Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Energy Conversion and Management 46 (2005) 563–575

www.elsevier.com/locate/enconman

Use of turbines for simultaneous pressure regulation and


recovery in secondary cooling water systems in deep mines
H.J. van Antwerpen *, G.P. Greyvenstein
School of Mechanical and Materials Engineering, North West University, Potchefstroom Campus, Private bag X6001,
Potchefstroom 2520, South Africa
Received 30 January 2004; accepted 29 April 2004
Available online 7 June 2004

Abstract
A system is proposed that uses a reverse-running multistage pump acting as a turbine to do simultaneous
pressure regulation and pressure recovery in secondary cooling water systems in deep mines. The turbine
drives a standard induction motor that acts as a generator. A constant pressure drop across the system is
maintained with bypass and inline valves. For design evaluations, a simple method was developed for
calculation of turbine head/flow and efficiency relations.
Ó 2004 Elsevier Ltd. All rights reserved.

Keywords: Mine refrigeration; Energy recovery; Reverse running; Multistage pump; Secondary cooling water; Pressure
relief; Turbine; Mathematical modelling

1. Introduction

A major operating cost of South African deep gold mines is refrigeration, which is mainly done
with chilled water piped down the mines. The descent of the water releases large amounts of
potential energy. This energy is either dissipated into heat by pressure reducing valves or can
partly be recovered with energy recovery equipment such as turbines [1].
Pelton turbines are well established equipment for energy recovery in primary water distribu-
tion systems. They are always installed above primary storage dams, as their operating principle
requires an atmospheric outlet pressure. However, in the secondary water system between primary
storage dams, pressure reducing valves are still used, presenting a further opportunity to recover

*
Corresponding author. Tel.: +27-18-297-0326; fax: +27-18-297-0318.
E-mail address: mgihjva@puk.ac.za (H.J. van Antwerpen).

0196-8904/$ - see front matter Ó 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2004.04.006
564 H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575

Nomenclature

a turbine modelling constant


b rotor passage thickness [m]
c fluid absolute velocity [m/s]
d diameter [m]
g gravitational acceleration [m/s2 ]
h head [m]
H pressure head [m]
k turbine modelling constant
N rotational speed [rpm]
Q volume flow rate [m3 /s]
r radius [m]
s number of pump or turbine stages
u local rotor velocity [m/s]
w fluid relative velocity [m/s]
y efficiency modelling constant
Subscripts
in inlet
l loss
out outlet
opt optimal
r radial
rot rotor
sys system
t turbine
x circumferential direction
Greek symbols
a rotor vane/radial angle [degrees]
c turbine modelling constant
/ flow coefficient
g efficiency
f turbine modelling constant
x angular speed [rad/s]

energy. Such a secondary pressure relief system should maintain a sufficient outlet pressure for
stoping water usage, ruling out the use of a Pelton turbine.
Other options are either turbine/generator sets or displacement pumping systems like the 3-
champer pipe feeder system [2] in which downward flowing cold water is used directly to displace
hot water to a higher level. Because of its operating principle, such a system requires integration of
hot and cold water networks, as well as matching flow rates, which is not practical at secondary
H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575 565

pressure relief stations. In contrast, the location and operation of a turbine/generator set is in no
way affected by the hot return water route or flow rates. Displacement pumping is also a relatively
recent development, while turbine/generator technology is already well established so that at
present, a turbine/generator system is the most practical option for secondary pressure relief and
energy recovery.
Previous work on secondary energy recovery did not focus on energy recovery from the total
secondary water stream but rather focused on using water pressure in the secondary network to
drive various machines. Ferguson and Bluhm [3] reported on the successful use of a turbine to
reduce the water pressure for an air cooling spray chamber, while driving the recirculation pumps
of the cooling chamber. Ramsden and Bluhm [4] tested small water turbines to substitute electric
motors for driving fans in portable air coolers. Another recent development is the use of water
driven stope drills.

2. Turbine selection

Other industrial processes where turbines can replace pressure reducing valves are cryogenic
systems, petrochemical refinery processes and seawater desalination. Custom built turbines used
in these industries to substitute pressure relief valves are either multistage standard speed units
[5,6] or high speed single stage units [7,8]. However, reverse running multistage pumps can also be
used for this purpose, albeit with a lower peak efficiency.
Laux [9] summarized the application of multistage pumps as turbines. Advantages of reverse
running multistage pumps are listed as large, adjustable output range by altering the number of
stages, low manufacturing cost due to pump standardisation, short delivery time due to pump
mass production, relatively low runaway speed and smaller absolute speed compared to Pelton
turbines. Torbin [10] compared reverse running multistage pumps with Pelton turbines for mining
energy recovery, noting a further advantage of multistage pumps in the mining context, namely
that mine maintenance personnel are familiar with multistage pumps as they are commonly used
to pump water to the surface. The head ranges of reverse running multistage pumps also include
the head ranges of secondary pressure reducing stations, which typically vary from 150 up to
1000 m [11]. In view of the above advantages and their robustness, reverse running multistage
pumps are considered the best option for secondary energy recovery turbines. For generality in
this study, no specific optimal flow rate is used, but the turbine mode performance of the available
multistage pumps is investigated over their respective flow ranges.
With the turbine type selected to be a reverse running multistage pump, a general model of its
characteristics is required to investigate its interaction with the other system components.

3. Derivation of a turbine model

In order to utilize manufacturer supplied turbine mode data easily and to obtain characteristics
at speeds other than synchronous speed, mathematical turbine models based on the underly-
ing physics are required. The models thus required are the head/speed/flow ðHt ; N ; QÞ and
566 H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575

efficiency/speed/flow ðg; N ; QÞ relationships of a reverse running multistage pump. The model for
head will be derived first.
For this analysis, a multistage turbine is considered to consist of inlet guide vane sections,
rotors and outlet return vane sections. For a turbine with s stages, the sum of the head losses in
these sections will equal the total head loss across the turbine, so that
Ht ¼ sðhin þ hrot þ hout Þ ð1Þ
with hin and hout , respectively, the inlet guide vane section and outlet return vane section losses,
while hrot represents the head loss across the rotor. Neither the inlet guide vane nor outlet return
vane sections comprise any moving parts, so the most significant head losses in those sections are
simply proportional to the flow rate squared and can be expressed as
ðhin þ hout Þ / Q2 ð2Þ
With the head change in the rotor, the rotation and backward curved profile of the rotor vanes
come into play. To establish the relation of rotor head change to speed and flow, Euler’s equation
is used:
ghrot ¼ u1 c1;x  u2 cx;2 ð3Þ
It is assumed that the flow leaves the rotor in a radial direction. Therefore, cx;2 ¼ 0, so
ghrot ¼ ucx ð4Þ
from which it follows that the head is proportional to the product of rotor tip velocity u and fluid
circumferential velocity cx .
According to the velocity triangle in Fig. 1, the fluid circumferential velocity cx can be written as
cx ¼ u  wx , but the relative velocity wx also equals cr tan a. The fluid radial velocity cr is deter-
Q
mined by the flow rate: cr ¼ 2prb , where r is the radius and b is the flow passage thickness. cx can,
therefore, be rewritten again as
Q
cx ¼ u  cr tan a ¼ u  tan a ð5Þ
2prb
Rotor tip velocity u equals xr, which can be substituted with Eq. (5) in Eq. (4) to obtain

u
wx cx
cr
c
w α

Fig. 1. Velocity triangle for a reverse running pump rotor.


H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575 567
 
Q
ghrot ¼ xr xr  tan a ð6Þ
2prb
Rearrangement gives
r2 tan a
hrot ¼ x2  xQ ð7Þ
g g2pb
Considering the fact that x equals 2pN
60
and d ¼ 2r, the above relation is simplified to show only the
functional relation of head to rotational speed and flow rate:
d2 tan a
hrot / N 2  NQ ð8Þ
g dg
Frictional head losses in the rotor are also related to the square of flow rate, so with inclusion of
Eq. (8), the head loss across a turbine, Eq. (1) has the following form:
Ht ¼ sðk1 Q2  k2 NQ tan a þ k3 N 2 Þ ð9Þ
with k1 to k3 being constants per stage and s the number of stages. The diameter d is included in
the constants because the performance will not be scaled geometrically. The model will only be
used for extrapolating manufacturer supplied performance curves.
Though the rotor vane angle is also a constant geometric feature of each turbine, it has been
included in the constants to demonstrate the effect of the vane angle on the turbine characteristics.
When the vanes are purely radial, i.e. a ¼ 0, k2 QN tan a becomes zero, so the general turbine
characteristic becomes:
Ht ¼ sðk1 Q2 þ k3 N 2 Þ ð10Þ
which is the same as the turbine model proposed by Kimmel [5]. He stated without deduction that
conservation of energy dictates the following for a turbine:
Ht ¼ cQ2 þ 1N 2 ð11Þ
where Ht is differential head, N rotational speed in rpm and Q volume flow rate. f and c are design
specific constants determined from simple empirical tests. It is thus concluded that the model
posed by Kimmel [5] was derived for a turbine with radial rotor vanes, as opposed to a multistage
pump that has backward curved vanes. Since the constants in Eq. (9) will be determined uniquely
for each turbine, the vane angle can be included in the constants, so that the eventual form of the
characteristic is
Ht ¼ sða1 Q2  a2 QN þ a3 N 2 Þ ð12Þ
with a1 , a2 and a3 empirical constants and s the number of stages.
The turbine model is demonstrated by fitting it to experimentally determined characteristics
presented by Bohl [12] shown in Fig. 2. A discussion of the shape of the head/speed/flow chart can
be found in Kimmel [5].
The turbine model describes the head/flow curves excellently as shown in Fig. 2 and by a
correlation coefficient R2 of 0.9995. The head/flow model is, therefore, considered suitable for this
study.
568 H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575

250

200

e
Head [m]
2500

qu
150

tor
d

ro
ee
2000 sp

Ze
100
Zero
1500
50
1000
0
0 10 20 30 40 50
Flow rate [m^3/h]

Fig. 2. Experimental (solid) and predicted (dashed) head/flow curves for the turbine data presented by Bohl [12].

4. Efficiency relationships

Head and flow rate data are scaled according to speed, but the corresponding efficiency curve
cannot be scaled exactly the same. The experimental results of Bohl [12], shown in Fig. 3, are used
as the starting point to find a scaling method.
On the head/flow chart in Fig. 3, the iso-efficiency lines are equally spaced between the zero
torque and zero speed lines, while the constant speed head lines converge to the zero speed line
after maximum efficiency.
It appears that the iso-efficiency lines can be approximated with parabolas. It will now be
shown that the head/flow lines for constant flow coefficient / (defined as Q=Nd 3 ) are also para-
bolic. By substituting the relation N ¼ Q=/d 3 in Eq. (12) and rearranging, the following is ob-
tained:
 2 !
1 1
Ht ¼ s a1  a2 3 þ a3 Q2
/d /d 3

This result shows that for a constant flow coefficient, head is a parabolic function of flow. This
indicates that the efficiency could be modelled well as a function of the flow coefficient. Indeed, the
efficiency data from Fig. 3 correlates very well to Q=N as shown in Fig. 4. The diameter d only has
to be taken into account when data is scaled to another machine size.
An appropriate modelling approach will be to replace the curves with a single polynomial
curve, as that will provide a simple calculation method. An important requirement for such an
efficiency polynomial is that it must be positive only between the zero torque and zero speed lines
while being zero at these lines. On account of this, only second and fourth-order polynomials will
be considered, as polynomials of uneven order always assume very large positive values. As shown
in Fig. 4, a fourth-order polynomial fits the data from all four efficiency lines well.
One may question the adequacy of this polynomial model on account of the sharp decrease at
Q=N ¼ 0:05 while the efficiency lines appear not to drop to zero. It must be taken into account
that region A in Fig. 4 corresponds to region A on the head/flow chart (Fig. 5), while region B
corresponds to the much smaller region B on the head/flow chart. This shows that in terms of
head and flow rate, region B is quite insignificant, the only requirement being that the efficiency
should decrease from the peak to zero within that region. Also, efficiency accuracy in region A is
H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575 569

250

200

Head [m]
%
68
150 %
% 60
40

%
ue ee
d

20
100 rq
to sp
ro ro
Ze
50 Ze

0 0.8

Efficiency
0.6

00

00
00

00
0.4
10

20
15

25
0.2

0
0 10 20 30 40 50
Flow rate [m^3/h]

Fig. 3. Head/flow rate chart and efficiency/flow rate chart of a six stage pump in turbine mode. Adapted from Bohl [12].

1
A B
0.8
Efficiency

0.6

0.4

0.2
Zero torque Zero speed
0
0 0.01 0.02 0.03 0.04 0.05 0.06
Q/N

Fig. 4. Efficiency lines for different speeds as a function of Q=N [12]. A fourth-order polynomial is fitted to the efficiency
line data. Regions A and B correspond to the same regions on the head/flow chart in Fig. 5.

250
cy
ien
200 A e ffic
ak
Head [m]

d B
ue

150 Pe ee
sp
rq
to

Z ero
ro

100
Ze

50

0
0 10 20 30 40 50
Flow rate [m^3/h]

Fig. 5. Illustration of the different efficiency areas on the head/flow chart. With reference to Fig. 4, region A lies be-
tween zero torque and peak efficiency, while region B lies between peak efficiency and zero speed.
570 H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575

of much greater importance than its accuracy in region B, so the model shown in Fig. 4 will give
satisfactory results for techno-economical evaluations.
Efficiency will thus be modelled by fitting a fourth-order polynomial to manufacturer supplied
efficiency curves, with head and speed described by Eq. (12). The efficiency polynomial is pre-
sented in
gt ¼ y1 /4 þ y2 /3 þ y3 /2 þ y4 / þ y5 ð13Þ
with / equal to Q=N and constants y1 to y5 are determined from experimental data and regression
analysis.

5. System design alternatives

An important operational requirement for the pressure relief system is that it must change its
own pressure drop characteristic according to the downstream flow demand, so that the down-
stream pressure remains constant. The two viable options are to regulate the pressure by varying
the turbine speed or to operate the turbine at constant speed with inline and bypass pressure
regulating valves. In the rest of this section these two options will be discussed in detail.
The two alternatives will be evaluated for an installation having an optimal turbine operating
point at 30 m head and 28 l/s flow rate, which, for ease of comparison, coincides exactly with the
maximum efficiency of a single stage SULZER HPH 24 pump running in reverse at 1500 rpm. The
comparison is done for a speed of 1500 rpm in order to stay within acceptable speed limits in
the case of variable speed.

6. Variable turbine speed

A typical mine system would require that the flow rate vary while the pressure drop is kept
constant at 30 m as indicated on Fig. 6. To vary the flow rate at 30 m head, the turbine speed has
to vary from 2200 rpm for 15 l/s to 900 rpm for 31 l/s. Variation of the speed is viable by using a
doubly fed induction generator system in which the exciting frequency is changed according to the
rotational speed in order to always maintain the synchronous output frequency [13].
Fig. 6 indicates that the turbine has a limited flow capacity at constant head, which necessitates
a bypass valve to cope with higher flow rate demands. On the other end, low flow rates require the
turbine to shut down and be shut off with an isolating valve so that the bypass valve can control
the flow. The resulting system layout is shown schematically in Fig. 7.
In the calculation of the system’s steady state efficiency, the turbine differential head is fixed, the
flow rate is varied and the speed is solved implicitly by using the head/speed/flow relation (Eq.
(12)). The flow rate/speed data pairs obtained from Eq. (12) are then used as input to calculate the
turbine efficiency using Eq. (13) at flow rates lower than optimum. The speed is only varied up to
the optimal flow rate, where the bypass pressure reducing valve opens to handle the excess flow
rate, so that the turbine flow, speed and power stays constant at turbine optimal condi-
tions. Because of the energy dissipation in the bypass valve, the system efficiency diminishes
H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575 571

50
0
250
40

e
qu
0
200

Head [m]

tor
30

ro
Ze
0
150
20
0
100
10
p e ed
s
Zero
0

0.8
Constant speed

Efficiency
Variable head 0.6
1500 rpm
0.4
Variable speed
Constant head 0.2

0
0 5 10 15 20 25 30 35 40
Flow rate [l/s]

Fig. 6. Speed and efficiency characteristics of a HPH 24 reverse running pump.

Constant
supply head Variable
Turbine speed
generator

Bypass
regulating valve Isolating valve

Varying demand

Fig. 7. Schematic layout of a variable turbine speed pressure reducing system.

1
System efficiency

0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60 70
Flow rate [l/s]
Constant speed Variable speed

Fig. 8. Efficiency versus flow rate for two pressure reducing systems using a HPH24 reverse running pump either
operating at a constant speed of 1500 rpm or at variable speed.
572 H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575

proportionately to the ratio Qt;opt =Qsys with Qt;opt the turbine flow rate at the optimum point and
Qsys the system flow rate. At flow rates larger than optimal, system efficiency is, therefore, given by
Qt;opt
For Qsys > Qt;opt : gsys ¼ gt  ð14Þ
Qsys
Thus, Eqs. (12) and (13) or (14) are solved simultaneously for system efficiency. Fig. 8 shows the
calculated system efficiency for the variable speed system, as well as the constant speed system,
which will now be discussed.

7. Constant turbine speed

An alternative to variable turbine speed for pressure control is to use a constant speed turbine
and affect pressure control with inline and bypass pressure reducing valves as described by Laux
[9]. Constant speed electricity generation is simply done by driving an induction motor faster than
synchronous speed, i.e. with positive slip, so that standard electrical machinery can be used.
Operation of the turbine system is now explained with reference to the head/flow chart in Fig. 6.
The optimal flow rate for maximum efficiency is indicated in Fig. 6 at 28 l/s, and it is apparent
that at flow rates less than optimal, an inline valve is needed to sustain the reduced pressure drop.
At flow rates larger than optimal, the turbine pressure drop exceeds 30 m so that a bypass valve is
required as in the case with the variable speed system. At low flow rates, when the turbine is shut
down, the inline pressure reducing valve also acts as an isolating valve. The resulting system
layout is exactly the same as shown in Fig. 7, with only the inline isolating valve changing to a
regulating valve and the generator changing to constant speed.
Because of the inline pressure relief valve, the turbine does not utilize the full system head at
flow rates below optimum, causing the system efficiency to be lower than the turbine efficiency as
calculated with the following expression:
Ht
For Qsys < Qt;opt : gsys ¼ gt  ð15Þ
Hsys
At flow rates higher than optimal, the system efficiency is also given by Eq. (14) because of the
bypass valve. The system efficiency is then obtained by solving Eqs. (12) and (13), simultaneously
with either Eq. (14) or (15) while speed is set constant. The system efficiency obtained for the
HPH24 case study is presented in Fig. 8.
Fig. 8 indicates the variable speed system has a marginally wider peak efficiency range, while
the constant speed system has a wider low efficiency range, so depending on the shape of the flow
volume/flow rate distribution, one may recover more energy with one configuration than with the
other.
Apart from recovery efficiency, reliability and therefore system complexity also has to be taken
into account. The variable speed system comprises a slip ring induction generator with a wound
rotor instead of a squirrel cage rotor, power electronics and an intricate controller, whereas the
constant speed system uses a simple squirrel cage induction generator and its industry standard
ancillaries. This means that the constant speed system is much less complicated than the variable
speed system.
H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575 573

Constant speed Variable speed

Induction generator,
Induction generator, slip
Generator type standard squirrel cage
ring wound rotor type
rotor type

Relative generator Generally 75% more than


Used as reference
cost constant speed

On/off, speed control


system, comprising
Generator control On/off only
pulse-width modulating
power electronics.

In line valve type Pressure reducing valve Isolating valve

Automatic mechanical Automatic mechanical


Valve control
pressure-sensing pressure-sensing

Relatively high compared


System complexity Used as reference to the constant speed
system.

Better midrange
Recovery
Used as reference efficiency than constant
efficiency
speed system.

Fig. 9. Summary of the comparison between the constant and variable speed systems.

The level of system complexity has a profound impact on maintenance costs in that it generally
increases the cost and frequency of maintenance. Maintenance of underground machinery is
particularly complicated, as their remote location requires quite an effort to get the necessary
equipment to the machine site. Scheduled maintenance is still manageable because it can be
planned outside mining shift times, while unscheduled maintenance poses a real threat to prof-
itability, particularly in the case of critical machinery since the loss of ore production is a far
greater loss than the direct cost of repairing the machine. By this reasoning, a measure of machine
energy efficiency could rather be sacrificed to gain reliability. Thus, in the present case, the
constant speed system is superior to the variable speed system on account of its simplicity.
The constant speed system also has a lower initial cost, as the price of a slip ring generator is
almost double that of a squirrel cage generator [14]. Apart from the generator, the variable speed
system also has power electronics and control circuitry that are not required with the constant
speed system, while the constant speed system only has a little more sophisticated inline valve. The
costs of these auxiliary components are relatively low in comparison to the generator, so the
generator cost is a decisive factor showing the variable speed system to be the most costly option.
The results of the above comparisons are summarized in Fig. 9.
The above analysis establishes the constant speed system as the most suitable for mine appli-
cation. The control methodology of the chosen approach can now be considered in more detail.

8. Modes of operation

As the turbine operating flow rate range is limited, more than one mode of system operation is
required to cover the full range of flow rates. For this system, three modes of operation can be
574 H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575

50 Mode A: Mode B: Mode C:


Inline: closed Inline: regulating Inline: open
40 Bypass: regulating Bypass: closed Bypass: regulating

Head [m]
Turbine OFF Turbine ON Turbine ON
30

20 pm]
1500 [r
ue
rq d
10 to pee
os
Ze
ro Zer
0
0 5 10 15 20 25 30 35 40
Flow rate [l/s]

Fig. 10. The three operating modes for a constant speed pressure reducing turbine.

identified as indicated in Fig. 10, namely mode A at flow rates lower than the turbine switch-on
flow rate; mode B at turbine operating flow rates and mode C at flow rates higher than the turbine
optimal flow rate.
During mode A operation, the turbine bypass valve regulates the pressure while the turbine
inline valve is closed. For the system to switch from mode A to mode B, the flow rate has to be
sufficient for the recovered energy to exceed the turbine and generator mechanical losses. When
switching from mode A to mode B, the inline valve starts regulating and the bypass valve closes.
As Laux [9] described startup for such a configuration, the generator starts as an electric motor
and changes to the generator mode when the turbine speed increases sufficiently to exceed the
synchronous speed. The generator torque angle changes so that power is fed into the grid, while
the power grid dictates the operating speed by means of the exciting frequency. With an increase
in flow rate, the inline valve opens more until the optimal flow rate is reached, and transition to
mode C operation happens when the bypass valve has to open again and regulate the excess flow.
Pressure reducing valves are easily controlled to open, close or regulate by changing the dif-
ferential pressure set point, so a simple electronic controller will be able to manage the turbine
startup and control the switching between operating modes. The actual pressure control is still
done by automatic mechanical action of the valves, and as this pressure control method has been
standard practise for years, no untried methods are used in this pressure reducing solution.
Therefore, there is little doubt about the technical feasibility of constructing and operating such a
system.

9. Conclusion

Two workable configurations to do energy recovery have been identified and evaluated, namely
a constant turbine speed system and a variable turbine speed system, with the constant speed
system emerging as the best option for mine duty on account of its simplicity while maintaining an
acceptable level of efficiency over a relatively wide flow range. The proposed system can be built
with off-the-shelf components using existing mining expertise, leaving little doubt about the
technical feasibility of recovering energy from the total stream of secondary cooling water.
H.J. van Antwerpen, G.P. Greyvenstein / Energy Conversion and Management 46 (2005) 563–575 575

Acknowledgements

The authors wish to thank the Deep mine Collaborative Research Programme whose financial
support made this work possible.

References

[1] Bluhm SJ. Deep mine Task 6.5.3 Report: Energy Recovery Equipment. Bluhm Burton Engineering, Johannesburg,
South Africa.
[2] SIEMAG. Fluid transportation services. Available from: http://www.mining-technology.com/contractors/winding/
siemag/index.html#siemag5 (Accessed December 2003).
[3] Ferguson DWB, Bluhm SJ. Performance testing of an energy recovery turbine at a three stage spray chamber.
J Min Ventilat Soc S Afr 1984;37(11):121–6.
[4] Ramsden R, Bluhm SJ. Energy recovery turbines for use with underground air coolers. In: Proceedings of the 2nd
US Mine Ventilation Symposium, Reno, NV, 23–25 September 1985. p. 571–80.
[5] Kimmel HE. Speed controlled turbine expanders. Hydrocarbon Engineering, May/June 1997. Palladian
Publications Limited, UK; 1997.
[6] Prescott M. HPRT developments improve payout. Hydrocarbon Process 1990;(January):53–4.
[7] Gopalakrishnan S. Power recovery turbines for the process industry. In: Proceedings of the Third Pump Users
Symposium; 1986. p. 3–11.
[8] Pump engineering. The hydraulic turbocharger; 2001. Available from: http://www.pumpengineering.com (Accessed
December 2003).
[9] Laux CH. Reverse-running standard pumps as energy recovery turbines. Sulzer Tech Rev 1982;2:23–7.
[10] Torbin RN. Alternate methods of energy recovery for the mining industry. IEEE Trans Ind Appl 1989;25(5):811–8.
[11] Butterworth M. Deep mine Task 6.5.1 Report: Futuremine. CSIR Miningtek, Johannesburg, South Africa; 2001.
[12] Bohl W. Str€omungsmaschinen 1, Aufbau und wirkungsweise. 7. Auflage. Vogel Buchverlag; 1998. p. 312.
[13] Van Staden CO. Energy recovery in a deep mine with an asynchronous induction motor. M.Eng Dissertation.
Potchefstroom University for Christian Higher Education. Work in progress, 2002.
[14] Zest Motors. Product price list; 2001. Available from: http://www.zest.co.za/price/index.htm (Accessed October
2003).

You might also like