reliability,
International Technical Conference on Refractories, Cancun, Mexico, 4–7 and long campaign life. Proceedings of the McMaster Symposium on Iron
November. pp. 200–207. and Steelmaking. pp. 37–47.
PETRUCCELLI, A., VAN LAAR, F., and HRIBLJ, F. 2003. Alternating tap-hole practice STEVENSON, P. and HE, Q. 2005. Slug flow in blast furnace taphole. Chemical
with a two-tap-hole blast furnace. Proceedings of the ISS Technical Engineering and Processing, vol. 44. pp. 1094–1097.
Conference 2003. pp. 457–468. STOKMAN, R., VAN STEIN CALLENFELS, E., and VAN LAAR, R. 2004. Blast furnace
PIEL, K.-TH , WILKENING, S., LECHTHALER, K., and SANTOWSKI, K. 1988. Improved lining and cooling technology: experience at Corus IJmuiden. Association
carbon blocks for blast furnaces. Aachen Proceedings. pp. 22–25. for Iron & Steel Technology, Warrendale, PA. pp. 21–29.
PLIKAS, T., GUNNEWIEK, L., GERRITSEN, T., BROTHERS, M., and KARGES, A. 2005. The SUNDSTRÖM, A.W., EKSTEEN, J.J., and GEORGALLI, G.A. 2008. A review of the
predictive control of furnace tapblock operation using CFD and PCA physical properties of base metal mattes. Journal of the Southern African
modelling. JOM, Oct. 2005. p. 37–43. Institute of Mining and Metallurgy, vol. 108, Aug. 2008. pp. 431–449.
POST, J.R., TEETERS, T., YANG, Y., and REUTER, M.A. 2003. Hot metal flow in the SZEKELY, J. and DINOVO, S.T. 1974. Thermal criteria for tundish nozzle or taphole
blast furnace hearth: Thermal and carbon dissolution effects on buoyancy, blockage. Metallurgical Transactions, vol. 5. pp. 747–754.
flow and refractory wear. Proceedings of the 3rd International Conference SZYMKOWSKI, S.J. and BULTITUDE-PAULL, J.M. 1992. The production of high quality
on CFD in the Minerals and Process Industries, 10–12 December 2003. silicon metal and Simcoa. Proceedings of INFACON VI, Cape Town.
CSIRO, Melbourne, Australia. pp. 433–440. pp. 185–191.
PERSONAL COMMUNICATION. 1999. Mississauga, 29–30 March 1999. TANZIL, F.W.B.U., NIGHTINGALE, R.J., LEE, R.C., DUNNING, J.P., and RATTER, J. 2001.
PERSONAL COMMUNICATION. 2003. Rustenburg, 8–9 April 2003. The application of novel techniques and new sensors to tap-hole
management practice at Port Kembla no. 6 blast furnace. Proceedings of
PERSONAL COMMUNICATION. 2010. Johannesburg, 27–29 April 2010.
the McMaster Symposium on Iron and Steelmaking. pp. 193–207.
PERSONAL COMMUNICATION. 2011. Victoria Falls, 2–8 October 2011.
TOMALA, J. and BASISTA, S. 2007. Micropore carbon furnace lining. Proceedings
ROBERTSON, D.G.C. and KANG, S. 1999. Model studies of heat transfer and flow of INFACON XI, New Delhi. pp. 722–727.
in slag-cleaning furnaces. Fluid-Flow Phenomena in Metals Processing. TRAPANI, M.L., KYLLO, A.K., and GRAY, N.B. 2002. Improvements to tap-hole
El-Kaddah, N. (ed.). The Metallurgical Society, Warrendale, PA. pp. 1–20. design. Proceedings of the 3rd International Sulfide Smelting Symposium
RODD, L., KOEHLER, T., WALKER, C., and VOERMANN, N. 2010. Economics of slag (Sulfide Smelting '02), Seattle, Washington, 17–21 February 2002.
heat recovery from ferronickel slags. Proceedings of Sustainability for pp. 339–348.
Profit, COM2010, Vancouver. pp. 3–17. TSUCHIYA, N., FUKUTAKE, Y., YAMAUCHI, Y., and MATSUMOTO, T. 1998. In-furnace
RÜTHER, H.P. 1988. Refractory technology and operational experience with conditions as prerequisites for proper use and design of mud to control
tapholes and troughs of blast furnaces in the Federal Republic of blast furnace tap-hole length. ISIJ International, vol. 38. pp. 116–125.
Germany. Aachen Proceedings. pp. 60–66. UENAKA, T., YULUBO, Y., SHIMOMURA, K., YORITA, E., OHARA, K., and OMORI, H.
SADRI, A., GEBSKI, P., and GEORGE-KENNEDY, D. 2008. Development of the taphole 1989. An evaluation method of taphole mud for blast furnace. Shinagawa
acoustic monitoring (TAM) system for water-cooled copper tapblocks. Technical Report, no. 32. pp. 29–44.
Proceedings of COM 2008, Winnipeg. pp. 7–19. VALENTAS, L.S. and THIERNEY, E.P. 2010. Furnace panel leak detection system.
SAGER, D., GRANT, D., STADLE, R., and SCHREITER, T. 2010. Low cost ferroalloy US patent 7,832,367 B2. 16 Nov. 2010.
extraction in DC-arc furnace at Middleburg Ferrochrome. Journal of the VAN DER WALT, N.J. 1986. The design and development of furnace linings for
Southern African Institute of Mining and Metallurgy, vol. 110. manganese alloys smelting. Proceedings of INFACON IV, Rio de Janeiro.
pp. 717–724. pp. 17–30.
SHAO, L. 2013. Model-based estimation of liquid flows in the blast furnace VAN IKELEN, J.P., DE PAGTER, J.J., and BELLEMAN, G.T.J. 2000. Experience with
hearth and taphole. Doctor of Technology thesis, Abo Akademi University, tap-hole drilling technology. Proceedings of the 4th European Coke and
Turku/Abo, Finland. (ISBN 978-952-12-2941-1). Ironmaking Congress, Paris, 19–21 June. vol. 2. pp. 610–616.
SHAO, L. and SAXEN, H. 2011. A simulation study of blast furnace hearth VAN LAAR, K. 2001. The tap-hole: the heart of the blast furnace. Proceedings of
drainage using a two-phase flow model of the tap-hole. ISIJ International, the McMaster Symposium on Iron and Steelmaking. pp. 1–21.
vol. 51, no. 2. pp. 228–235
VAN LAAR, R. 2014. Personal communication.
SHAO, L. and SAXEN, H. 2013a. A simulation study of two-liquid flow in the
VAN LAAR, R.J., HÖLL, H., DZERMEJKO, A.J., RISI, J., and MALAN, J. 2001. Improving
taphole of the blast furnace. ISIJ International, vol. 53, no. 6.
profitability by analysing ferroalloys furnace lining performance to
pp. 988–994.
improve lifetime. Proceedings of INFACON IX, Quebec City. pp. 475–485.
SHAO, L. and SAXEN, H. 2013b. Flow patterns of iron and slag in the blast
VAN LAAR, R., VAN STEIN CALLENFELS, E., and GEERDES, M. 2003. Blast furnace
furnace taphole. ISIJ International, vol. 53, no. 10. pp. 1756–1767.
hearth management for safe and long campaigns. ISSTech 2003
SHAW, A., DE VILLIERS, L.P.VS., HUNDERMARK, R.J., NDLOVU, J., NELSON, L.R., Conference Proceedings. pp. 1079-1090.
PIETERSE, B., SULLIVAN, R., VOERMANN, N., WALKER, C., STOBER, F., and
VEENSTRA, R., VOERMANN, N., and WASMUND, B. 1997. Prototype metal tap block
MCKENZIE, A.D. 2012. Challenges and solutions in PGM furnace operation:
design. Proceedings of Nickel-Cobalt 97, Montreal. vol. 3. pp. 595–607.
High matte temperature and copper cooler corrosion. Journal of the
Southern African Institute of Mining and Metallurgy, vol. 113. VOERMANN, N., GERRITSEN, T., CANDY, I., STOBER, F., and MATYAS, A. 2010.
pp. 251–261. Developments in furnace technology for ferro-nickel production.
Proceedings of INFACON X, Cape Town. pp. 455–465.
SHENG, Y.Y., IRONS, G.A., and TISDALE, D.G. 1998. Transport phenomena in
electric smelting of nickel matte: Part II. Mathematical modelling. WALKER, C., KASHANI-NEJAD, S., DALVI, A.D., VOERMANN, N., CANDY, I.M., and
Metallurgical and Materials Transactions B, vol. 29B. February. WASMUND, B. 2009. Nickel laterite rotary kiln-electric furnace plant of the
pp. 85–94. future. Proceedings of COM 2009, Sudbury. pp. 19.

SINGH, B.B, SURESH KUMAR, P.G., and MAHATA, S.K. 2007. Modern practices of WASMUND, B.O. 2003. Personal Communication.
post tap-hole operation in ferro chrome production and its advantages. WELLS, L.J. 2002. The rheology of a composite polymer ceramic plastic
Proceedings of INFACON XI, New Delhi. pp. 530–538. refractory. PhD thesis, University of Wollongong, Australia.
SMITH, M., FRANKLIN, S., and FONSECA, F. 2005. Improving blast furnace YAMASHITA, M., KASHIWADA, M., and SHIBUTA, H. 1995. Improvement of tap holes
operations by co-ordinated improvements in tap-hole clay, guns and drills. at Wakayama no. 5 blast furnace. Proceedings of the 54th Ironmaking
Proceedings of the 5th AIS Ironmaking Conference. Instituto Argentino de Conference, Nashville, TN. Iron and Steel Society, Warrendale, PA.
Siderurgia, pp. 197–202. pp. 177–182.
SPREIJ, M., TIJHUIS, G.J., TROW, J., and FRANKEN, M.C. 1995. A quantitative ZHOU, J. and SUN, L. 2013. Flash smelting and flash converting process and
selection of blast furnace hearth refractories. Proceedings of the Unite CR start-up at Jinguan Copper. Proceedings of Copper 2013, Santiago.
Congress. pp. 167–175. pp. 199–212. N
L

490 VOLUME 116


 
      

View publication stats

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy