Anda di halaman 1dari 135

COMMENTARY ON

SANS 10208:3 – 2001 Edition 3

THE DESIGN OF STRUCTURES FOR THE


MINING INDUSTRY
PART 3: CONVEYANCES

GJ Krige
Principal Structural Engineer
Anglo Technical Division

SOUTHERN AFRICAN INSTITUTE OF STEEL CONSTRUCTION


JOHANNESBURG, REPUBLIC OF SOUTH AFRICA
Copyright © 2001 SAISC
July 01

The Southern African Institute of Steel Construction is an autonomous, non-profit


organisation, funded by the subscriptions of its members. It has a permanent staff
operating from Johannesburg and is active countrywide. The Institute operates in the
technical and promotional fields and does not involve itself with contractual, business
or labour matters. It has three fundamental objectives:

• to advance the state of knowledge and expertise relating to the design and
construction of steel structures,

• to promote the use of steel in construction in preference to alternative materials,

• to guard over the interests of the steel construction industry with respect to
legislation, technical manpower, public relations, etc.

Although care has been taken to ensure, to the best of our knowledge, that all data
and information contained herein is accurate to the extent that it relates to either
matters of fact or accepted practice or matters of opinion at the time of publication,
neither the author nor the SA Institute of Steel Construction assumes any
responsibility for any errors in, or misinterpretations of, such data and or information
or any loss or damage arising from or related to its use.

Acknowledgement

This Commentary arises from the dedicated activity of the Institute's Regulations
Codes & Standards Committee and in particular that sub-committee responsible for
the SABS 0208 series of Codes of Practice for the Design of Structures for the
Mining Industry. The fact that the substantial contributions made by these committee
members are on a voluntary basis, makes their achievements that much worthier.
The author and the SAISC express their gratitude to all those who have made both
SABS 0208 and this Commentary a reality. Their efforts are much appreciated by the
South African Steel Industry and all those engineers who are responsible for the
design of structures for the mining industry.

i
July 01

FOREWORD

Codes of practice play a prominent and extremely important role in the work of the
structural designer. As a means for capturing in quantitative form what is known
about the behaviour of structural elements and disseminating this information widely
they are unequalled. Their value in bringing order to the design process and
promoting consistency in the quality of design calculations can hardly be
overestimated.

And yet, codes are widely criticised by structural designers, often with justification.
Criticism tends to centre on the following matters:
• Codes are legalistic and tend to force designers to follow set rules, thus
discouraging engineering judgement and innovation born out of an in-depth
understanding of structural behaviour.
• Codes often contain outdated information.
• Whilst appearing very authoritative, any code is actually far from perfect, and it
may lead designers to commit errors, for example when rules are applied to
situations where they are not applicable.
• Codes often demand from the designer more calculation than is needed to
establish the adequacy of a particular design, and every new generation of codes
seem to be more demanding of designers' time.
• The introduction of a new code inevitably results in a degree of confusion, and
expense, such as for the purchase or development of new design aids. There
may also be strife when at least some designers are not convinced of the need
for change.

The fundamental requirement for conciseness, and the fact that codes are produced
only after the laborious interaction of many people with divergent views and priorities,
makes it inevitable that any code will attract comments like these. One thing can
however be employed to soften the adverse effects and weaknesses of a new code:
a good commentary. A commentary allows the authors of a code the opportunity to
communicate with the users in a less formal and cryptic manner than demanded by
the style commonly adopted for codes. Thus the commentary affords the user some
insight into the reasoning behind every clause, so that it can be interpreted correctly,
and its limitations be known.

Inevitably, a commentary becomes somewhat of an extension of the code it deals


with. Additional information is given, and the interpretation of the code requirements
is made more precise. This will have unavoidable legal implications in any case
where the adequacy of a design is questioned, thus lending extra weight to the need
for designers to familiarise themselves with the contents of the commentary.

The purpose of this commentary on the Code of Practice for the Design of Structures
in the Mining Industry — SABS 0208 Part 3 Conveyances is as follows:

• To provide additional information which might be of use to designers.

ii
July 01

• To provide information regarding the origins of, and reasons for, the code
requirements.

• To explain the requirements of the code and its implications to the user where it is
considered that explanation may be helpful.

• To highlight and justify changes from the previous version of the code.

It is hoped that this commentary will facilitate the use of the code, and generate a
greater interest in discussion about the behaviour of conveyances, which can lead to
an even better code, that will ensure safer, more reliable, and perhaps more
economical conveyances, in the future.

iii
July 01

INTRODUCTION

SANS 10208 — The code of practice for the design of structures for the Mining
Industry, is currently published in four parts viz.:

Part 1 Headgear structures


Part 2 Stages
Part 3 Conveyances
Part 4 Shaft System Structures

This commentary has been written to cover part 3, Conveyances, and is intended to
provide a better understanding of the design loads and design procedures employed
in their structural design. References to "the code" are thus to SANS 10208:Part 3
"Design of Structures for the Mining Industry: Conveyances".

The code was introduced with the aim of achieving the following objectives:

1. Defining current practice clearly and succinctly, to facilitate new development.


Design codes have been accused of retarding innovation, but it is the hope that in
defining current practice, its strength and shortfalls will become clearer, and that
engineers will be enabled to more rationally and confidently introduce innovations.

2. Enabling more engineers to competently design the conveyances. Many structural


and mechanical engineers have the basic skills required to design conveyances,
but are not well informed about the specific requirements, and may thus easily
overlook certain items. This design code will help to address this difficulty.

3. Introducing a unified design procedure which is acceptable to all the mining


companies. In the past there have been substantial differences between the design
procedures and requirements of different mining companies. This has led to
confusion amongst suppliers, differing expectations, and even different margins of
safety at different mines. A unified design approach will benefit the whole industry.

Conveyances are among the most important pieces of mine equipment; their proper
design, maintenance, and workmanlike repair are essential to continued smooth
operations on a mine. Mancages, large and small, rate the highest in terms of safety.
Thousands of men are transported daily in mancages and their safety is of
paramount importance. Shaft accidents due to poorly manufactured or poorly
maintained mancages may be rare, but everyone should strive to eliminate such
accidents completely. The entire production of the mine must be hoisted by means of
skips which often work very full duty cycles, under severe loading and environmental
conditions. Other conveyances, too, are essential to the commissioning or operation
of South Africa’s mines.

1
July 01

As conveyances are guided by portions of shaft system structures, which are covered
in Part 4 of the code, there are definitions and clauses common to both Part 3 and
Part 4. Rather than demanding frequent cross-referencing, a certain amount of
material is repeated in both parts of the code.

In drafting the code, an effort was made to incorporate as much as possible of the
most recent knowledge and results of research undertaken under the guidance of the
mining industry in South Africa. There are however several instances of loads which
have been defined without the benefits of recent research. In such cases the
definition of loads has relied on the long operational experience of the South African
mining industry. These loads are defined at magnitudes which would lead to sizes of
members which are commonly known to be acceptable and reliable. The committee
has attempted to write the code in a form which is simple and explicit, and
furthermore to avoid unnecessary changes from procedures which are familiar to
designers. Much debate has taken place in the drafting committee in an effort to
reconcile differing design philosophies, so that the code can hopefully be universally
acceptable to the South African mining industry.

The numbering of paragraphs in this commentary corresponds to the


numbering of clauses in the code. Extra information dealing with legal and
safety aspects, fabrication, and other matters, is also given in addition to the
clause clarification.

LEGAL REQUIREMENTS

The Mines Health and Safety Act refers back to the Mines and Works Act and
Regulations, which describes minimum criteria for the construction and maintenance
of skips, cages and attachments. Regulations 16.11 to 16.15, inclusive, refer to the
Construction of Winding Plant Conveyances and Regulations 16.16 to 16.19 refer to
the Connection of Winding Plant Conveyances. These regulations require that:

• Conveyances shall be of substantial construction, and in addition cages shall be


fitted with a roof and securely fitting doors.

Consequently, design procedures have been developed on behalf of the mines, and
maintenance repair schedules have been introduced by the mining companies to
ensure that conveyances are kept in good and safe working order in the interest of
the mine and the safety of the persons travelling in the conveyances. Possibly the
most important phrase used in the Regulations is that the equipment shall be free
from visible defects as clearly these can lead to disastrous consequences. There are
many defects which are invisible to the untrained eye, and some which are invisible
even to trained personnel. "Visible" includes detectable by appropriate non-
destructive testing procedures. Manufacturing and maintenance criteria are, however,
not considered within the scope of the code, but because of their importance and
influence on design they are discussed in this commentary.

2
July 01

In terms of the Mines Health and Safety Act, the code forms an interpretation of how
conveyances are to be designed in order to be of "substantial construction".
Compliance with the code is thus deemed to satisfy the requirements of the Mines
Health and Safety Act.

DESIGN PHILOSOPHY AND SAFETY FACTORS

The safety critical components (bridles, transoms and attachment hooks) are
designed with high safety factors. This is not only to comply with the mining
regulations but also to ensure that the operating stress is kept low to ensure a long,
safe working life for these components. Traditionally, a minimum safety factor of 10
against either the ultimate, or the yield, strength of the material has been used. In
numerical stress terms, for Grade 300W steel which has a minimum yield strength of
300 MPa, a maximum working stress of 30 MPa was implied by this design
procedure.

The less critical components (e.g. cage floor beams and bracing, skip side plates
and stiffeners) are designed with lower safety factors. If these elements were
designed with the same safety factor as the bridles and transoms, the
conveyances would become extremely heavy. This would lead to increased
power demand or reduced payloads and larger diameter ropes. Since the
winding system is balanced (an empty skip moves down the shaft when a loaded
skip moves upwards) the running torque of the winder is not affected by the
conveyance mass except by low friction effects. However, the conveyance mass
does affect the total system inertial mass and therefore the peak power demand
during acceleration.

Generally, the less critical components have a safety factor of approximately 1,7
against the yield strength of the material (e.g. a tensile working stress of 185 MPa is
permitted for Grade 300W steel). Since the safety factors for these components are
reduced, they require as much care and attention as the more important components.

The code introduces three concepts which are fairly new to the design of
conveyances. The first is the introduction of limit states design procedures, which
replace the use of allowable stress design which has been the standard. The second
is that emergency loads, which are used for the design of the major structural
members, are no longer based on using a factor of safety of 10. They are now based
on the maximum rope load derived from a rational assessment or simulation, of the
entire winder/rope/conveyance system. For fixed rope winders this is likely to be the
rope break load, but for friction winders it may well be significantly less than the rope
break load. The third concept is linking maintenance conditions and operating
practices in the shaft with the design loads applied to the conveyances.

1. What are the major differences in philosophy between "allowable stress"


and "limit state" design methods? In the former, the stress in a loaded
structural member is not permitted to be higher than prescribed levels.

3
July 01

Maximum allowable stresses are restricted to provide an adequate "safety


factor" to guard against beam flange or web buckling, shear buckling,
column instability, or any other failure mode. In Limit States Design, various
factors are applied to the loading and the member properties to ensure that
the structural element will not collapse under a particular loading
combination. This "load factor" method is a more logical design approach
since the structural engineer can consistently predict the level of loading to
cause collapse. In allowable stress design methods, there are different
safety factors for bending, shear and axial loading and it is not always clear
which type of loading would cause failure, especially when the three types
of loading act in combination with each other.

There has been resistance to the introduction of limit states codes, especially
from older engineers who may be reluctant to learn the new design techniques.
The argument that "..... none of my previously designed structures collapsed"
has been used. Perhaps so, but the behaviour of structures is generally more
predictable and in many cases material can be used more efficiently by using
limit states codes.

The limit state design philosophy allows the definition of loads and load factors
by this code, and then design essentially in accordance with the steel or
aluminium codes in current use for structures in South Africa. It has been
necessary to introduce some specific modifications in section 6, but these have
been kept to a minimum.

2. The main reason for using maximum rope load rather than a safety factor of
10 in this new design procedure, is minimising the risk of a rope break
disaster following a slack rope incident. The design philosophy to be used
under accident or severe misuse conditions was debated at length by the
committee drafting the code. This new design for maximum rope load
reflects the philosophy adopted, which is that the main structural members
must not fail under the most adverse loading possible, whereas even quite
severe damage to secondary members can be accepted as a result of
accidents or severe misuse. Where the rope break load is used as the
maximum rope load, it is known that rope break usually occurs due to
kinking or accidental cutting of the rope, at loads well below the rope
breaking strength, but it was concluded that a lower load could not be
logically justified for design. Using rope break as the emergency design
load is based on recognising that if the conveyance falls any distance, the
likelihood of the rope breaking can be reduced by ensuring the maximum
possible energy absorption capacity. Typically, the rope has a high energy
absorption capacity whereas the conveyance has a very low capacity.
Should the conveyance transom fail first, the total energy absorption
capacity is small, as shown by the shaded portion in figure 1(a). However,
the new design procedure ensures that the rope fails first, maximising the
total energy absorption capacity, as shown in figure 1(b).

4
July 01

So, there are fundamental differences between the old approach and this code
regarding emergency loads:

- The old method used the weight of the conveyance and a permissible elastic
stress of 1 10 of the ultimate tensile strength of the conveyance transom
material. The major flaw was that the method did not recognise the
magnitude of loading which could actually be applied because the rope
strength was disregarded.

- In the approach of this code the maximum loading level possible is


considered and matched with the ultimate strength of the transom. (i.e. the
plastic collapse load). This will always give a factor of safety of at least 10,
but the factor may be much higher.

Load Load
Failure
y
Failure y

Conveyance Rope Conveyance Rope


Deflection Deflection Deflection Deflection

(a) Conveyance Failure (b) Rope Failure

Figure 1: Energy Absorption

3. The design procedure, in particular the loads which are specified are dependent
on both the maintenance condition of the shaft and the operating procedures.
Thus, for example, the vertical impact loads during loading of conveyances
depend on the type of rope stretch compensation in use. Also the lateral roller
and slipper loads depend directly on the shaft misalignment which can be kept
small with good shaft maintenance. The clause clarification section will
elaborate on these and other cases.

5
July 01

INFLUENCE OF NEW PROCEDURES ON CONVEYANCE DESIGN

It is worth considering how the introduction of this code is likely to affect conveyance
designs.

In the code, provision is made for appropriate operating loads, which is unlikely to
affect the mass of items such as floor beams, corner angles, bracing members, sides
and doors. These loads are similar to those typically used in the past, so the main
change in design of these items is in the change to limit states procedures. It is not
expected that this will significantly alter the resulting member sizes.

The provision for emergency conditions such as a slack rope event mean that
transoms and bridles need to be designed to resist maximum rope loads which may
be equal to rope break loads in order to ensure that these critical items will not
collapse prior to the rope breaking. This is a significant increase on the loads used in
the traditional method of designing the conveyances.

Checks on the design of old conveyances, either in the process of refurbishment or


which were designed using the old "safety factor of ten" philosophy, invariably show
that transom strengthening is necessary. The information in table 1 demonstrates
why this is so. These examples assume a maximum rope load equal to the rope
break load.

Notes to Table 1

1. Gr 300W steel is assumed here.

2. For allowable stress design, the design load is the maximum conveyance
weight, W. For the new limit state design, the design load is 1,1 x 1,05 x R,
where R is the specified rope strength.

3. Allowable stress design uses the elastic section modulus, whereas limit states
design allows the use of the plastic section modulus, but with a material factor
of 0,9. The ratio of plastic section modulus to elastic section modulus is typically
approximately 1,15 for the H, I and C sections normally used for transoms.

4. For a shallow shaft, the rope strength must be 10 times the maximum
conveyance weight, whereas for a deep shaft it will be more. As an example, it
is assumed here that R = 12,5W for a specific deep shaft.

5. The rope capacity factor is the nominal rope break strength divided by the load
suspended at the end of the rope. The rope safety factor is the nominal rope
break strength divided by the sum of rope self weight and load suspended at the
end of the rope.

6. The bridle section usually has a low stress, as practical details require a
minimum size. To allow space for guides, slipper plates and slipper clearance, a
minimum of 160 mm is required inside the bridle. This requires a channel at
least 220 m deep. 240 mm deep channels are usually used.

6
July 01

Table 1: Comparison of Design Methods


(a) General results Design Method

Allowable Stress
with Factor of Limit state design
Safety = 10 for Rope Break

Ultimate Tensile 450 x 106 N m2 450 x 106 N m2


Strength (1)
Yield Stress (1) 300 x 106 N m2 300 x 106 N m2
Maximum design 450 x 10 6 10 = 45x10 6 N m2 300 x 106 N m2
stress, sm
Design Load (2) w 1,155 R

Transom bending WL 4 = 0,250 WL 1,155RL 4 = 0,289 RL


moment, M
Required section (3) Ze = M s m Zp - M sm x 0,9
0,250 WL 0,289 RL
= =
6
45 x 10 0,9 x 300 x 10 6
−6
= 0,00556 x10 WL = 0,00107 x 10 -6 RL
Shallow shaft (4) Ze = 0,00556 WL x 10-6 Ze = Zp 1,15 = 0,00093 RL x 10-6
Deep shaft (4) Ze = 0,00556 WL x 10-6 Ze = 0,00931 WL x 106

Ze = 0,01163 WL x 106
Bridle tension W 2 = 0,5 W 1,155 R 2 = 0,578 R
Required section A = 0,5W (45 x 106) A = 0,578 R (0,9 x 300 x 106)
= 0,0111 W x 10-6 = 0,0214 R x 10-6
Shallow shaft (4) A = 0,0111 W x 10-6 A = 0,02141 W x 10-6
Deep shaft (4) A = 0,0111 W x 10-6 A = 0,02676 W x 10-6

(b) Specific example of 180 man cage

Maximum weight 250 700 N 250 700 N


Rope strength 2 990 x 103 N 2 990 x 103 N
Capacity factor (5) 2 990 x 103 250 700 = 11,93 = 11,93
Transom Ze 2,51 x 10-6 m3 5,01 x 10-6 m3
Transom section 2 457 x 191 x 67 I 2 610 x 229 x 101 I
Transom mass 242 kg 364 kg
Bridle A 2,79 x 10-3 m2 6,40 x 10-3 m2
Bridle section (6) 240 x 85 [ + 240 x 12P 240 x 85 [ + 240 x 12P
Bridle mass 281 kg 281 kg

7
July 01

As an example of the two design procedures, consider a two deck cage, which has a
mass of 12 955 kg and carries 180 men, giving a payload of 12 600 kg. The total
weight is thus 250 700 N. The cage is supported on a 62 mm diameter rope with a
breaking strength of 2990 x 103 N and a mass of 16,8 kg m. The transom has a
length of 1,8 m. The details of this design are shown in Table 1b.

Therefore, for this particular cage where the capacity factor is 11,9, there is a difference of
98% in the elastic moduli for the two design methods. Capacity factors are higher than 10
for rope selections dictated by the safety factor rather than the capacity factor and member
sizing will become more and more problematic with deep shafts (unless rope factors are
relaxed). Energy absorbing devices are another way to address the problem, by reducing
the maximum loads which can be applied to transoms. However, no such devices are in
use yet.

The mass increase in the transom would be 122 kg in this case.

The new limit state method will always result in larger sections for the transoms of
conveyances in deep shafts, because the rope strength is more than 10 times the
load carried. Figure 2 shows the variation of this capacity factor with shaft depth for
the rope considered above.

Capacity Factor
30

20

10
1780m

0
0 1000 2000 3000
Shaft Depth (m)

Figure 2: Varying Capacity Factor

Thus, for a shaft which is shallower than about 1 800 m, there should not be much
change in transom size and mass. Small conveyances, such as inspection cages,
suspended on large ropes, will also require much heavier transom and bridle
members than previously.

Drawing information from various comparative design exercises which have been
undertaken, the likely mass implications of the code may be summarised in
percentage form as in Table 2.

8
July 01

Table 2: Mass Implications of the Code


Conveyance Member Shallow Shaft < 1 600 m Deep Shaft > 1 600 m

Skip
Transom mass 0% to +10% +10% to +30%
Bridle mass 0% 0% to +15%
Body mass -3% to +3% -3% to +3%
Total Skip mass -1% to +4% 0% to 6%

Cage
Transom mass 0% to +10% +10% to +30%
Bridle mass 0% 0% to +20%
Body mass -2% to +2% -2% to +2%
Total Skip mass -1% to +3% +1% to +7%

Note that the values given in Table 2 are dependent on what previous design method
is used in the comparison, and on the size of the conveyance and rope. In general
the lower mass increases are for smaller conveyances, accurately sized ropes, and
where the previous design used a factor of safety of 9 or 10 against yield. The higher
increases are generally for larger conveyances, oversized ropes, and where the
previous design used a factor of safety of 9 or 10 against ultimate strength, or lower
factors of safety. The values quoted are also based on using the same grade of steel
in old and new designs.

The provisions of the limit state code appear to allow use of existing bridle sections
without any need to be strengthened. Generally, these members appear to be
proportioned to suit dimensional requirements and tend to be slightly overdesigned.
This is less likely to be true for deep shafts or very large conveyances where
strength, rather than minimum dimensions will govern the design.

MAJOR CHANGES FROM 1994 CODE

The major changes to be implemented in this second edition of the code are as
follows:

1. The definitions have been brought in line with the definitions used in Part 4 of
the code.

2. Four different shaft zones have been defined, as a mechanism of dealing with
the differing conditions of rock strain in shafts where shaft pillar mining is taking

9
July 01

place. Requirements in these four shaft zones give guidance for the
conveyance loads at, near, and remote from the reef intersection zone.

3. Roof loads have been added for cage design, to ensure a robust roof structure..

4. Rock loads in kibbles have been added, as there are distinct differences
between these and the rock loads in skips.

5. Several difficulties were apparent with the blanket application of rope break
loads as the emergency load on all conveyances. A number of different
conveyance types are now treated differently in clause 5.6.1.

6. Vertical friction load during slipper plate contact on the guide has been added.

10
July 01

CLAUSE CLARIFICATION

1 SCOPE
The scope of the code is intended to include all types of conveyances in
the permanent shaft configuration. This leads to a fairly wide scope, but it
is felt to be appropriate. The frequent multiple use of conveyances would
require confusing cross-referencing of codes if the scope was limited to
only certain conveyances. Many of the load cases and design concepts
are also identical, or very similar, for most types of conveyances.
Equipping conveyances and kibbles are also included within the scope of
this code, because of many similarities of loads and design concepts.

In order to make the code as inclusive as possible, it also covers


conveyances operating in vertical shafts and incline shafts. It is thus
important to distinguish which loads act vertically, and which act along, or
perpendicular to, the axis of the shaft. This is elaborated under the
appropriate loads.

Specific allowance is made for the use of acceptable analytical


procedures. In numerous clauses, the code requirements have been kept
simple, at the expense of leading to conservative results in certain
instances. These clauses are identified in this commentary, and alternative
analytical procedures are suggested and illustrated. This provision will
sometimes enable the engineer to move away from the specific
requirements of this code, in order to achieve an improved design. It also
enables designers to introduce innovations, or deal with unusual situations
which are not included in the provisions of the code.

It should be pointed out that loads specified in this code are not intended
to be employed for the design of conveyance ropes, sheaves, or other
associated equipment. Applicable reference should be sought from Mine
Health and Safety Act (Act 29 of 1996) and Regulations to the Mines and
Works Act (Act 27 of 1956).

3. DEFINITIONS AND SYMBOLS

3.1 Definitions

Bridle Hanger, Bottom Transom, Top Transom

These members make up the main structural harness, and are the most
important load carrying members of the conveyance. A typical example of
these members is shown on the layouts of a man material cage in figure
3, and a skip in figure 4.

11
July 01

Top Transom
Roof Beam
Roof Beam
Bridle Hanger
Floor Beam
Floor Beam
Bridle Hanger

Bridle Hanger

Bridle Hanger
Bridle Hanger
Floor Beam
Floor Beam

Bridle Hanger

Bottom Transom Floor Beam

Figure 3: Typical Cage Layout

Top Transom

Pivot Bar

Loading Lip
Bridle Hanger

Bridle Hanger

Bridle Hanger

Bottom Transom Door

Figure 4: Typical Skip Layout

Under normal operating conditions, these members carry most of the


loads applied to the conveyance. Under emergency conditions it is vital
that these members do not fail, as this would lead to a major disaster with
the conveyance falling down the shaft. The load path for any underslung

12
July 01

loads, the loads at the bottom of the conveyance, and tailrope loads in the
case of a friction winder, is that they are applied initially to the member (or
members) making up the bottom transom which transfers them to the
bridle hangers. The bridle hangers, which are usually single members on
each side of the conveyance, and which may or may not be, integral with
the sides of the conveyance, then carry the loads to the top of the
conveyance. The member (or members) making up the top transom then
carry the load to the point at which the rope attachments transfer it into the
rope.

The term "crosshead" is commonly used in the mining industry to refer to a


top transom, but it is given a more specific usage in this code, as indicated
in the definitions.

Typically, top transoms are made of two channels, back-to-back, but


spaced apart to allow the rope attachments to fit between them. Bridle
hangers are usually made of channel sections, or two angles and a plate
bolted together to form a channel shape. This "three piece" bridle hanger
is used because of unavailability of channel sections in the required steel
grades, and to prevent possible fatigue cracks developing right across the
bridle hanger. Typical transom and bridle sections are shown in figure 5.

(a) TYPICAL TRANSOM SECTION (b) TYPICAL BRIDLE SECTIONS

Figure 5: Typical Transom And Bridle Members

Conveyances

A list is provided of all the conveyance types included within the scope of
the code. Note that counterweights are not strictly conveyances, as they
do not convey anything in the shaft, but the code uses the generic term
"conveyances" to include counterweights.

Inspection cages were previously commonly placed above cages and


even skips. This practice has now been largely replaced by inspection

13
July 01

cages below the main conveyance because this allows inspection right to
the bottom of the guides, and if guides have been damaged this is
identified before the main conveyance passes through the damaged area.
Placing an inspection cage below the main conveyance also provides
better protection against falling objects.
Skips are separated into top discharge, or overturning, skips and bottom
discharge. The term bottom discharge is fairly loosely used, and should be
understood to include side discharge skips as well. Overturning skips are
seldom specified for new shafts, but they are still encountered in older
shafts, so it was felt that they must be included in this code.

Safety Critical Components


Certain members of conveyances are defined as being safety critical
components, because their failure could lead to a major disaster. The
failure, at any time, of the bridle hangers or one of the transoms could lead
to the conveyance plunging down the shaft. This has been know to
happen, in some cases with great loss of life, following slack rope
incidents. A slack rope incident occurs if the tension in the rope is lost,
usually due to the conveyance jamming in the shaft while the winder
continues to pay out rope. The extra weight of rope can then dislodge the
conveyance which falls some distance under gravity.

The jack catch devices are also safety critical components because they
support the conveyance following rope detachment in an overwind.

Jack Catch Device


The jack catch is a safety catch which is located in the headgear near the
top of the guide path, and serves to prevent the conveyance falling down
the shaft after the rope has been detached following an overwind. It
consists of a series of hinged lugs which project into the compartment. The
jack catch device, located at the top of the conveyance catches in these
lugs to hold the conveyance.

Rope attachments
Typical rope attachments are the following:
Drawbar — A high-grade steel link between the top transom of the
conveyance and the rope end attachment or detaching hook.
Shackle — A connecting device between ropes and drawbars which
consists of a U shaped section pierced for a cross bolt or a pin.
H-link — A connecting device between the ropes and the drawbars which
consists of a forged, machined, "H" section and two cross-bolts or pins.
Detaching hook — An appliance which automatically releases the
winding rope from the conveyance should an overwind occur. The hook is
placed between the conveyance drawbar, and the winding rope cappel or

14
July 01

thimble. A detaching plate (spectacle plate) is fitted to girders below the


winding sheave, and in the event of an overwind, the hook becomes
locked in the plate, thus suspending the skip or cage, while the winding
rope is simultaneously detached.
Cappel — A fitting at the end of the winding rope connecting the winding
rope to the detaching hook. The cappel may consist of a taper wedge and
locking device or a high grade steel container into which white metal or
resin is cast around the rope.
Thimble — A fitting at the end of the winding rope connecting the rope to
the detaching hook and arranged to distribute the rope end load without
undue stress concentration. The thimble is the component which fits
securely into the eye of the spliced rope, through which the connecting pin
passes.
Pins — Circular fittings to provide a shear connection between the
transom or deadeye and the drawbar, and between the drawbar, H-link or
shackle and the rope.
Deadeye — A fitting below the transom which serves to distribute the load
from the drawbar into the transom.
Compensating sheave — A sheave which may be located within the
transom or at the top of a drawbar, and serves to equalise the tension in
the two ropes of a BMR winder. In new installations, this function is
commonly provided by means of hydraulic sheave supports in the
headgear, thus eliminating this heavy attachment.
Doubling down sheave — A sheave which may be located within the
transom or at the top of a drawbar and is used during doubling down of the
rope, either during rope installation, or to allow for statutory rope tests, or
to tension the rope coils on the winder drum.

Winders and Winding


The operation of moving people, equipment materials and rock in a mine
shaft is referred to as winding throughout the code. The winder is the
motor and drums, which provide the power and move the hoisting rope.
Conveyances are usually operated in pairs on a single winder, so that an
empty conveyance, or a counterweight, counterbalances a full
conveyance. Kibbles and small Maryanne cages are sometimes operated
independently of another conveyance.
Winders in common use are shown schematically in figure 6. They are:
Double drum winder — Each conveyance is attached to a single rope,
which is in turn attached to the winder drum, being wound onto the drum to
lift the conveyance.

Blair Multi-Rope winder (BMR) — Each conveyance is attached to two


ropes, which are attached to the winder drums as for the double drum

15
July 01

winder. It is important with BMR winders that the tension in the two ropes
is equalised.

Sheaves Sheaves

Winder Winder

Conveyance Conveyance

Conveyance Conveyance

(a) Double Drum (b) Blair Multi-Rope

Winder
Sheave

Winder

Conveyance
Conveyance

(d) Single Drum


Conveyance
Tailrope

(c) Friction

Figure 6: Different Winders In Common Use

16
July 01

Friction (Koepe) winder –- Each conveyance is attached to two or more


ropes, which pass over the winder drum, and are then attached to the
other conveyance. This winder relies on friction between the ropes and the
drum in order to operate, so the tension must be maintained in the ropes.
Hence the use of "tailropes" which are ropes, or chains, suspended
between the undersides of the two conveyances.
Single drum winder (Winch) — A winder operating a single conveyance
which is attached to a single rope, which is wound a single drum.
The "winder system" is the term used in the code to include a single
winder and the ropes, sheaves and conveyances attached to it.

Legal requirements are that connection ropes and links to conveyances


and trailers should be free from visible defects, be of adequate calculated
strength and be correctly heat treated after every six months of use,
unless made of an approved steel. Exemption from heat treatment of rope
attachments manufactured from steels to BS.2772 Part 2 (1977), 150M12
and 150M19 has been granted by the Government Mining Engineer, to
individual mines. Heat treatment requirements are currently under review.

3.2 Symbols
The list of symbols is a comprehensive list defining all the symbols which
are used throughout the code.
In general, capital letters are used for loads and forces, small letters for
lengths, velocities, accelerations and ratios, and Greek symbols are used
for factors, such as impact factors (α) and load factors (γ).
Note that the units used throughout the code, and thus throughout the
commentary, are the SI system, ie:

Mass - kg
Force - N
Length - m
Velocity - m/s
Acceleration - m/s2

17
July 01

4. MATERIALS

4.1 Steel

Most of the conveyances in use on South Africa's mines are constructed


mainly, or at least partially, of structural steels. Other steels which have
been considered include:

• 3CR12. This is a low chromium, corrosion resisting steel. It is used


chiefly because of its enhanced corrosion resistance, although it does
not appear to perform well in the acidic water environment of many gold
mines. When used in skips, it appears to exhibit somewhat better wear
characteristics than structural steel.

It would appear that one of the chief reasons for the continued limited use
of these steels is their poor availability. Some are only available in plate, or
a very limited range of sections.

The widespread use of shafts as ventilation passages for cold, downcast


air, leads to operation at quite low temperatures being common. The
hoisting operation also unavoidably leads to impact loads, which can be
quite high. It is thus common practice to specify impact resistant structural
steels for the safety critical components, although this is not done by all
the mining companies. The code thus recommends, but does not
prescribe, the use of steels with a Charpy impact resistance of 27J at 0ºC.
This again appears to lead to poor availability of the specified steel grade.
It is thus common practice to acquire a lower grade of steel and perform
impact tests which are used to determine whether particular sections must
be rejected or may be utilised.

The use of commercial quality steel must never be contemplated, as its


quality is not guaranteed, so its properties are uncertain and they may be
very variable.

4.2 Aluminium Alloys

In modern mancages, a large percentage of material used may be


structural aluminium. It is still common for transoms and some bridles to be
constructed in steel, but aluminium has great mass saving advantages and
some aluminium alloys have tensile strengths as good as, and sometimes
better than, carbon steels. Aluminium is also being used for other
conveyances, particularly skips. Aluminium is however more likely than
steel to exhibit fatigue cracks.

All sections, angles I sections, etc are extruded, whilst sheet and plate
products are rolled. Code references to extrusions are thus referring to
sections, and references to rolled products imply sheet or plates.

18
July 01

When aluminium components need to be replaced, ensure that the correct


alloy is used. All plates and profiles in mine stores should be clearly
marked, particularly the heat treated products which may lose strength if
incorrect fabrication procedures are used.

The recommended alloys are given below:

Aluminium Alloys Suited to use in Mining Conveyances

The five most common structural alloys readily available in South Africa
and their common applications to the mining industry are listed. The
classifications given are the ISO registered alloys. Changes in demand do
lead to alloys and sections being discontinued, so availability should
always be checked with aluminium suppliers.

Category and
Alloy Characteristics Common Uses
Alloy
Sheet and Plate
5251 Non-heat-treatable. Medium Side Cladding for mine
strength work hardening alloy. cages, structures
Good weldability, formability exposed to industrial
and corrosion resistance. atmospheres.

*5083 Non-heat-treatable. Good Structures in general.


weldability and corrosion Mining man cages and
resistance. High strength after ore skips.
welding. Very resistant to
industrial atmospheres.
Extrusions
6063 Heat-treatable. Medium Secondary members
strength alloy. Good weldability where complicated
and corrosion resistance. Used sections required.
for intricate profiles.

6082 Heat-treatable. Medium Structural members to


strength. Good weldability and man cages and ore
corrosion resistance. skips, although most of
the previously popular
sections are no longer
made.

*6261 Heat-treatable. Properties very Main structural


similar to 6082. Preferable as members to man
air quenchable, therefore has cages and ore skips.
less distortion problems. Not
notch-sensitive.

* Alloys recommended for use as safety critical structural members.

19
July 01

The strength tables given in the code are adapted from BS8118 tables 4.1 &
4.2, to represent the characteristics of alloys and forms available in South
Africa. These two BS8118 tables are included here as tables 3 and 4.

Table 3: BS8118: Part 1:1991 – Table 4.1: Limiting Stresses,


Heat-Treatable Alloys
Thickness Limiting Stress (yield)
Up to and Po Pa Pv
Alloy Condition Product Over
including

mm mm N mm2 N mm2 N mm2


6061 T6 Extrusion – 150 240 260 145
6063 T4 Extrusion – 150 065 085 040
T8 Drawn tube – 12 180 190 108
T6 Extrusion – 150 160 172 096
6082 T4 Extrusion – 150 120 144 72
T4 Sheet 0,2 3 115 145 70
T4 Plate 3 012 105 140 65
T6 Extrusion – 020 255 275 155
20 150 270 290 160
T8 Drawn tube – 6 255 306 153
6 12 240 282 144
7020 T6 Extrusion – 25 270 290 162

Table 4: BS8118: Part 1:1991 - Table 4.2: Limiting Stresses,


Non-Heat-Treatable Alloys
Thickness Limiting Stress (yield)

Up to and Po Pa Pv
Alloy Condition Product Over
including
2 2 2
mm mm N/mm N/mm N/mm
1200 H14 Sheet 0,2 12,5 90 90 90
3103 H14 Sheet 0,2 12,5 110 120 065
H18 Sheet 0,2 3 150 150 090
3105 H14 Sheet 0,2 3 145 150 085
H16 Sheet 0,2 3 170 175 100
H18 Sheet 0,2 3 190 200 115
5251 H22 Sheet, plate 0,2 6 130 165 78
H24 Sheet, plate 0,2 6 175 200 105
5454 0 Sheet 0,2 6 80 147 48
H22 Sheet 0,2 3 180 215 110
H24 Sheet 0,2 3 200 235 120

20
July 01

Availability
Aluminium extruded sections, sheets and plates are available from semi-
fabricators or from "engineering" (as distinct from "architectural")
aluminium stockists. Material is available in two grades, commercial and
technical. The difference is in the degree of quality assurance and
documentation applied to the finished product. Technical grade, as
described in the code, should always be used for safety critical
applications.
Technical Grade consists of EMPS and SMPS, which are Hulett
Aluminium internal specifications monitored in terms of SABS 0157 to
achieve a consistent quality material suited to use on the mines. Technical
grade is thus a material which is subjected to a higher level of quality
control. The code recommends that this material should be used for safety
critical members. Where this material is specified, it should be roller
marked using etching paints, for clear identification.
A wide range of extrusions are regarded as standard in that they represent
general purpose shapes such as angles, channels, I beams, and square,
rectangular and circular tubes and solid sections. Amongst these are the
more commonly used sections that can be purchased at relatively short
notice. General purpose mining material in technical grade 6261 T6 is
available off the shelf in mining areas. Other standards may require a few
weeks notice.
Most of the some 25 000 extrusion dies available (and removed if not used
for eighteen months) are the non-standard special dies. These are either
special in the sense that they are not general purpose shapes, or because
they are privately commissioned.
Extrusions for mining purposes are normally available in T6 (Solution Heat
Treated) condition, as welding is not envisaged on safety critical sections.
Where welding is envisaged, e.g., on sections which are not safety critical,
F material would be a common choice as the heat affected zone normally
shows similar tensile capacity after welding.
As extrusion dies are cheap, costing about same as quarter a ton of semi
fabricated metal, many designers elect to design a shape to suit their own
purposes, so saving either metal or labour or both.
Most plates are supplied in H4 condition. These can be used for built up
beams bolted together.
Alloy 5083 is supplied in temper H2 modified above 6,5 mm thicknesses
because of local processing limitations. The relevant strengths at differing
thicknesses are given in table 5. Alloy 5083 plate is no longer rolled in
South Africa.

The H2 temper is obtained by forming plates thicker than the final thickness,
and then pressing them down to the final thickness, work hardening them in

21
July 01

the process. Presses used in South Africa are not capable of developing this
hardening through plates thicker than 6,5 mm, so the inner portion of thick
plates is not tempered to H2.

Table 5: Limiting Stresses for 5083 H2 (Modified)


Thickness Limiting Stress (yield)

Up to and Po Pa Pv
Alloy Condition Product Over
including
2 2 2
mm Mm N/mm N/mm N/mm
5083 H2 modified Plate 6,6 9,5 200 229 120
H2 modified Plate 9,5 12,7 175 202 105

Notch ductility is high in aluminium. The fracture toughness, or impact


resistance is thus sufficiently high for all structural applications. It is thus
not specified as a particular material requirement. Common general
purpose extrusions suited to mining and included in stock are listed in
Appendix II. Commonly used plate sizes and thicknesses held as stock are
also included in Appendix II.

Description of alloy and temper

One problem is the understanding of the alloy and temper systems. The
first number of the four digit series indicates the major alloying element(s).
The last two indicate the alloy number. The second shows variants to the
basic alloy. Thus 6061, 6161, 6261, and 6361, for instance, may be
expected to show similar mechanical properties. Other properties may be
slightly different and have been developed for specific applications. 6061
is the alloy usually used in South Africa because it behaves well in fatigue.
If it is not available, 6082 can be used, but is not available in technical
grade.

The alloys are divided into two classes, namely, heat treatable, designated by
T and non-heat treatable, designated by H. The T is normally followed by one
(UK), or two or more (US), digits. The second number (US) corresponds to the
UK figure and indicates the strength. The first UK figure shows the process
route through which the strength was obtained, the third and others indicating
modifications to the basic process. Annealed materials, whether extrusions or
plate, are designated O, and as fabricated materials are designated F.
Thus it is important to indicate not only the alloy but the temper as well.
This is because there are a number of ways in which aluminium alloys
may be strengthened. The first is before casting, by alloying, ie adding
small quantities of other elements such as Mn, Mg and Si into the basic
aluminium. The second is after casting and may be by cold working, by
solution heat treatment or by both. While the 6xxx series can be heat
treated, as indicated by the temper designator, the 5xxx series can only be

22
July 01

cold worked to further enhance its inherent strength, as indicated by the H


temper designator.
In general technical grade 6061 alloy in T6 condition would be used for
extrusions intended for mining use.
5083 plate is normally called up in H2 (modified) above 6,5 mm thick. The
strengths are slightly lower than the normal H2 due to local process limitations.
Sheets in other alloys are normally supplied in H4, half hard, condition.

New materials
Two new types of material are available but should only be used after
laboratory testing as the long term track record has not been established.
The first of these are MMC's, metal matrix composites. These give both
higher strengths and higher stiffness in the normal range of general
purpose extrusions.
The second is laminar composites of aluminium skins adhesively bonded
to an aluminium honeycomb centre. These are the main structural
component of aircraft fuselages. Lighter, stiffer and stronger than normal
plates of the same mass, abrasion and impact resistance can be obtained
through choosing thicker skins. South Africa has the capability to work with
skins up to 12 mm thick.

Limitations of scale
While the notes have indicated that extrusions may be developed in a wide
range of shapes there are size limitations to what can be achieved. Generally,
an extruded flat profile may not exceed 350 mm wide while a square or
rectangular section must fit within a circle of about 200 - 250 mm diameter. The
length of extrusions is limited by a mass of about 75% of the charge billet
mass. This is normally about 20 kg (equivalent to 60 kg in steel) although it can
be higher, depending on the extrusion press used. Consequently very long
lengths of heavy sections cannot be produced, while light sections can be
much longer.
Concerning plate, the general width limitation is between 1 000 and 1 750
depending on the alloy and thickness chosen. Lengths of 2 500, 3 000 and
4 000 are the most common. The maximum thickness available is 50 mm,
the minimum 1,6 mm. A full listing of all sizes and thicknesses commonly
held is included in Appendix II. The normal limitation on stretched (tension
levelled) material to 6 000 mm. Stretching plates is necessary to distribute
stresses evenly so that a plate, once cut, remains plain.
Stretched bridle (back) plates in 6082 T6 are supplied in 465 mm widths,
10 mm thick in 8,7 m and 10,7 m lengths.
Available lengths of particular sections should always be checked with
suppliers.

23
July 01

Connections and corrosion


While aluminium is corrosion resistant it is also subject to galvanic corrosion
because of its position on the galvanic scale. This is significant because
connections are normally bolted using a material different to aluminium. This is
because aluminium connectors tend to creep under load. The use of 316
stainless steel (self passivating) bolts overcomes this problem - but is
expensive. Where stainless steel cannot be used good practice would indicate
that connecting parts should be isolated through the use of PVC tape or a
bitumen paint.
However, hot dipped galvanised huck bolts are commonly used without
protection because the condition of conveyances is monitored regularly.
Where different materials are used for the bridle and body of the cage, it is
essential to isolate the two materials from one another using suitable
plastic tape or a bitumastic coating.
Basic Design Considerations
Because aluminium is more prone to fatigue than is steel, it is desirable to
situate joints and connections at positions of lower stresses or stress
changes. This can be achieved through the flexibility of extrusions.
Another consideration could be that of designing fatigue prone sections as
replacement sections specifically designed for easy removal after a
specific number of hours, or once a crack of given length has developed.
Again this can be achieved through choice of extrusion design with the
possible choice of a special shape to achieve the purpose.
Aluminium has an elastic modulus of 70 GPa, ie one third that of steel.
Consequently deflection and lateral buckling need to be considered more
carefully. While an unlikely consideration to affect man cages, local
deformation or buckling under point loads is also more common because of the
use of thin sections. This may be important in skip bodies.
Huck bolting is the recommended method of connecting parts together.
Welding is not normally allowed for man cages, or safety critical
components of any conveyances.

Fatigue Stress Endurance Curves

Different types of welded and bolted joints have been classified into
groups and rated, for fatigue purposes, from best to worst. The
classification varies from Class A (best) to Class W (worst).
In general fatigue failures are caused by repeated loading and unloading
of a component with tensile stresses. The fatigue life (the ability of the
component to withstand the tensile stress) depends on two elements:
• magnitude of the stress range and
• the frequency of repetition of the stress range

24
July 01

Figure 7 is a graphical representation of the relationship between the


stress magnitude and the frequency of application of the stress for steel
components. The graph shows that a component can withstand far fewer
repetitions of the higher stress ranges than the lower stress ranges. It also
shows that Class B components will survive longer than Class C, etc.
Similar curves could be drawn for aluminium structures but, in general,
aluminium components have approximately one third of the fatigue
endurance of steel components. Table 6 shows values extracted from the
graph for a single stress range of 100 MPa, which is a realistic stress
range for the welded stiffeners on a skip, for example. The fatigue class is
an approximate correlation of steel and aluminium fatigue reference
classes.

Note that Class A components were not included in the table because to
achieve a Class A detail, the material would need to be machined and
polished which is never practical for skips and cages.

Table 6 - Endurance Cycles for Steel & Aluminium components


Under a Repeated Stress Range of 100 MPa

Load Repetitions Before Failure Occurs

Fatigue Class Steel Components Aluminium Components

Steel Aluminium

B 60 No limit 120 000

C 50 1 700 000 93 000

D 42 750 000 63 000

E 35 360 000 48 000

F 29 140 000 16 000

W 14 130 000 5 200

Special Materials for Specific Components

Many specific components of conveyances, for example the rope


attachments, have special requirements relating to impact resistance,
strain ageing, etc. The minerals act and each of the mining houses have
their own specifications for these items, and several research papers (for
example Van Rooyen, 1972, and Taljaard, 1973) have dealt with the
special requirements. This code is only intended to cover design of the
structural members of conveyances, so no reference is made to any of
these special materials.

25
July 01

1000

Stress range, MPa

100 b
y b-1
w c
d
e

10
4 5 6 7 8
10 10 10 10 10
(a) Carbon Steel Cycle Life

Note: static requirements may Class (= reference strength in


2 6
limit fr in this region N/mm or 2 x 10 cycles)
3 4 5 6 7 8 9
10 10 10 10 10 10 10
2 3 5 7 2 3 5 7 2 3 5 7 2 3 5 7 2 3 5 7 2 3 5 7
300 300
*************

200 200
160 160
120 120
(log scale)

100 100
80 80
60
60 50 60
Stress range fr , N/mm2

50 42 50
40 35 40
29
30 24 30
20 20 20
16 17 16
14
12 12
10 10
8 8
6 6
5 5
2 3 5 7 2 3 5 7 2 3 5 7 2 3 5 7 2 3 5 7 2 3 5 7
3 4 5 6 7 8 9
10 10 10 10 10 10 10
(b) Aluminium Endurance N (cycles) (log scale)

Figure 7: Fatigue Stress Endurance Curves

26
July 01

5. NOMINAL LOADS

The definition of loads has been separated into three different categories. The
first is section 5.1, General Operating Loads, which are loads applicable to all
conveyances in either vertical or inclined shafts. The second is sections 5.2 to
5.4, which define the specific loads due to man hoisting, material and
equipment hoisting and rock hoisting respectively. The third is section 5.5,
Emergency Loads, which deals with emergency loads applicable to all
conveyances.
Several of the loads defined in this section are defined in terms of an
impact factor multiplied by another load. It must be realised that such
loads are still to be regarded as "nominal loads", because they are the
best empirical, statistical or rational estimate of the load, which is
available. Such loads must still have load factors applied, before they
become "ultimate loads" for design purposes.

5.2 General Operating Loads

The loads specified in this section are those regarded as important for the
design of conveyances. Some additional loads do exist, but are not
regarded as significant for design of conveyances. Loads neglected
include the following:

a. Air drag load


Ventilation air in shafts may flow at a velocity of up to 12 m s, and
conveyances travel at hoisting velocities of up to 20 m s. The relative
velocity may thus reach 32 m s, causing some air drag load. This is,
however, a small load when compared to other loads. Only in isolated cases
have air flow loads been known to be severe. One case was where two
large, high speed cages filled most of the shaft cross section. It was found
that buffeting as the cages passed each other led to high lateral
accelerations, and bad wear on the guides.
This does not occur where there is a single conveyance, or where there
are more than two conveyances in a shaft.
If it is felt to be necessary to include air drag, it may be calculated using
the formula:

W = Cp ρ ( V - Va )2 A

where: Cp = drag coefficient, taken as 1,5


ρ = air density, taken as 1,0 kg m3
Va = ventilation air velocity
V = conveyance hoisting velocity
A = plan area of the conveyance

27
July 01

A typical value of this force is thus:


W = 1,5 x 1,0 x (10 + 15)2 x 4 N
= 3 750 N

5.2.1 Permanent Loads

5.2.1.1 General

In general, dead loads are calculated in accordance with the loading code,
SABS 0160.

5.2.1.2 Conveyance Self Weight(Gc)

Note that Gc is the actual weight of the conveyance, excluding the rope
attachments. The self weight is thus:

Gc = 9,8 mc

where mc is the conveyance self mass

Many different considerations influence the self mass of conveyances, so


the initial determination of dead weight is discussed below. It should also
be noted that conveyances are always weighed after fabrication is
complete, so any design calculations can be checked against an exact
dead load.

(a) The mass of a counterweight is normally designed to be equal to the


average mass of the conveyance it counterbalances, loaded and
unloaded. Its self mass is thus:

Self mass, mc = Cage self mass + Payload 2

where the payload is the maximum mass conveyed.

(b) A survey of skips and cages in use on South African gold mines has
shown that their self mass is influenced most significantly by the
following factors:

• material of construction, aluminium conveyances being lighter


than steel conveyances, and skip liners contributing
significantly.

• type of conveyance construction, particularly whether the bridle


hangers are integral with the skip or cage body, or separate as
with a Job's bridle or swingbody skip.

28
July 01

• type of winder, because friction winders lead to conveyances


with heavier safety critical members than other winders,
because of the high tail rope loads which must be carried.

Table 7 gives guidance on typical self masses for the major components of
conveyances. All values in part (a) of this table give the self mass as a
ratio of maximum payload, and the values for specific equipment in part (b)
are masses in kg.

(Note that the values in this table are given for rough guidance only, and
may vary widely depending on design details and specific equipment used.
Most of the values in the table have a standard deviation of about 20% of
the tabulated average value. More accurate values should be obtained
wherever possible. In this table, a shallow shaft is less than 1 600 m deep,
and a deep shaft is up to 2 500 m deep.)

(c) Equipping conveyances vary so widely in design and function that no


typical self masses can be listed.

(d) Kibbles typically have a self mass which is of the order of the
payload.

Some concern has been expressed about the possibility of the self mass
of conveyances becoming too low to accelerate the conveyance down the
shaft, where a high headgear and the typical ground mounted winder
arrangement are used.

The rope catenary between the winder and the sheave develops a
significant tension in the rope which may lead to rope being paid out
onto the bank when the winder accelerates, if the empty conveyance is
not sufficiently heavy to overcome this tension and accelerate down
the shaft.

29
July 01

Table 7: Approximate Self Masses of Conveyance Components

(a) Main Structural Components (ratio of payload)


Steel Aluminium

Mancage body 0,35 0,20

Equipment cage body 0,60 –

Skip body 0,25 0,14

Friction winder
Integral bridle
Deep shaft 0,15 –
Shallow shaft 0,13 –
Separate Bridle
Deep shaft 0,17 –
Shallow shaft 0,15 –
Other winders
Deep shaft 0,12 0,06
Shallow shaft 0,11 0,05
Rubber liners 0,04 0,04
Steel liners 0,06 0,06
(b) Fitted Equipment (mass in kg)
Guide rollers (3-roller cluster) 100
Jack catch lugs 30
Slippers (4-sets of 1 face & 2 side slippers) 120
Cage doors (roller shutter) 140
Pivot bar 80
Pneumatic cylinder for air skip 400
Door rollers, links for swingbody skip 180
Tipping rollers, etc 160

(c) Rope Attachments (mass in kg)

Compensating sheave 1 000


Drawbar, pins and shackles 300
Detaching hook 150
Cappels and thimbles 100

30
July 01

The winder, rope, sheave, and conveyance can be modelled as shown in


figure 8.

L
(Horizontal sheave – winder distance)

Sheave

H
(Vertical distance of sheave

T2 T3
above winder)

Conveyance
D

Mass

T1

Figure 8: Winder and Sheave Layout

A linear relationship between rope mass and rope strength is assumed, T2


is assumed to be 1% less than T3 due to friction, and the weight of rope
between conveyance and sheave is ignored. If it is then assumed that the
limiting conditions during winder acceleration are either a 10º sag in the
rope angle at the winder, or a tension at the winder, T1, below 1% of rope
strength, the minimum conveyance mass may be calculated.

Figure 9 gives the minimum mass as equivalent to the mass of a length of


the rope, for varying winder to sheave distances and three different winder
acceleration values.

The conveyance self mass must never be less than this value, perhaps
multiplied by a safety margin of about 1,5 as considered by the engineer to
be appropriate. It will be seen, however, that it is unlikely that any
conveyance will have a self mass as low as this minimum value, except in
a deep shaft where the rope has a high capacity factor.

31
July 01

Sheave height above winder, D (m)

200

180 x x x x

160
o o o

140
x v x v x vx v

120

o o o
100 v v v v
x x x x

80

60 v v v v
o o o
x x x x
40

20

20 40 60 80 100 120 140 160


Horizontal distance between winder and sheave, L (m)

o – acceleration 0,4 m/s2


x – acceleration 0,8 m/s2
v –_ acceleration 1,6 m/s2
* indicates equivalent
length of rope

Figure 9: Minimum Conveyance Mass (equivalent to m of rope)

EXAMPLE
A particular installation has a sheave height of 90 m, and the sheave is
70 m in front of the winder. A 58 mm diameter rope is used, which has a

32
July 01

mass of 14,68 kg/m and a breaking strength of 2 500 x 103 N. Winder


acceleration is 0,8 m/s2
From figure 9, the minimum mass of the conveyance is:
mc min = 300 x 14,68 = 4 404 kg

Allowing a margin of 1,5 implies a minimum required mass of:


mc min = 1,5 x 4 404 = 6 606 kg

Assuming a deep shaft, where the capacity factor on the rope is 13, the
maximum full conveyance weight is:
2 500 x 103
Gc + Conveyed load = = 192 x 103 N
13
A payload ratio of 0,6 gives a self mass of:
192 x 103 0,6
mc = = 7339 kg
9,81 1,6

This exceeds the minimum required mass, so it is acceptable.

5.2.2 Vertical Imposed Loads due to Holding Devices


The "conveyed load" defined in this clause refers to the weight of whatever
is being conveyed by the conveyance, – i.e. equipment, material,
personnel or rock. The weight of underslung items is not included,
because conveyance holding devices normally cannot be used during
underslinging operations.
In assessing the load effects of the loads due to holding devices, it is
important to assess where, and how, the conveyance is supported at each
step of the operation.
This may be illustrated with reference to a conveyance held by guide clamps,
which are placed at the bottom transom in order to protect them from spillage.
If material or personnel are loaded on the upper decks, the effect will be to
reduce the dead load tension, or even to induce compression, in the bridle
members below that deck level. Once the clamps are released the payload is
carried by the head rope. The compression is thus released, and the load is
carried as additional tension in the upper portion of the bridles and additional
bending of the top transom. This process has specific stress and fatigue
implications for the members concerned.
The elastic stretch of the winding rope when loading, or unloading,
conveyances, is very approximately 1 1000th of the depth of the loading
station.
The actual values of this rope stretch depend on many things, such as tensile
strength of the rope and its construction, the factor of safety and capacity factor
used. While the design is governed by the capacity factor, the payload remains
constant, and rope stretch is linearly proportional to shaft depth. However, once
the factor of safety governs the design, rope stretch reduces, because the

33
July 01

payload which can be carried reduces. Figure 10 shows calculated values of


rope stretch, for a UHT triangular lay rope, 6 x 32 construction, having a
diameter of 49 mm and a tensile strength of 2 000 MPa. This figure also shows
the maximum payload possible at the different shaft depths. This elastic stretch
of the rope can lead to two problems. The first is that loading and unloading of
the conveyance becomes difficult, or impossible. If a skip moves down whilst it
is being loaded extra spillage may occur. If a cage moves up or down whilst
unloading or loading material, the material cars may jam, or high impact forces
may result. Access for men may become difficult if the movement is too great.
The second problem is increased dynamic loading on the rope.

4,0
Rope Stretch (m)

160

Maximun Payload (kN)


Maximum payload
3,0 120

2,0 80
Rope Stretch

1,0 40

0,0
0 1000 2000 3000 4000 5000
Shaft depth (m)

Figure 10: Rope Stretch for 49 mm diameter UHT Rope


Various methods of compensating for this rope stretch have been
introduced, and other methods are under consideration.

5.2.2.1 Kep Engagement Load


It was fairly common practice to use keps to hold cages in position whilst
loading and unloading at stations where the depth exceeds about 1200 m.
These are mechanical devices located at the stations. It appears to be
common practice to insert keps whilst the cage is travelling at a creep speed of
about 0,2 m s. This ensures that the cage stops at the correct level, but an
impact force, as defined in this clause, is also applied to the cage. This clause
recognises that both the dead load and payload of the cage are travelling with
a small velocity, when the cage is suddenly stopped.
A survey of kepping practice showed that in extreme cases, cages
travelling at up to 3 m s may be arrested by keps. This should be
regarded as extremely bad practice and is likely to damage the cage and
the station steelwork. The forces specified in this clause do not cater for
the forces generated in such a case.
Ideally, a spring, or other flexible device, should be located between the
hook and the hook support to limit the impact effects. Figures 11, 12 and

34
July 01

13 show the arrangement of a kepping system and the position of coil


springs and or hydraulic buffers used to absorb part of the kinetic energy.
The magnitude of this load depends on the total load which is stopped by
the kep, and the stiffness of the kep. The total load moving is the self
weight of the conveyance, the conveyed load and the tailrope weight
where a tailrope is used. The stiffness of keps which do not have any
spring is difficult to determine, so a factor of 1,5 is specified to cover this
case. Where a spring is used, the load can be rationally assessed from the
energy equation for stopping the conveyance, added to the static load, ie:
K = V m v k + (G c + Conveyed load + T )
where: V = travelling velocity, which may be taken as 0,5 m s
m v = the moving mass, ie (G c + Conveyed load + T ) 9,81
k = the kep spring stiffness
Once the conveyance has stopped and the kep is engaged, the tension in
the head rope is released by creeping the winder. This prevents the cage
bouncing up the shaft as it is unloaded. Once loading or unloading of the
cage is complete, creeping of the winder again takes the cage weight onto
the headrope and releases the kep. This load transfer takes place slowly
as the rope stretches, so there is no impact induced.
Kepping is seldom used in new shafts.

5.2.2.2 Guide Clamp loads


Another method of holding cages in position is guide clamps, which are
usually located on the cage, and clamp onto the guides. Most of these
devices cannot be operated until the winder is at rest, so there is no clamp
engagement load defined. There have however, been suggestions that
clamps might be used for emergency braking.
Where this is implemented, consideration should be given to clamp
engagement loads, which are likely to be similar in magnitude to kep
engagement loads. Clamps, however, tend to be released very quickly,
leading to dynamic forces approaching a suddenly applied load on the cage,
hence the clause requirement of an impact factor of 2 on the imposed load
only. Recent developments in these clamping devices have led to the
introduction of slow release clamps which reduce the dynamic overload to
about 20% to 30%. A reduced impact factor of 1,3 could be used in this case.
The self weight of the conveyance and any tailrope load continue to be
carried by the headrope whilst clamps are applied. In unusual cases the
head rope tension may be released, in which case the clamp loads should
be increased by the total of conveyance self weight and any tailrope load.

35
July 01

Buffer box,
see Fig. 13

Fig 11: View on centre Kingpost Fig 12: View on A

36
July 01

Figure 13: Spring on Kep

5.2.2.3 Loads from other Holding Devices

Any other types of devices should be assessed in terms of how they


operate, and what impact loads may be induced by their operation.

5.2.3 Lateral Imposed Loads

In order to guide any conveyance in a shaft, ensuring a motion which is


generally parallel to the shaft axis, certain loads which are nominally
perpendicular to the direction of travel, must be applied to it. The
magnitude of these loads is dependent on the guide system used and

37
July 01

various characteristics of the conveyance and the winder system. The


loads are defined for three different cases, i.e. fixed guide systems, rope
guide systems, and inclined shafts. These loads are applied to the
conveyance through slipper plates, guide shoes, rubbing blocks or wheels
in the case of inclined shafts.

Typical slipper plate and rubbing block details are shown in figure 14.

Application of Lateral Loads

In the two cases of fixed guides and incline shafts, the lateral loads
defined are to be applied at one point only. Previously, design procedures
have sometimes considered lateral loads to be distributed between, say,
two slipper plates, but in practice the random lateral motion of
conveyances makes it most unlikely that a high lateral guiding load will be
applied at more than one point at any one time.

The randomness of lateral motion however, leads to the possibility of the


lateral load being applied at any of the guide rollers or slipper plates on
fixed vertical guides, or any of the wheels in inclined shafts. The very high
flexibility of rope guides leads to equal distribution of the lateral loads for
all rubbing blocks.

In the cases of either fixed guides or rope guides in vertical shafts, it is


assumed that the lateral loads may be applied in any direction. Usually two
directions would be considered, the direction in the plane of the guides
and that perpendicular to this plane, as shown in figure 15.

38
July 01

(b) Typical rubbing block detail


(a) Typical slipper plate detail used with rope guides
used with tophat guides

(c) Typical guide rail shoe detail used with rail guides

Figure 14: Typical slipper plate and Rubbing block details

Plane of Guides

Figure 15: Direction of Application of Lateral Loads

39
July 01

Reaction of Lateral Loads

Any analysis of forces in the members of a conveyance is only possible if it is


known where the applied lateral loads are resisted. The difficulty arising here
is that the lateral loads are dynamic, so that they are essentially resisted by
the mass inertia of the conveyance. There are thus three possible
approaches suggested, the first two of which are static approximations.

1. The simplest procedure is to assume a rigid support at the centre of


gravity of the conveyance, and hence calculate member forces.
A modification to this procedure which may be considered, includes
the inertia force of the transom. In the typical case where the slippers
are placed adjacent to the top or bottom transom, any load applied to
the slipper must first cause an acceleration of the transom before any
force is transferred into the bridles and hence into the conveyance. A
high slipper impact can typically lead to a transom acceleration in
excess of 4 m s2. For the design of the conveyance, it is thus
possible to reduce the slipper force by the force required to cause this
acceleration. This force reduction may be assessed by means of a
comprehensive dynamic analysis of the conveyance, or it may be
approximated by:
FA = C A m t
where mt = mass of transom and attached floor or roof
CA = transom acceleration
Where the slipper is not placed adjacent to the transom, there is no
mass to be accelerated and so the slipper force cannot be reduced.
2. A slightly better procedure is to assume that the lateral load is resisted in
proportion to the mass distribution of the conveyance, and any resultant
moment is resisted at the centre of gravity of the conveyance.
3. The most sophisticated procedure is a full dynamic analysis. The SDRC
report, (ref Final Report on an Investigation into Shaft Steelwork and Skip
Dynamics. Phase 1 – Experimental Programme, June 1983) shows
measured bunton slamming forces which are approximately triangular
impulses over a time of 0,1 to 0,25 seconds. Measurements carried out by
Krige on a 105 man cage with no rollers at Hartebeestfontein show
measured slipper forces which are also approximately triangular impulses,
over a time of 0,15 to 0,25 seconds. It is thus recommended that a
dynamic analysis assume a triangular impulse with a magnitude of the
specified lateral load and a duration of 0,2 seconds.

EXAMPLE

As an example of the application of lateral loads, consider a 10 000 kg


skip, with body mass 3 500 kg, top transom mass 800 kg and bottom
transom mass 400 kg. The skip is 6 m long, with the top transom 0,5 m
above the body, and the centre of gravity 3,5 m below the top transom.

40
July 01

The results of the three methods are shown in table 8. The stiffness of the
connection between the top transom and the skip body is taken as 5 x
106 N.m, and that to the bottom transom as 20 x 106 N m, for the dynamic
analysis.

5.2.3.1 Fixed Guide Systems in Vertical Shafts – Shaft Zone A

The concept of different shaft zones has been introduced to deal with the
problem of different ground strains in the vicinity of the reef intersection when
shaft pillar mining is undertaken. This is explained in greater detail in the
commentary to Part 4 of the code. Shaft zone A is remote from the reef
intersection, so only nominal rock strain is anticipated.

Table 8 : Lateral Load Application


Method 1 2 3

6000 N 1000 N 6000 N 6000 N

Load Application
and Reaction
21000 N.m 18750 N.m

O 0,2 sec
6000 N 4500 N

500 N

Shear on each
3 000 N 2 500 N 2 800 N
bridle at top

Bending Moment
on each bridle at 750 N.m 630 N.m 700 N.m
top

The loads specified in this clause are determined in accordance with the
COMRO Guidelines. Although the COMRO Guidelines refer specifically to
skips, no other rationally derived information is currently available, so it is
assumed that the same procedure is valid for all conveyances.

In summary, the COMRO guidelines define both roller loads and slipper
plate loads. The roller loads are determined by calculating the natural
frequencies of the rigid-body modes of the conveyance and assuming a
damping value. Charts, based on a dynamic analysis which utilises typical

41
July 01

shaft misalignment data, then give the dynamic roller loads, Hfd. This must
be added to the quasi-static loads Hfp due to the roller preload, and Hfg
due to reduction in the guide gauge, where appropriate. The total roller
load, Hf, is thus:

Hf = Hfd + Hfp + Hfg.

This procedure assumes as much clear space between slipper and guide
as is required, which is typically twice the guide misalignment. Common
shaft detailing does not allow so much clear space, so the code requires
the use of a roller load of either 5 000 N, which assumes a typical roller
spring stiffness of 500 000 N m and a clear space of 0,01 m, or a load as
determined by:

Hf = kr ∆c

where kr = roller spring stiffness

∆c = clearance between slipper and guide

If sufficient clearance can be provided between slipper and guides, then


the procedure given by the COMRO guidelines may be used.

The slipper plate loads are based on the conveyance moving laterally
across its compartment. The slipper at the leading corner of the
conveyance thus strikes the guide and deforms it. The guide stiffness
force then accelerates the conveyance back, so that it can move past the
next bunton. The slamming load, defined as the slipper load by the code,
is the force developed between the slipper and the guide as this action
takes place. The load magnitude depends on several variables, some of
which are straightforward, but several of which need further explanation.

The equation is taken from the COMRO Guidelines:

400 m e v 2 e
Hs = α n Pb
L2

The statistical factor αn has been introduced to differentiate between the


likely load, which is used for fatigue calculations, and the maximum
nominal load, which is used for static strength calculations.

Values of αn arising from the calibration exercise undertaken during the


development of the COMRO Guidelines are shown in Table 9.:

42
July 01

Table 9: Correlation of Measurement with Prediction


Measured Forces (kN)
Slipper
Max. guide Rollers
Shaft Compartment Conveyance plate
load
misalignment correctly set No rollers αn
up
(kN) (mm) Ave. Max. Ave. Max.
FSG 5# 7 14t Jeto skip 20,8 6 9 30 9 24 1,15
FSG 5# 8 14t Jeto skip 21,8 6 7 19 8 13 0,60
Elandsrand 12,5t Sala
Siemens 18,0 14 4 14 50 98 5,44
RV skip
Elandsrand
BMR 21t Sala skip 30,7 10 8 17 18 53 1,73
RV
FSG 5# 5 18t Sela skip 50,0 18 15 34 72 130 2,60
FSG 5# C 3 deck cage 12,6 8 3 4 5 15 1,19

The information in Table 9 is extracted from various documents produced


in the course of the research leading to publication of the COMRO
Guidelines.

The code specifies that αn = 2 in order to achieve a reasonable average of


these values.

It is important to note that Hs is still a nominal load, which must be factored


for limit states design.

The slipper forces at the leading slippers will guide the conveyance, so
that the forces whilst travelling down the shaft should be calculated for the
bottom slippers, and whilst travelling up the shaft they should be
calculated for the top slippers. Where the centre of gravity is located at the
geometric centre of the conveyance, as is typically the case for cages
these two values are the same.

In other cases they may differ substantially because the effective mass,
me, changes.

Conveyance Mass and Inertia

ms and I, the conveyance mass and mass inertia should be based on the
fully loaded mass for skips and material or equipment in cages, as the
payload moves together with the conveyance. In the case of men in cages,
the men move relative to the cage, so ms and I should be based on the
empty mass of man cages.

Where compensating sheaves or doubling down sheaves are used, their


mass must be included with ms and I if they are located immovably within
the transom. Where they are located at the top of a drawbar their mass is
not included, as they can move laterally during a slamming event.

43
July 01

As the leading corner of the conveyance moves to negotiate a bunton, the


conveyance will both move laterally and rotate. This means that only a
portion of the mass of the conveyance is accelerated, and not the full
mass. This portion of the mass, referred to as the effective mass, me, can
be shown for motion in one plane to be:

msI
me =
( I + msh2 )
where ms = conveyance mass

I = mass inertia

h = distance from slipper to conveyance centre of gravity

The above equation applies where rotation about one axis takes place,
which is true when the force direction and the conveyance centre of gravity
both lie in the plane between the guides. In all other cases, such as where
more than two guides are used, or out-of-plane slamming is considered,
rotation may take place about two axes, and the effective mass may be
adjusted to:

ms I1I2
me =
I1I2 + ms I1h12 + ms I2 h22

where I1 and I2 = mass inertia about the centroidal axes


perpendicular to the desired load direction

h1 and h2 = distance of the load along the relevant axis

The code defines this more general equation.

EXAMPLES

Example 1 A skip with two guides, where

ms = 25 000 kg

Ix = 245 000 kg.m2

Iz = 220 000 kg.m2

Iy = 20 000 kg.m2

hx = 0,9 m

hy = 5,3 m

hz = 0,0 m

44
July 01

For evaluation of the in plane force, F1


y
h1 = hy = 5,3 m z

I = Iz = 220 000 kg.m3 F1


hy,
25 000 ⋅ 220 000 F2
(220 000 + 25 000 ⋅ 5,3 )
me = = 5 964 kg x x
2
hx

z
For evaluation of the out-of-plane force, F2
y
h1 = hy = 5,3 m
h2 = hx = 0,9 m

I1 = Ix = 245 000 kg.m2

I2 = Iy 20 000 kg.m2
=
25 000.245 000.20 000
me =
(245 000.20 000 + 25 000.0,92.20 000 + 25 000.5,32.245 000) kg = 691 kg

Example 2 A man cage with four guides, where:

ms = 8 000 kg
Iz = 45 000 kg.m2
Iy = 40 000 kg.m2
hz = 2,8 m
hy = 1,8 m

For evaluation of the plane force, F1

h1 = hz = 2,8 m
h2 = hy = 1,8 m
I1 = Iz = 45 000 kg.m2
I2 = Iy = 40 000 kg.m2
8 000 ⋅ 45 000 ⋅ 40 000
me =
(45 000 ⋅ 40 000 + 8 000 ⋅ 2,8 2
⋅ 40 000 + 8 000 ⋅ 1,82 ⋅ 45 000 ) kg = 2 630 kg

45
July 01

Note: Where the mass inertia, I, is not known, an approximate value may
be estimated by use of the following relations:

(a) For a solid rectangular body:

Ix =
12
(
M 2
a + b2 )
(b) For a thin-walled rectangular box with all wall thicknesses equal:
M 2 2abc (a + b ) 
Ix =  a + b 2 + 
12  ab + ac + bc 

(c) For inertia about an axis a distance h from the centroidal axis of a
body:
Ix = I + Mh2
where M = the total mass of the body
Steelwork stiffness at centre of guide, KG (x 106 N/m)

40

35

30
48EI (x 106N/m)
25 L3
20
20

15
10

10
5
5
2
1
5 10 15 20 25 30 35 40

Steelwork stiffness at Bunton, KB (x 106 N/m)

Figure 16: Guide Stiffness

46
July 01

Guide and Bunton Stiffnesses


The guide and bunton stiffnesses should be obtained by considering the
complete guide and bunton structure. The guide stiffness is then obtained
by applying a unit point load at the centre of the guide span, and the
bunton stiffness is obtained by applying a unit point load at the bunton.
In modelling the structure, it should be noted that in practice, guide splices
and bunton joints, all act almost as fixed joints under the action of the short
duration slamming loads. Typical guide to bunton joints are pinned, unless
the guide is joined above and below the bunton, in which case the joint
becomes a fixed joint.

If a three dimensional analysis of the shaft steelwork cannot be


undertaken, an approximate value of bunton stiffness may be obtained
from a two dimensional analysis of a bunton grid, but increasing the
8 EI
bunton stiffness by 3 to allow for the influence of the guides.
L
8 EI
ie k B = kB 2 +
L3
where kB2 = bunton stiffness from two dimensional analysis
I = guide inertia
L = guide span

An approximate guide stiffness may then be obtained from the chart in


figure 16, which is based on calculating the central stiffness of a three
span section of guide as shown in figure 17.

L L L

Bunton Bunton

Figure 17: Three span guide

47
July 01

Shaft Classification

The variable e, the guide misalignment, and the number of applications of


Hs used for fatigue life calculations are based on the classification adopted
for the shaft. The COMRO guidelines recommend the use of a “poor”
classification, class C, which gives e = 20 mm.

It would be unwise, except under very closely monitored conditions to


assume a shaft classification better than “average”, class B, which gives
e = 10 mm.

It is possible for shafts to be classified differently at different locations


because conditions vary considerably within each shaft. For example,
class B may be used for the upper half of a shaft, where very small rock
movements are expected, whereas class C or worse may be appropriate
at greater depths where greater rock movements are expected.

Shaft classifications may be interpolated between the classes specified, or


extrapolated beyond them, to better or worse classifications. All
interpolation and extrapolation is done linearly.

Reduction of Slipper Loads

In certain cases it is possible that the conveyance itself has a significant


flexibility in the bridles, or that flexible slipper plate mounts are used. In
such a case the bunton stiffness may be reduced to an effective bunton
stiffness which makes allowance for the extra flexibility. The effective
bunton stiffness is given by:

1 1
k b' = +
kb kc

where kb = the bunton stiffness

kc = the stiffness of conveyance bridles or slipper plate


mounts

It is important that this conveyance flexibility be considered when


determining out-of-plane slipper loads, because the shaft steelwork
stiffness at the bunton tends to be very high, as the bunton is loaded
axially.

Notes:

1. The derivation of the slipper plate load assesses the total force
required to accelerate the conveyance around a bunton. It is not to
be added to the roller force. Physically, it is likely that the roller will
experience a high force concurrently with a slipper plate load
occurring, because the roller spring must be compressed if the

48
July 01

slipper is to strike the guide, but this should be assumed to be


included in the slipper load as defined in the code.

Usually slippers are located very close to the rollers, so there is no


difficulty with the location of this force. Where the slippers and rollers
are not located together, it should be assumed either that the entire
slipper load is applied at the slipper, or that the maximum roller load
is applied together with a slipper load which is reduced by the
magnitude of the roller load. The engineer should use good judgment
in evaluating these alternatives.

2. The terms "in-plane" and "out-of-plane" refer to the plane defined by


a pair of shaft guides, as illustrated in figure 18.

Out of plane force

In-plane force

Plane defined
by the guides

Figure 18: Guide plane

3. Ongoing research into this phenomenon is being conducted on


behalf of COMRO. As new results become available they should be
considered in place of the specific requirements of this code.

5.2.3.2 and 5.2.3.3 Fixed Guides in Vertical Shafts


Shaft Zones B, C and D

In shaft zone B rock strains are significant in proximity to reef intersection.


This has the effect of increasing guide misalignment, so an increased
value of e is defined.

In shaft zones C and D at reef intersection, special design requirements


are recommended.

49
July 01

5.2.3.4 Rope Guide Systems

Rope guides are not frequently encountered in the newer mines in South
Africa, but there are cases where they offer advantages and where they
are thus used. In new Australian mines and Canadian production shafts,
rope guides are frequently the preferred option. Rope guides are attached
to the headgear and are suspended in the shaft, being tensioned by a
block of concrete or steel fixed to their lower end. Various different
recommendations for the mass of this block are shown in figure 19.

The rope guides in any one compartment typically have tensions which
differ by a maximum of about 5% in order to prevent simultaneous
resonant vibration from occurring. Misalignment does not occur, so the
guiding forces are small.
Guide rope end load (kN)

300
250
200

150

100
50

0
0 500 1000 1500 2000
Shaft depth (m)
Bridon
TAS
SA

Figure 19: Rope Guide Tensioning Mass

At stations it is however necessary to use fixed guides, as the rope guides


do not locate the conveyance sufficiently positively for loading and
unloading to be safely accomplished.

The code defines three different lateral loads, to account for these different
guiding conditions, (a), (b), and (c).

(a) Whilst guided by rope guides in the shaft, the load is nominally taken
as 1% of loaded conveyance weight. This is much higher than the
forces actually present, but is considered appropriate from the point
of view of robustness of designs. This load should not be used for

50
July 01

assessing lateral movement of the conveyance in the shaft, because


far too large a value will result.

(b) and (c) Whilst entering and running in fixed guides at stations the
loads are calculated as for fixed guides. It should be noted, however,
that the hoisting velocity is very much reduced at this point. If the
conveyance is stopping at the station the velocity is unlikely to be
more than 1 m s, and if it is passing through the station the velocity
will normally not exceed 3 m s. The appropriate reduced velocity
should be used for calculating the lateral loads here.

The shaft condition is the "no roller" case, as rope guided


conveyances do not have rollers. The number of cycles of this load
should nominally be taken as the number of times the conveyance
enters fixed guides.

Observation of rope guided conveyances suggests that the upper


rubbing blocks or slipper shoes impact the fixed guides when the
conveyance is ascending, and the lower rubbing blocks or guide shoes
impact the fixed guides when the conveyance is descending.

There may be some load reversal on rebound of the conveyance, but


observation suggests that this is small, so it is probably reasonable to
neglect it.

(b) The values of L, e, and Kg given for entering the flare assume a
straight tapered flare and that the conveyance strikes the extreme end
of the flare, as shown in figure 20. Other cases may be envisaged,
with L and e as indicated in figure 20. Kg is always taken as the
steelwork stiffness at the point at which the conveyance is assumed to
strike the flare.

Where a conveyance was guided by four ropes near the corners, which is
recommended practice, it may have a centrally located channel to engage
on tophat type fixed guides. The flare is then replaced by a spearhead, as
shown in figure 20. All load calculations remain as for a flare.

The amount of flare was defined as 2e in SANS 10208:3 Edition 1 (1994)


because the COMRO (1990) Guidelines, followed in the code, assume a
curved guide profile with a maximum slope of 2e L. The flare is typically a
fixed, straight slope, which must thus also be 2e L. The flare width has
now been defined as e, which effectively doubles this load in Edition 2 of
the code. This is based on observation of the behaviour of conveyances
entering fixed guides. Frequently, the conveyance is moving or rotating
significantly, thus increasing the impact velocity. This problem is worse
where the rope guide anchor points are located further than normal from
the entry point to fixed guides. Usually this distance is of the order of 50 m.
In one known case where this distance was 140 m, extreme difficulty was
experienced in getting the conveyance to enter the fixed guides without
high magnitude slamming events, even at very slow speed. Fatigue

51
July 01

cracking of skips in the vicinity of slipper shoe brackets has been observed
on large skips.

An appropriate rational analysis here would be simulation of the behaviour


of the conveyances during hoisting.

The forces during travelling are quite small, and are probably not
appropriate for sizing the connection between the rubbing block and the
conveyance. A robust rubbing block connection will be obtained if it is
assumed that the fixed guide forces are applied to the rubbing blocks,
even where they do not run in the fixed guides.

to 2000 mm)
(Typical 750
L

e (Typical 100 to 250) e

(a) Flare Guide (b) Spear Guide

Figure 20: Typical Flare or Spear Guide Dimensions

52
July 01

5.2.3.5 Inclined Shafts

Very little information is available for the lateral loads on conveyances in


inclined shafts. These conveyances typically operate at hoisting speeds of
less than 6 m s, and lateral design loads of 5% of conveyance weight
have been known to produce satisfactory results. Higher hoisting speeds
or poor rail alignment conditions may well increase the magnitude of this
load.

5.2.4 Winder System Loads

This clause defines additional loads which are induced on conveyances by


acceleration and deceleration of the winder system.

In this clause, reference is deliberately made to acceleration and


deceleration of the winder system rather than of the winder or the
conveyance. The reason for this is that rope stretch leads to differing
accelerations and decelerations at the winder and at the conveyance.

The direction in which these loads are applied will always be parallel to the
direction in which the conveyance is travelling. The load may be applied
either upwards or downwards in a vertical shaft, or in either direction in an
inclined shaft. This depends on the direction in which the conveyance is
travelling, and whether it is accelerating or decelerating.

5.2.4.1 Acceleration Load (Ao)

The values intended for use in this clause are the nominal maximum
acceleration or deceleration of the winder including a dynamic factor of 2
which measurements have shown to be appropriate. The dynamic factor of
2 is based on a constant winder acceleration, and measurements at the
sheave in the headgear. Tests are proceeding on winder control which
slowly increases the acceleration, and then slowly reduces it back to zero.
This will reduce the dynamic factor. Measurements at the top transom also
indicate a reduction of this factor to about 1,7, due to rope stretch during
acceleration.

The value of ao is commonly about 0,8 m s2 .

5.2.4.2 Trip-Out Load (At)

Due to loss of power, or some malfunction, it is possible for the winder to


trip-out. The law covering hoisting requires that in these circumstances the
winder brakes are applied in a controlled manner, so that a specified
maximum deceleration is not exceeded. The maximum nominal
deceleration in the event of a winder trip-out, at , is commonly set at about
5 m s2.

53
July 01

Trip-out deceleration is included in this clause, and not as an emergency


load because it is known to occur on all winders. It must not cause any
acceptable force levels to be exceeded. It will be noted, however, that this
load is not included in any fatigue calculations, because it is applied very
infrequently.

5.2.4.4 Vertical friction load

The code defines lateral loads on slippers and rubbing blocks. Friction
between these components and the guides must lead to co-existent
vertical loads of µH, where µ is the friction coefficient and H is the lateral
load. This load would, however, only influence the slipper or rubbing block
connecting bolts, but for practical reasons these are much larger than
required to resist the vertical friction.

Typical friction coefficients for polished surfaces are listed in table 6.


These coefficients may vary widely depending on the actual conditions.
New slippers or guides whose surfaces are still rough, may have 2 or more
times this friction coefficient.

Table 6: Friction Coefficients for Slippers

Guide Friction coefficient


Slipper material
material Dry Lubricated

Steel Phosphor – bronze 0,09 0,07


Cast iron 0,18 0,06
Mild steel 0,12 0,08
Bronze 0,18 0,06

Where guide ends are mismatched, a slipper may strike the end of a guide,
imparting or high vertical load to both guide and conveyance. This is however
virtually eliminated by careful matching of adjacent guide lengths, or grinding
guide surfaces near joints, and by bevelling slipper plates.

The load is defined as 0,5 Hs. This does not imply a friction factor of 0,5,
as allowance must also be made for vertical impacts on mismatched guide
joints.

5.3 Man Winding Loads


Men are conveyed in cages. When men enter the cage, they are
supported primarily by the floor plates. The loading is transferred to the
floor beams and, where installed, through diagonal bracing members in
the sides of the cage to the bridle. From the bridle, the load transfers into
the top transom beam before being supported by the winding rope. There

54
July 01

are no sudden shock or impact loads developed during the loading and
unloading of men because the rate of loading and unloading is relatively
slow.
It should be noted that there are no horizontal loads specified for the sides
of man cages. When a cage is full of men, they will press against the sides
and exert some loads, but it is assumed that the side plates will deflect
and resist this load by diaphragm action. This is an acceptable assumption
for typical multi-deck cages with a deck-to-deck height of about 2,5 m and
plan dimensions not more than about 3,0 m to 3,5 m. Where larger cages
are used, the requirements of SANS 10208:4 for screens at stations
should be consulted.

5.3.1 Standing personnel loads


The load for standing personnel assumes 8 men per m2 at 75 kg each.

5.3.2 Seated Personnel loads


The load for seated personnel assumes 5 men per m2.

5.4 Material and Equipment Winding Loads

5.4.1 Floor Loads


It should be noted that this clause refers to loads induced by equipment
carried inside the cage. Some of the loads defined are vertical loads
applied to the cage floors, whilst other loads are horizontal loads applied to
buffers or posts in the cage.
The loads which may be applied to each deck of a cage must be
separately assessed. Commonly, the lower deck is designed to carry the
total payload in order to allow for winding heavy equipment, such as
locomotives. These are usually only loaded on the lower deck, and are
frequently the heaviest payload catered for in the design.

5.4.1.2 Impact Loads

In addition to conveying men, cages are used to convey, materials and


equipment in different ways. Severe impact damage can occur when
material cars are loaded and unloaded, because of the sudden application
of high payloads. Impact damage can be minimised by fitting:
– clamps or keps to control rope stretch in the vertical direction (see
comment on clause 5.2.2.1)
– leaf springs in the rear of the cage to reduce impact in the horizontal
direction (as shown in figure 21).

55
July 01

(a) Vertical Impact Loads


A closer examination of the sequence of loading the cage with
material cars is worthwhile. On entering a well docked cage, the front
wheels cause a suddenly applied load, with an impact factor of 2,0.
After the front wheels have entered the cage, the winding ropes
stretch when holding devices are not used. Thus, the level of the rails
on the cage floor is lower than the bank level when the rear wheels
then fall onto the rails, and a vertical impact occurs. If decking
alignment is not accurate, the front wheels may also cause higher
impact. The magnitude of the vertical impact depends on several
factors including the weight of the material car, the docking
misalignment, and the diameter and length of the winding rope. The
vertical impact factor can be calculated and generally tends to have
the effect of doubling or trebling the axle load of the material car,
even for nominal falls of 50 mm to 100 mm. As a consequence the
whole structural system (rail beams, their connections, edge beams,
bracing, bridles and top transom) suffers the effects of the impact
loads. The reverse occurs on unloading.
Certain assumptions are necessary, in order to be able to define
vertical impact factors for this loading procedure. The first
assumption made is that material cars have two axles, carrying loads
of M1 and M2. The second assumption regards the magnitude of
docking misalignment. Where cage holding devices are used, this is
assumed to be zero. A survey of cage operating procedures revealed
that in isolated cases a misalignment of as much as 700 mm is
tolerated although in most cases 50 mm or 100 mm is the maximum
acceptable amount. There is little uniformity, however, so for
simplicity it is assumed that the initial misalignment is equal to the
static rope stretch due to half the material car load.
Two cases of the vertical impact load are considered, in both cases
taking the impact load as:

Impact Factor x Axle Load 2 + Axle Load 1

However, the most important effect of this load is bending of the floor
beam at the cage door. The wheel impacts are applied directly to these
beams, which are known to commonly sustain substantial damage.

56
July 01

Figure 21: Spring Buffer in Cage

CASE 1

Where conveyance holding devices are used, it is assumed that accurate


level alignment is achieved, and the impact factor is thus 2,0. The axle
impact load is given by:

First axle = 2,0 x M1

Second axle = 2,0 x M2

The total vertical floor load thus becomes:

2M1 + M2

CASE 2

Where conveyance holding devices are not used, a level misalignment as


discussed above is used. Referring to the impact factor curves given in
figure 21, it is seen that a similar factor then results for all static deflection
values.

The load in this case is thus given by:

57
July 01

First axle = 2,8 x M1

Second axle = 3,4 x M2, if equal axle loads are assumed

The vertical impact load is thus specified as:

Cv = 3,5M1 where M1 is the heavier axle load, assumed to be applied


after the lighter axle

The total floor load thus becomes:

3,5M1 + M2

In both of the above cases it is assumed that the lighter axle is loaded first,
but that the resulting deflection is half the static deflection due to the
weight of the material car.

Where different level alignment is known to occur, or where the heavier axle
is loaded first, the engineer may consider it appropriate to calculate a better
specific impact load. As an example, consider a case where level
misalignment is a maximum of 50 mm, the front axle carries 2 3 of the
material car weight, and static deflection of the cage under the full material
car weight is 30 mm. Mt is the full material car weight.

First axle: Drop = 50 mm


Static deflection = ⅔ 30 = 20 mm
Impact factor = 3,4 (from figure 22)
Impact load = 3,4 x ⅔ Mt = 2,3 Mt

Second axle: Drop = 50 + 20 = 70 mm


Static deflection = ⅓ 30 = 10 mm
Impact factor = 4,8 (from figure 22)
Impact load = 4,8 x ⅓ Mt = 1,6 Mt
Total floor load = 1,6 Mt + ⅔ Mt = 2,3 Mt

Other vertical impact loads may also be applied to cage decks, although
the code does not make specific provision for them.

For example, cases are known where long lengths of timber or pipes are
dropped into a cage through an opening in the roof. The lengths exceed
what can be conventionally loaded in material cars, but loading into the
cage obviates the need for underslinging. These items may weigh 2 kN or
3 kN, and they fall perhaps 6 m, loading to an impact factor from figure 22
which may be as high as 10 or 15. Such loads should be identified by the
design engineer, as they may constitute a significant design load.

58
July 01

(b) Horizontal Impact Loads

If material cars are loaded into the cage too rapidly, the momentum may
cause the car to collide with the cage rear wall, causing further impact
damage. Leaf springs can be fitted to the cage rear wall as shown in
Figure 5. These springs serve only to reduce the damage caused by
rolling material cars; they do not prevent damage entirely. The springs
have limited capacity to absorb kinetic energy and can easily bottom out
when material cars move too quickly. Cars should be loaded as slowly as
practicable to minimise rear wall damage.

The horizontal impact loads caused by these loading procedures depend


on many different things. The stiffness of the cage, guides, rollers, and any
buffers used are important. Less well defined influences include the speed
at which material cars are loaded, how far the cage can move before
rollers or slippers contact the guides, the cage holding devices used, and
whether the first or last material car is being loaded. The wide practical
ranges of these variables make it impossible to perform any accurate
calculations, and no careful measurement work is available.

Impact Factor

100
8
6
5
4
3
2

10
8
6
5
4 0
3 X Second axle
2
First axle
1,0
2 34568 2 3 4568 2 3 4568 2 3 4568
0,1 1,0 10 100 1000
H/S

H - is the height the material car falls.


S - is the rope extension under static load.

Figure 22: Impact Factors

59
July 01

In view of these difficulties the horizontal impact load has been


empirically derived by considering the load carrying capacities of
members which are known to constitute adequate buffers in cages.

Typical buffer members are:

152 x 76 [ Mp = 11,2 kN.m

200 x 20 PLT Mp = 6,2 kN.m

200 x 25 PLT Mp = 9,7 kN.m

80 mm φ pipe Mp = 7,4 kN.m

(A yield stress of 310 MPa is assumed in calculating Mp to give a realistic


actual plastic moment value.)

The buffer members would typically be simply supported beams, spanning


about 1,5 m horizontally across a cage or 2,1 m vertically between floors.

W 0,4 1,7
M= = 0,325 W
2,1
1,7 m

W
0,4 m

W
WL W 1,5
M= = = 0,375 W
4 4
1,5 m

In order to cause damage to the buffers, as is know to happen, W should


be large enough that M = Mp

This means that, for the various typical members:

152 x 76 [ W = 29 kN

200 x 20 PLT W = 17 kN

200 x 25 PLT W = 26 kN

60
July 01

80 mm φ pipe W = 23 N

A typical maximum load is thus of the order of 25 kN, and typical material
cars have a loaded mass of 2,5 to 5 tons.

The impact factor is thus between 0,5 and 1,0.

The energy required to yield a beam with W = 25,0 kN is: (assuming


centre point load)

L 2 M2
E =2 1 ∫ dx
2 0 EI

=∫
L2 (W 2 x )2
dx
0 EI
0,11x10 − 6
= W  L  = W L =
2 3 2 3
kJ
4EI  24  96EI I

This must be equal to the kinetic energy, 1 2 Mv2, on impact.

Thus, for the various typical members, the impact velocity is:

M = 2,5 t 5,0 t

152 x 76 [ V = 0,20 m s 0,14 m s

200 x 20 PLT V = 0,57 m s 0,40 m s

200 x 25 PLT V = 0,41 m s 0,29 m s

The clause thus specifies impact loads of 0,5 M for cases without any
spring buffers, and where spring buffers are used an impact velocity of 0,5
m s is assumed. The force using spring buffers is then given by:

Spring stiffness k s
V
car mass, M 9,81

9,81 k s
= 0,5
M

This is limited to a maximum value of 0,5 M to allow for the practical


realities of cage flexibility.

5.4.2 Roof Loads

A nominal load is specified to allow for shaft inspections and to give the
roof a measure of robustness.

61
July 01

5.4.3 Underslung Load

Long material, of heavy items of equipment, are commonly suspended


below conveyances when they must be moved up or down mine shafts.

This procedure is referred to as underslinging, and hooks or lugs, are


commonly provided on the bottom transom for this purpose. A heavy
underslinging operation is shown schematically in figure 23.

The clause requires that all loads induced during this operation must be
properly assessed and included in the design procedure. For example, it
can be seen that in position 2 shown in figure 23, there are quite large
horizontal loads applied to the bottom transom. Load transfer during
underslinging is typically gradual, so impact factors in the region of 1,2 to
1,4 are probably appropriate, but this must be carefully assessed by the
engineer.

62
July 01

Position 3 1 Tension = 0
1
1 Tension has
2
2 value for
Position 1 equilibrium

Figure 23: Underslinging operation

It may be possible for a mine to obtain special permission to reduce the


rope safety factor down to as low as 3, during slinging of infrequently
moved heavy equipment. The conveyance would usually be removed for
these infrequent, heavy slinging operations, but this should be confirmed
before designing the conveyances. If the conveyance is not removed, it is
presumably acceptable to allow an equivalent reduction in the factors of
safety applicable to the conveyance design for these underslinging loads.
Special consideration must also be given to low cycle fatigue damage
which may occur, as working stresses may be rather high.

Directive C1 from the Government Mining Engineer, dated 1 February


1992 requires a factor of safety of 15 for devices attaching trailers to
conveyances. It is generally accepted that this is intended to be a factor of
safety against ultimate failure, so a load factor of 9 or 10 as implied by
designing for rope break loads tested against the normal design strength
of structural steel members provides approximately the same margin of
safety. It is thus not necessary to modify the requirements of this code, as
it already complies with this directive.

5.5 Rock Winding Loads

The term "rock" as used in this clause includes ore, reef, and waste.

5.5.1.1 Static Loads (R)

The static load is taken as the total volume of the skip below the loading lip
multiplied by the density and gravity acceleration. This exceeds the payload
in all cases, but is used because it is possible that skips be overloaded.
This volume is shown schematically in figure 24, and can be defined as
the hydraulic volume of the skip.

63
July 01

The density of rock hoisted varies depending on the type of ore, blasting
procedures, and the quantity of fines included. It may sometimes be very
wet, almost like a slurry. Cohesion may be high as well, leading to
difficulties with tipping. It is usually assumed that gold ore and waste has
an angle of repose between 32° and 35° and a wall f riction angle of about
60% of this value, but these values may vary widely. For gold mines the
density of loose rock, as loaded into the skip, varies between 1 600 kg m3
and 2 100 kg m3. For design purposes a density of 1 800 kg m3 is
typically used, but a more accurate value should be used if it is available. It
is common practice to calculate the required volume of skips by assuming
a low value of density, otherwise the skips may well be too small to contain
their design payload adequately. Conversely, a high density is assumed to
assess the skip strength, in order to ensure adequate robustness.

F – Pressure on skip front


B – Pressure on skip back
Freeboard Travelling Loading Tipping
Top
F B
F
Front

F
Back

F
B B B

F B F
B
F B

Door 20 40 60 80 20 40 60 80 20 40 60 80
Ore Pressure kPa Ore Pressure kPa Ore Pressure kPa

Measured ore pressures in 13,5t Swingbody Skip.

Skip volume
assumed by code

Figure 24: Skip Loads and Pressures

64
July 01

Typical density ranges for different ore types are given in table 9.

Table 9: Typical Broken Rock Densities (kg m3)

Ore Mined High Density Design Density Low Density

Gold 2100 1 800 1 600

Platinum 2800 2 500 2 300

5.5.1.2 Bridle and Transom loads while filling

Skips tend to be loaded quite fast relative to their fundamental natural


period of bounce, particularly in deep shafts where the fundamental period
is typically of the order of 3 to 5 seconds. This leads to a dynamic impact
factor which measurements show to be of the order of 1,3 to 1,5, so the
code specifies an impact factor of 1,5.

This factor may reduce under certain conditions. In a shallow shaft, or


where rock loading is done more slowly, the dynamic factor reduces.
Where the rock is loaded in a time equal to, or greater than the
fundamental period of bounce, the dynamic factor may be reduced to 1,2
(Greenway, 1989).

Some experimental work has been undertaken with a beam below the skip
to support it during loading. After loading, the beam is lowered, slowly
transferring the payload weight to the rope. This has also been found to
produce a dynamic factor of about 1,2. In this case an important
consideration is that the ore weight may induce compression in the lower
portion of the bridle hangers.

5.5.1.3 Gravity rock pressure

All rock pressures are defined in terms of a factor multiplied by the


"hydraulic" pressure at the maximum rock depth in the skip, which is
defined as ρgz. This is the characteristic pressure for skips, po.

Rock Pressures

Three distinct phases of rock pressure loading on the skip body occur
during the rock hoisting operation. These are the pressures during filling of
the skip, during travelling up the shaft, and during emptying of the skip.
The pressures during travelling up the shaft are ignored by the code, as
they are always less than the pressures during filling and emptying.

65
July 01

Should the pressure during travelling be required for any reason, it may be
taken as:
p = 0,75 po for the skip bottom
p = 0,3 po for all sides of the skip.

The rock pressures given in the code assume a fairly dry, granular
material, having a reasonable angle of repose. It is not common practice to
design skips for sludge conditions where higher, hydrostatic, pressures will
exist. If it is known that hoisting of very wet material or sludge will
frequently occur, the pressures specified in the code should be carefully
reviewed.

5.5.1.4 Pressures while filling


Measurements of rock pressures in skips have shown high pressures at
points where the rock impacts against the skip body, and fairly low
pressures elsewhere. There does not appear to be a significant pressure
increase with depth, as would be predicted by any of the theories of
pressures developed by contained materials.
The most severe impact is on the bottom of the skip, because impact is
almost perpendicular to the plane of the skip bottom. Less severe impact
is evident on the back of the skip, where rock impacts the plane of the skip
side obliquely. There is evidence of similar impact pressures on the
loading lip of the skip where this is angled towards the loading flask chute.
Other side surfaces experience a lower pressure.
Smith (1992) comments that skip doors designed for a hydrostatic
pressure with a factor of safety of 2 give member sizes of the proportions
typically encountered in practice. The code provision of hydrostatic
pressure with a load factor of 1,6 will give similar, but perhaps slightly
lighter, designs.
The normal side pressure factor here has been increased from 0,2 in
Edition 1 of the code, to 0,3 in Edition 2 because experience has
suggested that 0,2 led to pressures lower than those actually applied in
practice. Where skips have one dimension significantly greater than the
other dimension, the pressure may be even higher.

5.5.1.5 Pressures while emptying


Measurements show fairly low pressures while emptying the skip, on all
sides except local to surfaces where the rock is forced to change direction
as it flows out of the skip. A high pressure is thus defined for the skip
surfaces close to such points. It should be noted that all skips have a
change of flow direction at the back of the skip. On fixed body skips this joint
is effectively stiffened by the skip body plating, whereas on swingbody skips
it occurs at the door hinge point which is not stiffened by the door.
Maintenance problems have been experienced in the vicinity of the door
hinge, so attention must be given to adequate stiffening at this point.

66
July 01

5.5.1.6 Load on tipping rollers

This load applies to any mechanically opened swingbody skips.

Severe problems have been experienced with skip tipping and tipping
roller damage, where the tipping rollers, their fixings, and the tipping path
are not adequately designed. In some cases the locking device has
ceased to lock in, potentially leading to dangerous skip opening in the
shaft. In other cases incorrect positioning of monkey rails to release the
locking device, and tipping path rails, has led to pulling the tipping rollers
against the securely locked locking device. This has ripped tipping rollers
off the skip.

Few research results or measurements are known to be available, so a


conservative definition of this load is given.

The forces required to tip the skip can be calculated by considering the
tipping mechanism. For simplicity, the skip body self weight is
conservatively assumed to be the total skip self weight, Gc .

Two different cases of tipping roller force must be considered, both of


which are shown schematically in figure 25.

Case (a) is lifting the skip in order to tip it. Taking moments of static forces
about the pivot:

(Gc + R) x − 2Rt sin (θ2 − θ3 − θ1) L2 = 0

where x is the horizontal distance between the C.G. and the pivot
≈ L1 sin θ1 + L3
L1 is the distance between the pivot and the C.G.
L2 is the distance between the pivot and the tipping roller
L3 is the horizontal distance between the pivot and the skip C.G. in
the hoisting position
θ3 is the angle between the line through the pivot and the tipping
roller and the vertical, shown in Figure 25.
θ1 is the angle through which the skip rotates during tipping
θ2 is the angle of the tipping path above the horizontal
θi are angles as indicated in figure 25.

The angle θ1 should be assessed on the basis of the angle of tip when the
skip unloads its contents.

Rt =
(Gc + R) (L1 sin θ1 + L3 )
2 L 2 sin (θ2 − θ3 − θ1)

67
July 01

Typical values of θ1 , θ2 and θ3 are 8º, 50º and 12º respectively, L2 is


commonly about 1,2 L1 and L3 is typically small (say 0,05 L1), so that:

R t ≈ 0,122 (Gc + R )

Tipping Path

Tipping Path
Pivot
Pivot
θ3 + θ1
Tipping
Roller
θ2 θ3
Tipping
90 – θ2 Roller
Rt GC + R
Rt

θ4

(a) (b)

Figure 25: Tipping Mechanism

Case (b) is pulling the skip over the door support. This occurs on a near-
vertical portion of the tipping path. Taking moments of static forces about
the pivot is approximately:
(Gc + R) L 4 tan θ4 − 2Rt L2 cos θ3 = 0
where L4 is the distance between the pivot and the door support (camel
back), and θ4 is the slope angle of the door support.

Assuming that L4 is typically approximately 20% more than L2, we have:


(Gc + R ) 1,2 L2 tan θ4
Rt =
2 L 2 cos θ3

The angle θ4 is typically of the order of 10º, so that:


R t ≈ 0,108 (Gc + R )

Conservatively, both of these cases are covered by putting:


Rt = 0,25 (Gc + R )

68
July 01

However, where an unusual skip configuration is used it would be prudent


to check both possibilities using these more accurate calculations.

In these equations αt is a dynamic impact factor which can be taken as


1,5. This load is considered adequate, provided the tipping path and rollers
are well maintained.

In considering how these forces are resisted by the skip it should be noted
that the load in case (b), is essentially resisted at the door support and
thus the lower portion of the bridle hangers. The loads in case (a), are
essentially resisted by the pivot bar and thus the upper portion of the bridle
hangers.

More detailed calculations are given by Soutar (1973).

5.5.1.7 Skip return-stop load

This load applies to swingbody and overturning skips.

High impact loads are generated as swingbody and overturning skips


move back into the winding path after tipping, and hit against the return-
stops. This load is particularly high in overturning skips, and has been
known to produce fatigue cracking and failure of bridle members adjacent
to the stops. The magnitude of loads applied to the return-stops is difficult
to determine, because the loads arise from manufacturing tolerances and
wear of components. Ideally, the skip body is moved back into the winding
path, by the tipping rollers running in the tipping path, in such a way that it
comes to rest against the return stops and no load is applied. In the real
world of tolerances and wear, one of two possible actions occurs. For
simplicity, the skip body self weight is conservatively assumed to be the
skip self weight. Gc.

a. Over-return load

The position of the tipping path may be such that the tipping rollers tend
to over-return the skip by pushing it beyond its neutral position in the
winding path. The tipping rollers thus force the skip against the return
stops. The skip body is rotating about its pivot point, as indicated in
figure 26.

The force applied to the return stops in this case is thus given by:

 Distance from pivot to tipping roller 


Rs = Tipping roller force  
 Distance from pivot to return - stop 

For want of better information, it is assumed that the total tipping roller
force is equal to the weight of the skip body multiplied by an impact
factor of 0,5, ie

Tipping roller force = 0,5 x Weight of skip body on each side.

69
July 01

Tipping Path
Pivot

Tipping Path

Return Stop
Tipping
Roller Tipping Roller Centre of gravity

Pivot
Tipping
Path

Tipping Roller Tipping Path

Tipping Roller
Return Stop

Centre of gravity

Pivot

Figure 26: Skip Tipping

70
July 01

b. Under-return load

The position of the tipping path may be such that it does not move the
skip fully back into its neutral position in the winding path. This under-
return, leads to the skip falling back against the return-stops under the
action of induced motion and gravity. The magnitude of this load
depends on the velocity of motion and the stiffness of the return-stop,
making it difficult to determine. The code thus assumes a force which is
similar to that assumed for the over-return case, ie:

 Distance from pivot to C.G. of body 


Rs =0,5 x Weight of skip body  
 Distance from pivot to return - stop 

In both cases it is assumed that this load is applied in equal proportions


to the return-stops on both sides of the skip, and it is assumed that the
same load applies to both swingbody and overturning skips.

When applying this load in design, experience has shown that it is only
necessary to check its influence on the return-stops themselves, their
connection to the skip body and the bridle, and the bridle members to
which the return-stops are connected.

This load may be reduced by using a rubber pad or spring on the


return-stops, but in the case of an over-return load the effectiveness of
this proposal is doubtful.

EXAMPLE

15 000 kg swingbody skip, with a body mass of 5 500 kg.

Distance from pivot to tipping rollers = 5m

Distance from pivot to C.G. of body = 4m

Distance from pivot to return-stops = 4,5 m

 5 
Over-return load = 0,5 x 5 500 x 9,81   x 1,6 N
 4,5 
= 47 960 N

 4 
Under-return load = 0,5 x 5 500 x 9,81   x 1,6 N
 4,5 
= 38 368 N

The return-stop is connected to the bridle hanger, which is a 240 x 85 [ ,


7 m long with guide rollers and slippers at each end. The return-stop is
connected 5,5 m below the top end of the bridle hanger.

71
July 01

47 960 2 x 5, 5 x1, 5
Hanger B.M. = = 28 262 N.m
7

28 262 28 262
Hanger σm = = 10 6 MPa
Zx 300 x10 −6

= 94 MPa

5.5.2 Kibble Loads

Kibble loads are dealt with in a manner similar to skip loads, but adding
different pressure factors for sludge, as this is a common load case for
kibbles. Sludge is treated as having a hydraulic pressure distribution.

5.5.2.4 Pressures during emptying

Kibbles are tipped over sideways to empty them, as shown in figure 27.
The effect of this orientation of load on the side of the kibble must be
assessed, giving particular attention to the lack of symmetry of loads
around the sides.

Figure 27: Emptying of Kibble

5.6 Emergency Loads

Various emergency loads are defined to allow for the possibility of


accidents which are known to occur. All possible precautions are taken to
prevent accidents, but should an accident occur, these load cases are
intended to minimise loss of life and other damage.

72
July 01

5.6.1 Rope Load

5.6.1.1 Permanent operating conveyances with fixed ropes

The most severe load that can conceivably be applied to a conveyance is


the load at which the rope breaks. This may happen due to the
conveyance jamming in the shaft, or falling under gravity. It has been
argued that the rope will usually break at a lower load than its nominal
breaking load because it will kink, or end connections will fail. Although
there is some evidence to suggest that splice connections fail at 85% to
90% of the rope strength, the intention is that the connection should be as
strong as the rope. There is too little information to justify any assumption
of a load lower than the nominal rope breaking strength. Resin or white
metal connections, which are now becoming more common, are usually
capable of resisting the full rope break load.

There are records of actual instances of rope break events on fixed rope
winders, so this emergency load must be considered, as it is known to
occur. It must be pointed out, however, that no instance of rope break has
occurred for many years.

The rope break load may be taken as the actual breaking strength of the
rope, if this is known from tests of the rope. Statutory requirements are
that all ropes must be regularly tested, so actual strengths are likely to be
available for conveyance replacements. Ropes must be rejected when the
test strength reduces to less than 90% of the initial specified strength. This
loss of strength is due to many causes, including wearing of the outer
strands, breaking of strands, and corrosion.

The code, however, recognises that a tested rope breaking strength may
not be available, and the rope break load is thus taken as 10% more than
the manufacturers specified strength. There are two reasons for this.

1. The manufacturers ensure that ropes are at least as strong as is


specified by slightly exceeding this specified strength. Small
variability in actual rope strengths will thus not lead to any under-
strength ropes. Rope manufactures are currently investigating
upgrading all their ropes so that this margin is only 5%, so 10% may
be considered a conservative value.

2. Many ropes show a small gain in strength after initial use, before the
strength later starts deteriorating.

Typical rope strength variations with rope life are shown in figure 28.

73
July 01

Rope Strength

110%
Nominal
90% X
Discard

Rope Life

Figure 28: Rope Strength Variations over Rope Life

The rope break load, specified as 1,1 times the manufacturer’s estimated
strength, is the emergency load.

It may appear to be inconsistent to define the emergency load as the rope


break load, rather than the total conveyance and payload weight multiplied
by a large load factor. This is done, however, because the mines Health
and Safety Act specifies a wide range of safety factors for ropes, and
these may be modified from time to time. Regulations 16.33 to 16.38 of the
Minerals Act, 50 of 1991, currently include the factors of safety shown in
Table 10, for ropes where periodic testing is done.

The safety factor is the ratio of the rope strength to the sum of conveyance
weight, payload, and rope weight.

The capacity factor is the ratio of the rope strength to the sum of
conveyance weight and payload.

The conveyance weight includes the weight of all head rope attachments,
and half the weight of any tail rope attachments.

74
July 01

Table 10: Factors of Safety

Conveyance Safety Factor Capacity Factor

Man/material cage
5 10
without tail rope

Skip without tail rope 4,5 9

Man/material cage on 2
8,1 – 0,00135L
or 3 ropes with tail
≤ 5,94
ropes
Skip on 2 or 3 ropes 7,29 – 0,0012L
with tail ropes ≤ 5,34
Man material cage on 4
8,1 – 0,00135L
or more ropes with tail
≤ 5,62
ropes
Skip on 4 or more ropes 7,29 – 0,0012L
with tail ropes ≤ 5,06

It is logical to link the safety factor to the rope length because:

1- rope length has important effects in reducing shock which are likely
with longer ropes.

2- higher levels of control, which are likely with longer ropes, reduce the
likelihood of an emergency occuring.

In Regulation 16.34.2, the safety factor is defined as:

25000
SF =
4000 + L
where:

L = the effective length of the winding rope.

This definition of safety factor requires monitoring of the rope in


accordance with SANS 10294.

5.6.1.2 Permanent operating conveyances with friction winders

The situation with friction winders is very different. In general, it is not


possible to break the ropes, because they will slip on the winder drum,
although there have been rare instances when the ropes have broken.

75
July 01

The rope slip load is the specified emergency load. The magnitude of the
slip load depends on the rope tensions in the system, and what
emergency condition has caused the rope to slip. Three possible
emergency conditions are shown in figure 29.

There are computer packages available which can simulate the dynamic
behaviour of these emergency conditions (ATD in Johannesburg are able
to do this simulation), to predict the maximum rope tensions. Alternatively,
but less accurately, energy calculations may be performed. As an
example, consider the rope snagging event shown in figure 30.

T1 T2 T1 T2 T1 T2

Cage strikes crash beams

Tail rope snags Cage sticks then falls

Figure 29: Possible emergency conditions on a friction winder

It is extremely conservative, and therefore unnecessary, to assume that


the tailrope can snag just below the cage, but this could be argued to be a
worst case scenario. A more rational assumption is that snagging below
the conveyance allows at least a small amount of movement of the tail
ropes, so that the tail ropes do not break.

76
July 01

Example in Figure 30:

The total kinetic energy in the system is:

25 000 ⋅ 15 2
Cage : = 2,81 x 106 J
2
17 500 ⋅ 152
Counterweight : = 1,97 x 106 J
2
810 x 4 x 3,0 x 15 2
Head ropes : = 1,09 x 106 J
2
810 x 3 x 4,0 x 152
Tail ropes : = 1,09 x 106 J
2
25 000 2 x 10 2
Drum : = 1,25 x 106 J
2
8,21 x 106 J

T1 T2
60 m
750 m

Tail ropes snag


at shaft bottom

Figure 30: Example of rope snagging condition

77
July 01

All the kinetic energy must be absorbed by the ropes acting as elastic
springs, and by friction losses if the winder drum slips. The elastic energy
in a rope is given by:
1 T2 L
Elastic Energy =
2 EA

The elastic energy absorbed by the head ropes is this: (conservatively


ignoring tail rope energy)
1 T12 x 810 1 T22 x 60
EE = + J
2 110 x 10 9 x 4 x 335 x 10 − 6 2 110 x 10 9 x 4 x 335 x 10 − 6

= (
3,39 x 10 −9 810 T1 + 60 T2
2 2
) J

The ratio between the rope tensions on the two sides of the winder drum is
given by:
T1
= eµα
T2

where µ = the coefficient of friction between the ropes and the drum
inserts

α = the angle of wrap of the ropes on the drum.

In order to establish the maximum emergency loads, a high value of the


coefficient of friction should always be used. The coefficient of friction
varies widely with different insert materials, so the value used should
always be checked with the winder manufacturer. It will be assumed here
that a high value of the coefficient of friction is 0,35.
T1
Thus: = e0,35 x π x 1,1 =3,35
T2

The total elastic energy absorbed is thus:


 T2 
EE = 3,39 x 10−9  810 T12 + 60 1  = 2,76 x 10− 6 T12
 3,352 

Equating kinetic energy to elastic energy gives:

8,21 x 106 = 2,76 x 10-6 T12

T12 = 1,72 x 106 N

This represents 86% of the rope break load. This calculation suggests a
scenario where the rope will not slip, but where the maximum rope load is
nevertheless less than the rope break load.

78
July 01

5.6.2 Emergency drop back

Emergency drop back following detachment of the hoist rope is onto jack
catches, which are mounted in the headgear. The resulting impact
constitutes part of the loads applied to headgears. The loads are thus
defined in SABS-0208:Part 1 "Headgear Structures".
The loads are induced by the conveyance falling under gravity after the
rope has been detached in an overwind. The magnitude of the loads thus
depends on the distance which the conveyance falls and the stiffness of
the jack catches.
Two cases are recognised:

Case 1: Jack catches rigidly mounted in the headgear. The very high
stiffness in this case makes calculation of the forces very difficult, so the
total force is often nominally taken as being equal to the rope break force.
Because the structure is very stiff, small misalignments may well lead to
unequal distribution of the load where there are more than two jack
catches. It is thus good practice to distribute this load to two diagonally
opposite jack catches.

Case 2: Jack catches mounted on buffers.


Where jack catches are mounted on buffers, it is possible to use energy
considerations to determine the buffer force. It can be shown that for a
body of mass m, falling a distance H, onto a spring of stiffness K, the
maximum spring force is given by:
 
2 HK
F=mg 1 + 1 + 
 mg 
 
where g = gravity acceleration constant
= 9,81 m/s2
In the case of jack catch loads, the mass falling is the full mass of the
conveyance, mc, the stiffness is the buffer stiffness, Kj, multiplied by the
number of jack catches, as the greater flexibility results in the conveyance
falling onto all the jack catches, and H is determined by the detailing of the
jack catches. Thus the force on each catch is:

mc g  2 HnK j 
Fc = 1 + 
2  mc g 

where n = number of jack catches
Note: Where the jack catch load is determined on the basis of this type
of rational analysis, it would be prudent to use a load factor higher
than the normal factor of 1,05. The normal live load factor of 1,6 is
recommended.

79
July 01

5.6.3 Roof impact loads

20 000 N is regarded as a reasonable level of accidental load due to an


object falling on the roof of the conveyance.

Under emergency conditions permanent plastic deformation is considered


to be acceptable, so no deflection calculations are required.

5.6.4 Application of Emergency Loads

It is unusual in codes of practice for design, to specify how loads are to be


applied to structures. In this code however, the manner in which the
emergency loads are to be applied is specified. This is done mainly
because the dynamic and emergency loads do not lead to static
equilibrium conditions, which complicates the calculation of the member
forces caused by these applied loads considerably.

The rope break load is distributed throughout the conveyance in proportion


to the distribution of mass, as shown for a Job's bridle, for example, in
figure 31 and tables 11 and 12. A more rigorous analysis would require a
formal dynamic analysis of the entire conveyance, but it most cases the
results would not differ significantly.

It should be noted that the above distribution of emergency loads is true


for both the rope break and jack catch emergency loads. The points at
which the load is resisted, ie the support point, is however different. The
rope break load is resisted at the rope attachment point. The jack catch
load is resisted at the jack catch lugs, with the total load being carried on
all the lugs, or on two diagonally opposite lugs, depending on whether
buffers are used or not.

The possibility of a conveyance colliding with, or lodging against, some foreign


body in the shaft also exists. This may lead to the rope break load being
applied eccentrically if the foreign body is lying across one side of the
compartment. In the extreme, the rope break load may conceivably be applied
at the corner of the transom, leading to a large distorting moment as shown in
figure 32. A design to resist this magnitude of bending moment would lead to
an unreasonably heavy bridle, and the actual collision behaviour is impossible
to predict. This case is thus not specified by the code, but the engineer should
nevertheless ensure that there is some resistance to out-of-square distortion of
the conveyance, particularly for cages which have doors at both ends, and
skips which do not have an integral bridle.

When designing skips it should be noted that the skip body is not
considered to be subjected to increased bursting pressures during
emergency load events. This is likely to result in quite large plastic
deformations to the skip body in the event of an accident occurring, but
these deformations will lead to movement in the contained rock and
reduction in the pressure.

80
July 01

V V
3 4 3 4 4
5
1 1
1 1

2 2
3 3

2 2

A) Cage in Job's Bridle B) Skip in Job's Bridle

Figure 31: Job's Bridle – Load application points

81
July 01

Table 11 – Job's Bridle Containing Cage –


Assumed application of load

Load Application Applied Load Applied Load


point Cage full Cage empty
1 0,5 m3 + 0,5 p 0,5 m3
2 0,25 m1 0,25 m1
3 0,25 m1 0,25 m1
4 m2 m2

where m1 = Weight of bridle and attachments excluding top transom


m2 = Weight of bridle top transom
m3 = Weight of cage complete
p = Payload
V = (m1 + m2 + m3 + p)

Table 12 – Job's Bridle Containing Skip –


Assumed application of load

Load Application Applied Load Applied Load


point Cage full Cage empty
1 0,5 m3 + 0,25 p 0,5 m3
2 0,5 m4 + 0,25 p 0,5 m4
or 0,5 m4 + 0,5 p
3 0,25 m1 0,25 m1
4 0,25 m1 0,25 m1
5 m2 m2

where m1 = Weight of bridle and attachments excluding top transom


m2 = Weight of bridle top transom
m3 = Weight of skip body excluding door
m4 = Skip door and attachments
p = Payload
V = (m1 + m2 + m3 + m4 + p)

82
July 01

Collision Rope
Force Force

Moments tending to
distort the conveyance

Figure 32: Out-of-Square Distortion

83
July 01

6. LOAD FACTORS AND LOAD COMBINATIONS

6.1 General

The load factors and combinations to be considered for design are


specified for normal operating conditions, and also for emergency
conditions.

6.2 Operating Conditions

6.2.1 Partial Load Factors

The partial load factors specified essentially follow those given in SABS 0160.
SABS 0160 specifies a partial factor if 1,3 for the ultimate limit state for
contained water loads, because they are very accurately known. This code
specifies the same factor for tailrope loads, as these are also accurately
known. A reduced factor of 1,4 is specified for personnel loads, because it is
not possible to exceed the design assumption of 8 people per square metre.

6.2.2 Load Combinations

A general load combination equation is given, which is similar to that


specified by SABS 0160. The load combination factors, are however
omitted, as the times and positions at which forces are applied during the
hoisting cycle are well determined. It is thus not considered to be
appropriate to use these reduction factors for conveyance design.

There is, however, a wide range of different load combinations to which


conveyances are known to be subjected.

Typical load combinations are suggested in tabular form, table 13 dealing


with vertical shafts, and table 14 dealing with inclined shafts. In principle
these two tables are similar, but certain different loads apply. For instance,
conveyances in inclined shafts run on rail tracks, rather than between
guides, so that slipper and rubbing block loads do not apply.

When combining loads in inclined shafts it is important to note the direction


in which the loads are applied. All loads relating to acceleration,
deceleration and rope break are parallel to the axis of the shaft, whilst
other loads are vertical.

It will be noted that some different loads are specified for inclined shafts.
The roller, rubbing block and slipper plate loads are replaced by a lateral
track load, and tail ropes are not used in inclined shafts, so the tail rope
load is omitted. It is, however, conceivable that a friction winder might be
used in an inclined shaft, requiring the use of tail ropes supported on
idlers, in which case the tail rope load must not be omitted from the

84
July 01

design. The “trap-rail” system uses a head and trailing rope arrangement,
in which the rope is held within a rail, and released when the conveyance
passes. This enables hoisting around corners, and down an incline shaft
with widely varying angles of inclination along its length. The appropriate
loads and load combinations to be used, must be carefully assessed by
the Designer under these conditions.

6.2.3 Fatigue Load Combinations

Section 4 of the code describes various types of loadings to which


conveyances are subjected. The loading and unloading is repeated
thousands, and sometimes hundreds of thousands of times during the
working life of a conveyance. Repeated loading and unloading, even if the
yield stress of the material is never exceeded, may result in cracks forming
in the material, leading to eventual failure. This phenomenon is known as
fatigue, and appropriate stresses are discussed in Section 3.

This clause defines what loads are to be considered in fatigue


calculations. Clause 7.4 defines the numbers of cycles of each load which
must be applied. It should be noted that the serviceability load factor of 1,0
for all loads, is specified for fatigue design. This is because fatigue is
always related to working stresses.

In general all the loads defined, with the exception of self weight and
emergency loads, may lead to cycling stresses, and must thus be
considered for fatigue. It will be found, however, that certain loads are
typically significant in terms of fatigue life of conveyances and others are
not. The fatigue calculation in Tables 15 to 17 show the different influence
of all the specified loads. Emergency loads are not considered as a fatigue
load because no conveyance is used after an emergency event without
very extensive overhaul and repair, if indeed it is ever used again.

6.2.3.2 Slipper action during travelling

It should be noted that no statistical factor is applied to this load here, in


contrast to the load specified in clause 5.2.3.1.2. What is desired for
fatigue design is an estimate of the likely load levels.

6.3 Emergency Conditions

Partial Load Factor, γE

Under emergency conditions, only the ultimate strength limit state is


considered to be relevant. Any conveyance which is subjected to
emergency loading will be fully inspected before it is used again, if indeed
it can be used again. The loads are thus only applied once, before careful
inspection, so that fatigue is not a criterion. Provided the conveyance does
not break, deflection is considered to be irrelevant, even if it leads to doors

85
July 01

jamming or some other malfunction of equipment. There are thus no


serviceability factors quoted.

The rope break load is accurately known, and substantial damage is


tolerable, so a rather low partial load factor of 1,05 is considered
appropriate.

86
July 01

Table 13 — Load combinations for cages and skips in fixed vertical guide and rope guide systems
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19
General operating loads Man winding Material and equipment winding Roof Rock winding
Lateral imposed load Winder system loads Floor loads Floor loads Underslung loads Loads Internal loads
Holding Static Impact Static Impact Impact
Permanent device/Kep Roller or Slipper Acceleration Trip-out Tail rope Standing Equip. or Rolling stock Other impact Vertical or Vertical or Static in Impact Impact
engagement rubbing block plate personnel rolling stock loads Horizontal Horizontal Transit Filling Discharge
Vertical Horizontal Vertical and Horizontal
Loading symbol Gc Ki Kc or Kr Hf or Hr Hs Ao At T P M Cv CH M U U po p1 & p2 p3
Partial factor γfo γfi γfi γfi γfi γfi γfi γfi γfi γfi γfi γfi γf i γfi γfi γfi γfi
Loading of cage 1.2 1.3 1.6
1.0 1.0 1.0
Loading of cage 1.2 1.3 1.6
1.0 1.0 1.0
Loading of cage 1.2 1.3 1.6
1.0 1.0 1.0
Loading of cage or skip 1.2 1.6
1.0 1.0
Loading of skip 1.2 1.3 1.6
1.0 1.0 1.0
Unloading of skip 1.2 1.3 1.6
1.0 1.0 1.0
Holding of cage or skip at station 1.6
1.0
Travelling (cage) 1.2 1.6 1.6 1.3 1.4
xxxxx 1.0 1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.6 1.3 1.4
1.0 1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.6 1.3 1.4
1.0 1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.6 1.3 1.4
1.0 1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.3 1.4
1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.3 1.4
1.0 1.0 1.0 1.0
Travelling (cage or skip) 1.2 1.6 1.6 1.4
1.0 1.0 1.0 1.0
Travelling (cage or skip) 1.2 1.6 1.6 1.4
1.0 1.0 1.0 1.0
Travelling (cage or skip) 1.2 1.6 1.4
1.0 1.0 1.0
Travelling (skip) 1.2 1.6 1.6 1.3 1.4
1.0 1.0 1.0 1.0 1.0
Travelling (skip) 1.2 1.6 1.6 1.3 1.4
1.0 1.0 1.0 1.0 1.0
Travelling (skip) 1.2 1.6 1.3 1.4
1.0 1.0 1.0 1.0
Roof loading 1.2 1,6
1.0 1,0
Key to partial load factors
1.6 ← Ultimate limit state
Serviceability limit state → 1.0

87
July 01

Table 14 — Load combinations for cages and skips in inclined shaft systems
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
General operating loads Man winding Material and equipment winding Roof Rock winding
Winder system loads Floor loads Floor loads Underslung loads Loads Internal loads
Lateral Static Impact Static Impact Impact
Permanent imposed Accel. Trip-out Seated Equip. or Rolling stock Other Vertical or Vertical or Static in Impact Impact
loads personnel rolling stock impact Horizontal Horizontal Transit Filling Discharge
loads
Track level Vertical Horizontal
Vertical and Horizontal
Loading symbol Gc Ht Ao At P M Cv CH U U po p1 & p2 p3
Partial load factor γfo γfi γfi γfi γfi γfi γfi γfi γfi γfi γfi γfi γfi γfi
Loading of cage 1.2 1.6
1.0 1.0
Loading of cage 1.2 1.6
1.0 1.0
Loading of cage 1.2 1.6
1.0 1.0
Loading of cage or skip 1.2 1.6
1.0 1.0
Loading of skip 1.2 1.6
1.0 1.0
Unloading of skip 1.2 1.6
1.0 1.0
Travelling (cage) 1.2 1.6 1.6 1.4
1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.6 1.4
1.0 1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.4
1.0 1.0 1.0
Travelling (cage) 1.2 1.6 1.4
1.0 1.0 1.0
Travelling (cage or skip) 1.2 1.6 1.6 1.4
1.0 1.0 1.0 1.0
Travelling (cage or skip) 1.2 1.6 1.4
1.0 1.0 1.0
Travelling (skip) 1.2 1.6 1.6 1.4
1.0 1.0 1.0 1.0
Travelling (skip) 1.2 1.6 1.4
1.0 1.0 1.0
Roof loading 1.2 1.6
1.0 1.0
Key to partial load factors
1.6 ← Ultimate limit state
Serviceability limit state → 1.0

88
July 01

89
July 01

90
July 01

91
July 01

7. DESIGN PROCEDURES

This clause of the code defines what codes are generally applicable to
design of the structural members of conveyances. It also defines certain
special requirements which apply for particular members and loading
types.

7.1 Design Loads

This clause simply states the obvious, that loads are to be determined as
specified in this code.

7.2 Design Codes

The design codes specifically permitted by the code for design of structural
members are those used in normal structural design practice in South
Africa for steel and aluminium. These are all limit states design codes.
BS8118 Part 1 deals with structural design of aluminium members, whilst
BS8118 Part 2 deals with such matters as corrosion and the strength of
the heat affected zone.

Limit states codes are increasingly being applied internationally as more


reliable actual factors of safety can be achieved. In the case of this code,
the only limit states considered are the ultimate and fatigue conditions.
Deflections are small in all cases, as spans are short.

The designer may however feel that for a particular very large cage,
deflections cannot be overlooked. In such a case, the appropriate limits
suggested by SABS 0160 (1989) should be used as guidance. There are
no close tolerance operational constraints applicable to conveyance
deflections.

7.3 Design for Emergency Loads

The design philosophy for emergency loads is that damage is acceptable,


provided conveyances do not fail completely.

Structural steel is usually considered to have failed when yield occurs and
there is large deformation with little or no load increase. Strain hardening
does however add further strength to steel members before actual failure
occurs, provided that local buckling does not take place. Hymers (1992)
has shown at least a 20% increase in the strength of steel transoms
beyond their yield strength. The code thus allows the use of a resistance
factor of 1,1 instead of the 0,9 which is generally used by SABS 0162. This
reflects a 22% increase in strength.

Local buckling does not occur in tension members. Bridles are essentially
tension members, so the higher resistance factor may be applied to any

92
July 01

steel bridles. Local buckling also does not occur where steel members
have "plastic" section dimensions. In order for a section to be a plastic
section, the following dimensional limits must be met:
B
145
B T≤
fy
T
1100
h t≤
fy

where fy = material yield stress


t h
For the commonly used steel grade,
GR 300, these limits become:

B T ≤ 8,37
ht ≤ 63,5

It should be noted that conveyance transoms are frequently made up of


two channels, as already indicated in figure 3. SABS 0162 only allows the
use of the plastic moment capacity for plastic sections which are also
doubly symmetric. For this purpose, the two channel transom may be
considered as one, doubly symmetric section.
Certain further important design considerations arise from the tests
reported by Hymers (1992).
(a) For deep transoms, where the span/depth ratio is less than 4, shear
lag effects in the flanges are significant and probably reduce the
ultimate bending strength. However, on a short span, the load
spreading effect of the deadeye, or thickener plates in the vicinity of
the drawbar pin, significantly reduces the bending moment. Hymer's
(1992) quotes a 15% reduction. The resistance factor of 1,1 can thus
be used, but only if the load is assumed to be a central point load.
Shear failure of the transom web may predominate in deep transoms,
so it cannot be ignored. The resistance factor of 1,1 is also applicable
to shear failure.
(b) Where bridle angles or channels are connected over the full depth of the
transom, bending and eccentricity effects reduce their tensile resistance.
Where they are only connected to the bottom of the transom, shear lag
effects reduce their tensile resistance. It appears that in both cases the
resistance reduces to about 70% of its full value. It is thus recommended
that the bridle plates are widened, or that thickener plates are used, at the
transom to bridle joint and for some distance below this joint.

Aluminium does not exhibit the same yielding and strain hardening
behaviour as steel, so no increase in resistance factor is permitted.

93
July 01

7.4 Fatigue

The varying loads to be considered for fatigue design are defined in clause
6.1.3. In addition to the load magnitude, it is necessary to know the
number of cycles of each load.

The winding schedules for different conveyances vary widely at different


mines, so the design engineer must establish the actual winding schedule
for any specific conveyance. The winding schedule for conveyances in a
typical high production shaft may be as shown in table 18.

Table 18: Example of Conveyance Winding Trips

Conveyance Kibble (during Skips Cages


sinking)
Trips per hour 5 10 6
Hours per day 12 20 20
Days per month 30 24 26
Months per year 12 12 12
Trips per year 21 600 57 600 37 440
Life, in years 1 1/2 3 3
Total trips 32 400 172 800 112 320

A winding "trip" is the term used to describe the movement of a


conveyance from surface to an underground station and its return to
surface, including all loading and unloading operations undertaken. So, for
example, a single trip for a skip would be as follows:

stationary at the tipping level in the headgear.

- accelerate down shaft.


- travel at constant hoisting speed down the shaft.
- decelerate to bottom loading station.
- creep at very slow speed into position for loading.
- load skip with ore.
- accelerate up shaft.
- travel at constant hoisting speed up the shaft.
- decelerate at top of shaft
- creep at very slow speed into tipping position.
- tip ore into chutes.

94
July 01

This process is shown schematically in figure 33.

It should be noted that this is twice the “duty cycle” time usually calculated
on winder duty cycle sheets. The duty cycle time is the time between
stopping an empty skip at the loading box and stopping the full skip at the
tip.

Winding Speed
8
Full speed
up

7 9

10
Standing still 1
6 11
Time
5
2
4

Full speed 3
down

Figure 33: Skip Winding Trip

A trip for a cage may be somewhat more complex, with loading or


unloading at different levels.

Each winding trip implies load variations, which leads to stress cycles in
the members of any conveyance. Two different categories of cycling loads
may be distinguished.

The first is those due to the operational function of the winder, ie forces
induced by loading, unloading and acceleration of the winder. These are
essentially applied once each time the conveyance is loaded. The second
is those loads due to the shaft guide misalignments, which are applied as
the conveyance travels in the shaft.

(a) The main operational load cycles derive from four sources. These are
the payload weight, impact during loading or unloading, acceleration
or deceleration, and the tailrope. The tailrope weight on a conveyance
is a maximum when the conveyance is at the top of a shaft, and a
minimum of close to zero when the conveyance is at the bottom of the
shaft. This weight variation can be high. Consider, for example, a
conveyance having 2 50 mm non-spin head ropes, and operating in a

95
July 01

2 000 m deep shaft. The tailrope weight variation is about 430 kN,
which is 11% of the rope break load. Typical (simplified)
measurements of the rope tension at the conveyance are shown in
figure 34 for one winding cycle. In this example, a conveyance without
a tailrope is shown.

Several different cycle counting techniques have been proposed, but it


is suggested that the cycles reflected in figure 34 should be assessed
as:

a
Rope Tension
Acceleration
Loading

Dead load
plus payload
Deceleration

Deceleration

Dead load
a
Acceleration
Tipping

Travelling down Travelling up


Time

Figure 34: Typical Rope Tension

1 major cycle between points marked a

2 acceleration cycles

0, 2 or 4 bounce cycles, depending on holding devices used.

The payload is always applied, as are the acceleration cycles. However the
loading and unloading bounce due to dynamic overload may be eliminated,
or at least reduced, by the use of certain conveyance holding devices, as
discussed under clause 5.2.2. There is no loading bounce during man
hoisting because the load is applied slowly. The load cycles during each
hoisting cycle must thus be carefully assessed for each individual case.

(a) Secondary operational load cycles are applied to specific members of


the body of the conveyance, such as cage floor beams and buffer
beams.

96
July 01

Horizontal loads on buffers are applied once for each loading


operation at a magnitude equal to the load defined, as shown in
figure 35.

Vertical material loads on cage floor beams may be applied up to four


times for each loading operation at a magnitude which may be taken
as 1½ times the load defined, as shown in figure 34. This applies to
isolated beams spanning across the cage. For larger cages with a grid
of floor beams, axle loads are not fully applied and removed as each
axle passes across the beam, so less cycles are appropriate. Beams
towards the back of cages only have one axle passing over them, so
only two cycles are applied for each loading operation.

Impact Loads

Horizontal
Buffer

Vertical

Axle 1 Axle 2 Axle 2 Axle 1

Loading Unloading

Figure 35: Fatigue on Cage Members

(c) The application of loads due to shaft guide misalignment depends on


the condition of the shaft.

The shaft condition factors given in table 4 of the code are derived
mainly from information given in the COMRO Shaft Steelwork Design
Guidelines. Different factors are given for four different shaft
conditions. The most common method of assessing shaft condition is
the use of decelerometer traces, which are described in the chapter
dealing with maintenance.

It is not necessary to assume that an entire shaft has the same


classification. It is possible that different positions of a shaft may be
subjected to different ground movement, or different installation
accuracy. Strength calculations for conveyances must be based on

97
July 01

the worst classification within a shaft, but fatigue calculations should


be based on a rational assessment of the portions of the shaft that fall
into different classifications.

1. Good shaft. This refers chiefly to good alignment of the shaft


guides. In order to achieve this, initial installation must be good,
and regular maintenance must include guide realignment.
Decelerometer trace values would typically not exceed 3 m s2 in
a good shaft.

2. Average shaft. This shaft is not as well installed or maintained as


a good shaft. Decelerometer trace values would typically not
exceed 5 m s2, but may reach 7 m s2 at a few isolated places.

3. Poor shaft. This is the worst shaft condition normally


encountered. A shaft with a worse condition is generally
considered unacceptable. Decelerometer trace values would
typically not exceed 7 m s2 but may reach 10 m s2 at a few
isolated places.

4. No rollers. This is the shaft condition to be used for conveyances


which have no rollers.

The factors given in the table are the following:

1. Maximum misalignment, e. This is the maximum local


misalignment, or "moving beam" misalignment, which is
discussed in detail in the chapter dealing with maintenance.

2. Roller load factor, αr. The lateral roller load defined in clause
5.2.3.1.1 is the maximum value of this load. The typical roller
loads required to guide the conveyance in good or average
shafts will be less than this maximum value. It is possible to
define the roller load as:

R = D+P
where D = dynamic load
P = preload

The preload, P, is set very low in almost all shafts because the
guide rollers in common use do not last long if any significant
preload is applied. Thus:

R ≈ D

The dynamic load is a function of shaft condition, so


conveyances in average and good shafts are subject to roller
loads which are lower than the maximum.

98
July 01

3. Slipper load distance, ss. The maximum slipper load is applied


more frequently in poor shafts than in good shafts. The COMRO
guidelines define the number of slipper load occurrences as a
percentage of buntons passed. The distance defined in the code is
based on this percentage and the assumption that bunton spacing
is 5 m. The slipper load is applied once for every distance, ss,
through which the conveyance moves during hoisting. Thus, for an
average shaft, 1 900 m deep, the number of slipper load cycles is:

1900
DOWN : = 38 for empty conveyance
50
1900
UP : = 38 for full conveyance
50

For each slipper load application there is assumed to be an


equal slipper load on the opposite slipper a short time later. The
magnitude of slipper load cycles for most typical arrangements of
slippers should thus be taken as twice the slipper load.

4. Roller load distance, Sf. This is similar to the slipper load distance,
but applies to the roller load. Note that slipper load and roller load
are not additive, so there is never a case where the roller load
increases the magnitude of the slipper load cycles.

The code permits the extrapolation of these shaft condition factors. Linear
extrapolation, as shown in figure 36 for maximum guide misalignment, is
envisaged.

The other factors are extrapolated as:

αr = 0,25 x EF

Ss = 100 EF

Sf = 20 EF

where EF is the extrapolation factor from figure 36.

The selection of the shaft classification should be carefully considered, as


it has a very important influence on the fatigue life of conveyances. If a
conveyance in a class C shaft has a particular fatigue life, the fatigue life
would be increased by 16 times if the shaft were upgraded to class B and
by 256 times if the shaft were upgraded to class A. Two simple examples
in Table 19 illustrate this:

99
July 01

Table 19: Example Fatigue Lives

Fatigue Life
shaft class
Example 1 Example 2

A 256 years 5 years

B 16 years 4 months

C 1 year 1 week

It should be noted that, in general, fatigue is not a critical design criterion


for conveyances, because of the high safety factors applicable to working
loads. A fatigue analysis of the skips at Vaal Reefs 10 shaft, assuming
10% slamming and a class D fatigue detail, showed the bridles to have a
50 year life, and the transom a 25 year life. Fatigue damage has, however,
been known to occur in some specific locations:

(a) The bridles of overturning and swingbody skips have cracked at the
return stops, which have a high impact load applied each time the
skip is tipped. This cracking is due to bending stress about the x-axis.

(b) The bridle on a fixed body skip have cracked just above the top of the
skip body, due to bending stresses about the y-axis. This was due to
high slamming forces in a poorly designed shaft.

(c) Skip body plates have cracked in areas of high ore impact, where
thin wear plates were used, and where stiffeners were intermittently
welded. They have also cracked at the pivot bar reinforcing plates
and at the door hinges on swingbody skips.

(d) In cases where guide rollers have been mounted on brackets, above
the top transom, or below the bottom transom, these brackets have
been subjected to fatigue damage due to the bending stresses
resulting from roller load fluctuations.

(e) The front floor beam, in cages used for the transportation of heavy
items of equipment, has sometimes suffered fatigue damage.

Further examples of typical fatigue life calculation are included with the
skip and cage design examples in Appendix III.

100
July 01

Extrapolation Factor, EF
8

10 20 30
Maximum Guide Misalignment, e (mm)

Figure 36: Shaft Condition Factor Extrapolation

101
July 01

CONVEYANCE LAYOUT CONSIDERATIONS

The design of conveyances is subject to a number of practical considerations, which


derive from operating conditions and requirements, rather than technical design
constraints.

1. CONVEYANCE MEMBER SIZES

1.1 Typical Minimum Member Sizes

In order to ensure robustness of conveyances to withstand everyday


bumps and handling into and out of the shaft, certain minimum sizes are
commonly applied. Table 20 gives minimum sizes for skip members,
reported by Smith (1992), and Table 21 for cage members. These
minimum sizes are based purely on common experience at mines.

Table 20 : Common Minimum Sizes for Skip Members

Member Minimum Size

Rectangular skip up to 8 tons


Body plate 5 mm steel
Door plate 6 mm steel

Rectangular skip greater than 8 tons


Body plate 6 mm steel 8 mm alum.
Door plate 8 mm steel 8 mm alum.

It should be noted that the information given in these tables is neither


comprehensive nor a requirement of the code. It is not comprehensive
because experience with operating conveyances is not comprehensively
documented, and because new concepts and materials may not have
been sufficiently used to develop sound experience with them. It is not a
requirement of the code because it is not based on any of the specified
loads or design criteria.

For example, Smith (1992) shows that the specified ore pressure loads in
a typical skip can be resisted by a 3 mm body plate, but this is likely to be
buckled or damaged by the slinging required to manoeuvre the skip into
the shaft.

102
July 01

Table 21 : Common Minimum Sizes for Cage Members

Member Minimum Size

Corner angles 50 x 50 x 5 Steel


50 x 50 x 6 Aluminium
Floor and roof plate 4,5 mm Steel
4,5 mm Aluminium
Side plate 3 mm Steel
4,5 mm Aluminium

1.2 Skip Stiffeners

The most common sections used for skip body stiffeners are angles, and
channels formed by bending plates. Bent plate channels are commonly
preferred because they provide greater stiffness for the same mass.
Appendix II lists properties for a range of bent plate stiffeners.

Smith (1992) has found that stiffener spacing should not exceed 500 mm,
and that more closely spaced smaller stiffeners may be preferable.

Typical angle stiffeners are 70 x 70 x 6 steel angles.

Typical channel stiffeners are bent from 6 mm plate into channels 60 x 40


to 140 x 70.

1.3 Corrosion Allowance

A corrosion allowance of 2 mm is frequently included in thickness of skip


body plates. This appears to be adequate for a total skip life of 6 years.

2. CAGE VENTILATION

An allowance of 0,1 to 0,143 m2 person is typically required in the side


panels of cages for ventilation purposes. Any floor level openings should
be manufactured from 6 mm thick perforated plate and higher level
openings are typically fitted with thinner perforated plate or woven wire
mesh of such size as to prevent persons pushing their fingers through the
holes.

One third to half of the total area is usually allowed at the bottom of the
cage sides and the remainder at the top, to ensure good circulation of air.

103
July 01

3. SKIP LINERS

Skip liners serve two functions. They reduce ore impact pressures, and
they protect the skip from abrasive wear. Rubber liners, with a thickness of
50 mm to 100 mm appear to be the most effective in high impact areas,
whilst cascade liners, or wear resistant steel plates are the most effective
liners in high wear areas.

Where rubber liners are used their edges need protection, to prevent
damage due to ore impact. It is recommended that all liners in a skip
should be the same size, to facilitate replacement, and to allow rotation of
liners between high wear and low wear areas of the skip.

104
July 01

FABRICATION AND MAINTENANCE

Fabrication and maintenance are not specifically included in the code, but it will be
clear that some clauses rely on fabrication tolerances and maintenance conditions.
The code relies on normal good practice in fabrication, even where tolerances are not
specified. It is thus considered appropriate to discuss fabrication in this commentary.

1. CONVEYANCE FABRICATION

The fabrication of conveyances should generally be in accordance with


SABS-1200H. It will commonly be found, however, that mining companies
have their own, more stringent requirements which will take precedence.

1.1 Welding

Welding leads to metallurgical changes in both steel and aluminium, so it


is usually not permitted on the main safety critical members. It is never
permitted on aluminium members.

Where welded elements carry significant stresses, continuous welding


should be used in preference to intermittent welds, to reduce the likelihood
of fatigue cracking.

1.2 Fasteners

All conveyances are subjected to some level of vibrations, which has


frequently resulted in ordinary bolts shaking loose. The preferred fasteners
for steel conveyances are thus rivets or swage-lock fasteners. The
preferred fasteners for aluminium conveyances are swage-lock fasteners.
An exception to this is the fasteners used for skip liner plates. These must
be frequently removed and replaced, so the most popular fasteners are
ordinary Gr 8.8 bolts. For removal, they are simply flame cut off.

2. CONVEYANCE EXAMINATION & REPAIR PROGRAM

Details of the use, inspection, maintenance and responsibilities for


conveyances rest with the particular mining house, although certain
requirements are listed by the Mines and Works Act. However, a
discussion of typical procedures is necessary for a proper understanding
of the code.

The objective of this section is to describe how conveyances are inspected


and maintained on a regular basis so that dangerous situations do not
develop, whilst ensuring that economics are considered and that
conveyances are not stripped and rebuilt unnecessarily.

105
July 01

2.1 Inspection and overhaul of conveyances

Three distinct types of inspection and overhaul are typically carried out on
conveyances.

(a) The Mines and Works Act (Regulation 16.74) requires a daily
examination of all conveyances. Guide rollers and door mechanisms
are checked and adjusted, and the conveyance is visually inspected in
the shaft.

(b) A major inspection is an inspection, which may be visual or include


non-destructive testing procedures, and which is carried out in the
workshop at the mine.

(c) A major overhaul involves stripping down the safety critical members
of the conveyance for careful nondestructive examination, cleaning
and repainting, and then rebuilding of the conveyance. This may be
carried out by the mine, or by a conveyance supplier.

Rope front ends are tested on a six monthly basis. The conveyance can
be removed from the shaft when the rope front end is cut and this makes it
a convenient time to conduct a major inspection, or a major overhaul, of
the conveyance.

Thus, after six months of service, the conveyance is removed from the
shaft and is delivered to the workshops for a major inspection which
includes non-destructive examination of the suspension gear, whilst
another conveyance is in service. Six months later, the conveyance is put
back into service for a further period of six months. At the end of that
period i.e. after a total 12 month service, the conveyance is removed again
from the shaft but this time, a major overhaul is conducted when the
conveyance must be dismantled completely. It is not necessary to replace
any items (including bridles and transoms), during that major overhaul
unless the nondestructive examination reveals that cracks are beginning to
develop in the safety critical components in which case replacement is
essential.

Once the conveyance has had a total service life of 3 years, it is strongly
recommended that the safety critical components e.g.: transoms, bridles,
slinging eyes, hooks and pins are replaced. Corrosion, physical damage
and general fatigue will have brought these components to the end of their
useful life. Structural components other than safety critical items can be
repaired for re-use. The conveyance suspension gear is only replaced
after six years of actual service. In addition to these inspections
supplementary daily and weekly examinations are recorded whilst the
conveyance is in use in the shaft.

106
July 01

A typical schedule of major inspections and overhauls for conveyances is likely to be


as summarised in Table 22.

Table 22 : Major Inspection and Overhaul Schedule

Conveyance Period in Entire Safety Critical


age Service Conveyance members

6 months 6 months Inspection Inspection

18 months 12 months Overhaul Overhaul

30 months 18 months Inspection Inspection

42 months 24 months Overhaul Overhaul

54 months 30 months Inspection Inspection

66 months 36 months Overhaul Discard and replace

3. SHAFT ALIGNMENT AND MAINTENANCE

The shaft condition has a significant influence on the forces applied to


conveyances, and the likelihood of damage occurring. A discussion of
certain aspects of shaft alignment and maintenance is thus pertinent.

3.1 Shaft Alignment

There are several phases in the installation of well aligned shaft steelwork.

- Pre-Installation
Bunton sets are generally preassembled in a jig at surface to ensure
dimensional accuracy. The actual sizes of each guide length are
commonly measured, taking particular cognizance of the depth of the
guide. Guides are then sorted into lots of very nearly equal size. Any
one guide string is then installed using guide lengths from only one lot,
or ensuring careful grinding to size when transferring to a different lot.
This process ensures a minimum of size mismatch at guide joints.

- Installation
Bunton sets are aligned during installation, using plumb wires,
tensioned wires, or more recently lasers. Plumb wires are disturbed by
any airflow in the shaft, and laser beams are diffused by moisture and
dust and diffracted by temperature gradients. Both of these
procedures can thus only be used over a few hundred metres for each
set up, and under the right environmental conditions. Tensioned wires

107
July 01

are used between carefully surveyed top and bottom points, which
may also be several hundred metres apart.

- The guides are then attached to the buntons, and may be further
aligned using the plumb wires and templates which define the exact
guide positions.

- Replacement
Similar procedures are used if damaged shaft steelwork has to be
replaced.

3.2 Shaft Inspection


The condition and alignment of shafts can deteriorate with use and rock
strains in the vicinity of the shaft. Certain inspection procedures are thus
carried out.

− Regular Visual Examination


Every shaft is carefully visually examined along its entire depth every
one or two weeks. This will locate loose joints, bad corrosion, fatigue
cracking, or foreign bodies lodged on the shaft steelwork.

− Decelerometer Traces
Lateral accelerations of conveyances travelling in shafts may be
measured and recorded using 'decelerometers' (named thus because
the early units were developed to measure braking deceleration of
winders). Some mines take regular decelerometer traces in all their
shafts at six monthly or yearly intervals, whereas other mines have no
regular programme. The quantitative interpretation of decelerometer
trace results is difficult, depending on many factors such as the type of
equipment, used, the exact position of the measuring device, the
conveyance size, stiffness and load, and previous conveyance
movements in the shaft. A reasonable qualitative assessment of the
shaft alignment condition, and approximate locations of bad sections
can however be achieved. The procedure is also simple and quick to
execute.

Two typical decelerometer traces are shown in figure 37, one a "good"
measurement, and the other a "bad" measurement which indicates that
guide realignment is necessary at the position shown.

− Misalignment Measurements
Equipment is available (from ATD or CSIR) to measure the actual guide
alignment in certain compartment configurations. At present the only
configuration in which full measurements can be made is where there
are only two guides, located one on each side of the conveyance, as
shown in figure 38. Equipment may be developed in future to measure
misalignment in compartments with any guide configurations, but at
present this is not available.

108
July 01

Misalignment may be recorded in one of two ways. It may be recorded


as an absolute misalignment from the true plumb position in the shaft,
or it may be recorded as a "moving beam" value. A moving beam
misalignment is the maximum displacement of a guide from a straight
edge beam which has a length equal to twice the bunton space, and
which is held against the guide. This is shown schematically in
figure 42. The absolute and moving beam misalignments for the guide
in figure 42 are given in table 23.

Table 23: Guide Misalignment

Absolute Moving Beam


Bunton No Misalignment Misalignment
(mm) (mm)

1 2 -2

2 8 7

3 0 1,5

4 -5 1,5

5 - 13 - 0,5

6 - 20 -3

7 - 21 –

The dynamic behaviour of conveyances travelling in mine shafts is


determined by the local misalignment of the guide, rather than the
absolute misalignment. The important misalignment value is thus the
moving beam misalignment, which is given by:

q + qi+1
ei = qi − i−1
2

where qi is the absolute misalignment at butons i.

109
July 01

14 16 18 20 22 24 26 28
2
3m/s
2
-3m/s

Test no. 7 (15m/s )


14 16 18 20 22 24 26 28
2
3m/s
2
-3m/s

Test no. 9 (15m/s )

(a) Good measurement

P3 IPC P2 P1 Bank

Realign
guides

(b) Bad measurement, indicating realignment

Figure 37: Decelerometer Traces

110
July 01

Misalignment measurements possible

Misalignment measurements
currently not possible

Figure 38: Misalignment Measurements

Bunton 0
(0 mm)

Bunton 1
(2 mm)

Bunton 2
(8 mm)

Bunton 3
(0 mm)

Bunton 4
True Plumb Position

(-5 mm)

Bunton 5
(-13 mm)

Bunton 6
(-20 mm)

Bunton 7
(-21 mm)

Figure 39: Guide Misalignments

111
July 01

REFERENCES AND BIBLIOGRAPHY

REFERENCES

COMRO Design Guidelines for the Dynamic Performance of Shaft Steelwork and
Conveyances. COMRO User guide No 21, Johannesburg, 1990.

Greenway, M.E. Dynamic Loads in Winding Ropes – An Analytical Approach. Anglo


American Corporation of S.A. Ltd, 1989.

Hymers, .T. Strength of Conveyance Transoms and Bridles. Anglo American


Corporation of S.A. Ltd, Report no 92-S-32, Johannesburg, November 1992.

Minerals Act 1991 (Act No 50 of 1991) and the Regulations.

SABS 0160. The general procedures and loadings to be adopted for the design of
buildings. South African Bureau of Standards, Pretoria, 1989.

SABS 0162. The structural use of steel. South African Bureau of Standards, Pretoria,
1989.

SANS 10208:Part 1. Design of Structures for the Mining Industry. Part 1: Headgear.
South African Bureau of Standards, Pretoria, 1986.

Smith, P.N. Survey of Design parameters for Skips. M.Sc Thesis, University of the
Witwatersrand, Johannesburg, 1992.

Soutar. 1973.Ore Handling Arrangements in Shafts of the Republic of South Africa, in


Robinson, L.R., Bunt, E.A. and Kraft, K. (eds). "International Conference on Hoisting
– men, materials and minerals", S.A.I.M.E., Johannesburg 1973.

Taljaard, L.C.J. Metallurgical aspects of Winder Attachments and Suspension Gears.


In Robinson, L.R., Bunt, E.A. and Kraft, K. "International Conference on Hoisting –
men, materials and minerals", S.A.I.M.E., October 1973.

Van Rooyen, G.T. Materials for Winding Plant Components. J. of S.A. Inst. of Min.
and Metal., November 1972

112
July 01

APPENDIX III: DESIGN EXAMPLE

1. Shaft Section
In order to clarify the loads and design procedures embodied in the code a
typical cage design and a typical skip design are given. These designs
should not be regarded as comprehensive working conveyance designs,
but rather they are intended to highlight important aspects. Both the cage
and the skip design will be for conveyances in the shaft shown in figure 40.

2,0 m 2,0 m 2,0 m 2,0 m


0 (typ)

Shaft Diameter 9,5 m


200 x 100 x 8

Shaft Depth 2000 m


Skip Skip Skip Skip
3 2 1
1,0 m

350 x 100 x 8 0 1,0 m


5 4
Cage Cage 1,8 m

7 6

Average Shaft Alignment


2,5 m 2,5 m

Figure 40: Example Shaft Layout

Bunton end connections, particularly to the shaft sides, are flexible in most cases,
so the bunton stiffnesses are taken to be a rough average of the fixed ended and
the pin ended conditions. Bunton stiffnesses, at the guide locations numbered 1
to 7 in figure 40, are given in Table 23. For buntons spaced at 5 m, and using a
152 x 69,8 kg m tophat guide, the guide stiffness is also given in Table 23.

All conveyance design will assume the use of Grade 300W or Grade 300WA
materials, as appropriate.

113
July 01

Table 23: Bunton Stiffnesses (N m)

Bunton Stiffness Guide Stiffness


Guide Position
(N m) (N m)
1 4 138 000 2 679 000
2 1 775 000 1 545 000
3 1 745 000 1 522 000
4 1 608 000 1 600 000
5 1 745 000 1 522 000
6 1 780 000 1 740 000
7 980 000 732 000

2. Skip Design

2.1 Basic Skip Parameters

The skip to be designed is a 15t swingbody skip, with main parameterss as


given in Table24.

Table 24: Skip Parameters

Payload 15 000 kg
Hoisting Velocity 18 m s
Winder Acceleration 0,8 m s2
Winder trip out deceleration 5,0 m s2
Rope: Type 51 mm Comp Triang
UTS 1 900 MPa
Breaking load 1 994 000 N
Ore density 1 850 Kg m3
Holding devices None
Underslinging None

114
July 01

15 000
Required skip volume = = 8,11 m3
1850

8,11
Ore depth = = 4, 58 m
1, 29 x 1, 37

Skip payload = 15 000 x 9,8 = 147 000 N

147000 x 2000 x 103


Rope stretch under payload = = 1400 mm
3 512
103 x 10 x π
4

This gives the layout dimension of the skip as shown in figure 41.

850
Backing plates

1750
4900

1580
160

Rubber liner
1370

C.G
Steel wear .
plates
4600
70

1200

70 1290 70
Return stops
900

Tipping rollers
450

Figure 41: Skip Layout

115
July 01

2.2 Skip Loads

(a) Self weight Cl 5.2.1.2


Self weight is approximately given by table 7 in the commentary.
Assume construction in steel.
Dead weight ratios: Body 0,25
Bridle 0,17
Rubber liner 0,04
Steel liner 0,06
____
0,52
Attachments mass 4 rollers 400 kg
Jack catch lugs 60
Slippers 120
Pivot bar 80
Door equipment 180
Tipping equipment 160
____
1000 kg
Total dead load: Gc = (0,52 x 15 000 + 1000 ) 9,81 = 86 328N

(b) Lateral imposed load Cl 5.2.3


Using fixed guides: Roller load, Hf = 5 000 N
400me v 2e
Slipper plate load, Hs = Pb
L2
For worst compartment, guide 1, whilst travelling upwards in the shaft,
we have:
4138000
rk = = 1,544
2679000
4138000 ⋅ 5 2 319290
kb = =
m e ⋅ 18 2 me
86 328
mc = 15 000 + = 23 809 kg
9,8
86328 15000
I ¡Ö
9,8 x12
( )
1,43 2 + 7,65 2 +
12
( )
1,29 2 + 4,602 kg.m 2

= 44 461 + 28 530 = 72 992 kg.m2


23 809 72992
me = =2696 kg
72992 + 23 809 x 4,90 2
319 290
kb = =118
2696

116
July 01

From figure 1 we get:


Pb = 0,054

The slipper plate load is thus:


400 2696 18 2 0,01
Hs = 2(0,054 2 = 15094 N
5
Whilst travelling down in the shaft, we have:
Hs = 24 211 N (from similar calculations).

(c) Winder systems loads Cl 5.2.4


Normal Ao =
0,8
(86 328 + 147 000 ) = 19 028 N
9,8

Tripout At =
5,0
(86 328 + 147 000) = 118 924 N
9,8

(d) Rock loads Cl 5.5


Static Cl 5.5.1.1
R = 1,37 x 1,29 x 4,60 x 1 850 x 9,8 = 147 389 N
Bridle transom loads while filling Cl 5.5.1.2
RD = 1,5 x 147389 = 221 084 N
Load on tipping rollers (for bridle design) Cl 5.5.1.6 (comm)
Rt = 3 x 0,1 (147 389 + 86 328) = 70 115
Skip return load Cl 5.5.1.7
7000
Rs = 0,25 x 86 328 x = 24 766 N
6100
Gravity rock pressures Cl 5.5.1.3
po = 1 850 x 9,8 x 4,6 = 83 398 N m2
Pressure on door Cl 5.5.1.4
p1 = 1,0 x 83 398 = 83 398 N m2
Pressure on back of skip Cl 5.5.1.4
p2 = 0,5 x 83 398 = 41 699 N m2
Pressure on lower portion of skip back Cl 5.5.1.5
p3 = 1,5 x 83 398 = 125 097 N m2
Pressure on front and sides of skip Cl 5.5.1.4
p2 = p3 = 0,2 x 83 398 = 16 680 N m2

117
July 01

(e) Emergency loads Cl 5.6.1


Er = 1 994 000 N

2.3 Skip Element Design


2.3.1 Top transom

Design for emergency load.

1 994 000 ⋅ 1,58


Mu = 1,05 = 827 012 N.m
4

1 994 000
Vu = 1,05 = 1 046 850 N
2

Try 2 300 x 100 PFC with 4 100 x 20P cover plates

Check section proportions


100 − 9, 6
Flange: = 5,8 < 8,37 SABS 0162 Cl 11
15, 5
300 − 2 x 15 , 5
Web: = 28,0 < 63,5
9,6

This is thus a class 1 section.

   15,5  9,6 x (300 − 2 x 15,5 )2  3


Zp1 = 22 100 x 20 x 160 + 100 x 15,5 x 150 −  + mm
   2  4 

= 2 509 x 103 mm3

Mr = 1,1 x 300 x 2 509 x 103 x 10-3 N.m SABS 0162 Cl 13.5


= 827 970 N.m > Mu OK Cl 7.3
Vr = 2(0,55 x 1,1 x 9,6 x 300 x 300) N SABS 0162 Cl 13.4.1.2
= 1 045 440 N ≈ Vu OK

Design for Operating Conditions

Maximum load = Dead load + Rock load filling

= 1,2 x 86 328 + 1,6 x 221 084 N

= 457 328 N Cl 6.2.2

457 328 x 1, 58
Mu = = 180 645 N.m
4

118
July 01

457 328
Vu = = 228 664 N
2

MR = 0,9 x 300 x 2 509 x 103 x 10-3 N.m SABS 0162 Cl 13.5

= 677 430 N.m » Mu OK

VR = 2(0,9 x 0,66 x 300 x 300 x 9,6) N SABS 0162 Cl 13.4.1.1

= 1 026 432 N » Vu OK

(Note: It can be seen here that the transom strength for operating conditions
is much greater than required. The emergency load will always be
the critical design case.)
Design for fatigue
(assume 3 000 trips/month for 2 years)
Total number of trips = 3 000 x 24 = 72 000 trips
Cycling loads as recommended in the commentary.

(a) Major cycle load = Rock loading + 25% (assumed) of dead load
= 221 084 + 0,25 x 86 328 N
= 242 666 N (72 000 cycles)
242 666 x 1, 58 x 10 3
σa = = 38 MPa
4 x 2 509 x 10 3
(b) Acceleration cycles = 2 x Ao
= 38 056 N (144 000 cycles)
38 056 x 1, 58 x 103
σb = = 6,0 MPa
4 x 2 509 x 103

(c) Bounce cycles = 20% (assumed) of payload


= 0,2 x 147 000 N
= 29 400 N (288 000 cycles)
29 400 x 1, 58 x 103
σc = = 5,0 MPa
4 x 2 509 x 103
Assuming a class C fatigue detail, σa, σb and σc are all too low to have
any effect as the endurance limit is 70 MPa (SABS 0162 Cl 14)

(Note: This result is typical. Fatigue is unlikely to cause any problems on


transoms, provided that careless weld details are avoided.)

119
July 01

2.3.2 Bottom transom

Design for Emergency Load


1994000 ⋅ 1,58
Mu = 1,05 = 653 339 N.m
8
1994000
Vu = 1,05 = 1 046 850 N
2
Try 2 300 x 100 PFC with 4 100 x 12P cover plates

   15,5  9,6 x(300 − 2x15,5 )2  3


Zpl = 22100 x12 x156 + 100 x15,5 x150 −  + mm
   2  4 

= 1 978 x 103 mm3


MR = 1,1 x 300 x 1 978 x 103 x 10-3 N.m SABS 0162 Cl 13.5
= 652 740 N.m ≈ Mu OK Cl 7.3
Vr = 1 045 440 N ≈ Vu OK

2.3.3 Bridle Hangers

Design for Emergency Load

Each hanger will carry half the rope break load in tension.
1994000
Tu = 1,05 x = 1 046 850 N
2
Try 260 x 90 PFC with 250 x 10P backing plate at top

Tr = 1,1 x 4 630 x 300 N SABS 0162 Cl 13.2(a)


= 1 527 900 N > Tu OK
A ′ne = 0,85 x [4630 + 250 x10 − 2x26(8,8 + 10 )] mm2 SABS 0162 Cl 12.3.3.2
= 5 229 mm
Tr = 0,85 x 0,9 x 5 229 x 450 N SABS 0162 Cl 13.2(b)
= 1 800 269 N > Tu OK

(Note: A material factor of 0,9 must be used here as a factor of 1,1 with fu would
imply tensile rupture. The area includes the backing plate, and assumes 2
transverse rows of M24 bolts).

Design for Operating Loads

120
July 01

Two sections of the hangers are considered.

(a) At the pivot bar connection

Four loads occur here

86 328
(i) dead load tension is = 1, 2 x = 51 797 N
2

dead load bending moment = 51 797 x 0,03 = 1 554 N.m

(This bending results from the eccentricity of support of the pivot bar. It
is about the y-y axis of the channel)

221084
(ii) rock load tension = 1,6 x = 176 867 N
2
rock load bending moment = 176 867 x 0,03 = 5 306 N.m
15094 x0,85
(iii) slipper plate load bending moment= 1,6 x = 5 132 N.m
4
(This bending is about the y-y axis of the channel, and is calculated as
shown in the sketch below.)

HS

HS x 0,85
BM = 4

(iv) tipping bending moment 0, 85 x 6, 80


= 1,6 x 70 115 3 x N.m
at pivot bar 7, 65
= 28 254 N.m

(This bending is about the x-x axis of the channel. It assumes that
the tipping roller forces must be resisted through the pivot bar as
discussed in the clause commentary, and that slippers are located
at the top and bottom of the skip.)

Loads (iii) and (iv) cannot occur simultaneously, and neither can occur
together with the rock loading impact, so three different load
combinations must be checked.

121
July 01

(i) (ii)
+ + (iii)
1,5 1,5
51797 176 867
Tu = + = 152 443 N
1, 5 1, 5
1554 5306
Muy = + + 5132 = 9 705 N.m
1,5 1,5
Tr = 0,9 x 4 630 x 300 = 1 250 100 N SABS 0162 Cl 13.2
Mry = 0,9 x 56 300 x 300 x 10-3 = 15 201 N SABS 0162 Cl 13.2
152 443 9705
+ = 0,76 < 1,0 OK SABS 0162 Cl 13.9
1250100 15201

(i) (ii)
+ + (iv)
1,5 1,5
Tu = 152 443 N

Mux = 28 254 N.m


Tr = 1 250 100 N
Mrx = 0,9 x 370 x 103 x 300 x 10-3 = 99 900 N SABS 0162 Cl 13.5

(The channel is considered laterally supported by the pivot bar.)


152 443 28 254
+ = 0, 41 < 1, 0 OK SABS 0162 Cl 13.9
1 250 100 99 900

(i) + (ii)
Tu = 51 797 + 176 867 = 228 664 N

Muy = 1 554 + 5 306 = 6 860 N.m


228 664 6 860
+ = 0, 63 < 1, 0 OK SABS 0162 Cl 13.9
1 250 100 15 201

(b) At the return stops camelback position

Four loads occur here

(i) and (ii) are as for case (a), but with no bending moments.

(iii) slipper plate load bending moment

24211x1,10
= 1,6 x N.m
4
= 10 653 N.m

122
July 01

0, 45 x 7, 2
(iv) tipping bending moment = 1,6 x 70 115 x N.m
7, 65
= 47 513 N.m

1, 55 x 6,10
return-stop bending moment = 1,6 x 24 766 x N.m
7, 65
= 48 975 N.m

Load combinations to be checked:

(i) (ii)
+ + (iii)
1,5 1,5
51797 176 867
Tu = + = 152 443 N
1, 5 1, 5
Muy = 10 653 N.m
152 443 10653
+ = 0,82 < 1,0 OK SABS 0162 Cl 13.9
1250100 15201
(i) (ii)
+ + (iv)
1,5 1,5
Tu = 152 443 N

Mux = 48 975 N.m


152 443 48 975
+ = 0, 61 < 1, 0 OK SABS 0162 Cl 13.9
1 250 100 99 900

(i) + (iii) is better than for case (a)

Design for fatigue


Maximum axial tension fluctuation is:
221084
σa = + 0,25 x 86 328 = 26 MPa
2 x 4 630

This is below the endurance limit, so ignore it.


(a) at pivot bar connection:

Skipper plate load bending moment gives:


5132 x 103
σs = 2 x = 91 MPa
2 x 56,3 x 103

number of cycles is:


2 000
72 000 x = 288 000
50

123
July 01

Tipping force bending moment gives:

28 254 x 103
σt = = 76 MPa
370 x 103

number of cycles is 72 000

With joint b detail, Miners law gives:

0 + 0 = 0,0 SABS 0162 Annex K

This implies that fatigue will not be a problem at this position, within the
2 year life required.

(b) At the return-stop camelback position:

Slipper force bending moment gives:

10 653 x 103
σs = 2 x = 189 MPa
2 x 56,3 x 103

Tipping and return-stop stress:

47 513 x 103 48 975 x 103


σt = + = 261 MPa
370 x 103 370 x 103

(This arises because the skip must be opened by the tipping roller force,
and then closed against the return-stops during each winding cycle)

Detail b and number of cycles as for case (a).

288000 72000
+ = 0,88 SABS 0162 Annex K
550000 200000

This implies that fatigue cracking is unlikely to occur at this position


within the 2 year life expectancy.

2.3.4 Skip Body

This example will only cover the design of the back of the skip.
(a) Space stiffeners at 0,5 m centres.
Stiffener UDL = 1,6 x 41 699 x 0,5 = 33 359 N m
Stiffeners are welded at corners, so that:
33 359 x 1, 43 2
BM ≈ = 6 822 N.m
10

124
July 01

6 822 x 10 3
Zxx required = = 25 267 mm3
0, 9 x 300
Use 60 x 100 Bent Plate Channel x 6 thk

(b) Plate between stiffeners (continuous beams):


41699 x 0 , 52
BM = 1, 6 x = 1 390 N.m
12
1390 x 10 3
Zxx required = = 5 148 mm3
0, 9 x 300
5148 x 6
t required = = 5,6 mm
1 000
Use 6 mm Body Plate

(c) Lower 0,3 m of skip back, space stiffeners at 0,2 m.


Stiffener UDL = 1,6 x 125 097 x 0,2 = 40 031 N m
40 031 x 1, 43 2
BM ≈ = 8 186 N.m
10
8186 x 10 3
Zxx required = = 30 318 mm3
0, 9 x 300
Use 60 x 100 Bent Plate Channel x 6 thk

3. Cage Design
3.1 Basic Cage Parameters

The cage to be designed is a 3 deck, 120 man cage. Provision for a 10t
equipment load on a material car is to be made on the lower deck only. The
important parameters for design of this cage are given in table 25.

120
Cage area required = = 5 m2 deck .
3x 8

This gives the layout dimensions of the cage as shown in figure 42.

10 000 x 9, 8 x 2 000 x 10 3
Rope stretch under equipment load = 2
= 1 052 mm
103 x 10 3 x π x 484
Weight of men = 120 x 75 x 9,8 = 88 200 N

88 200 x 2 000 x 10 3
Rope stretch under man load = 2
= 946 mm
103 x 10 3 x π 484

125
July 01

Table 25: Cage design parameters

Number of men 120

Payload (on lower deck only) 10 000 kg

Hoisting Velocity 16 m/s

Winder Acceleration 0,8 m/s2

Winter trip-out Acceleration 5,0 m/s2

Rope: Type 48 mm Comp. Triang.


UTS 1 900 MPa
Breaking load 1760 000 N

Holding Device Kep (attached to centre of


top front beam)

Underslinging None

X
2080 Kep
2650

attachment
3725
350

0
52
Guide Channel

Guide Channel
1800

Bridle Hanger

C.G
2300

.
350

2000
2500

Figure 42: Cage Layout

3.2 Cage Loads

(a) Self weight Cl 5.2.1.2


Assume a cage factor of 0,55 on equipment load.
Gc = 0,55 x 10 000 x 9,8 = 53 900 N

126
July 01

(b) Kep Engagement Load Cl 5.2.2.1


K = 1,5 (53 900 + 10 000 x 9,8) = 227 850 N

(c) Lateral Imposed Load Cl 5.2.3


Using fixed guides, Roller Load, Hf = 5 000 N

400me Ve2
Slipper plate load, Hs = αh Pb
L2
For worst guide, guide 6, whilst travelling upwards or downwards in the
shaft, we have:
1780 000
rk = = 1,02
1740 000

1780000 ⋅ 52 173828
kb = =
me ⋅ 162 me

53 900
mc = 10 000 + = 15 500 kg
9, 8
Following example 2 in the commentary on clause 4.1.3,

I1 = Iy =
15500
12
(
2,002 + 7,452 ) = 76 857 kg.m2

I2 = Iz =
15500
12
(
2,002 + 2,502 ) = 13 240 kg.m2

15 500 ⋅ 76 857 ⋅ 13 240


me = kg
76 857 ⋅ 13 240 + 15 500 ⋅ 76 857 ⋅ 0, 9 2 + 15 500 ⋅ 13 240 ⋅ 3, 725 2
= 3 240 kg
173 828
kb = = 53,7
3 240
Pb = 0,038 Figure 1

400 ⋅ 3 240 ⋅ 16 2 ⋅ 0, 01
Hs = 2 x 0,038 = 10 086 N
52

127
July 01

(d) Winder System Loads Cl 5.2.4


Acceleration load

Men: Ao =
0,8
(53900 + 88200) = 11 600 N
9,8

Equipment: Ao =
0,8
(53900 + 98000 ) = 12 400 N
9,8
Trip-out load
Men: At =
5,0
(53900 + 88200) = 72 500 N
9,8

Equipment: At =
5,0
(53900 + 98000 ) = 77 500 N
9,8

(e) Man Winding Loads Cl 5.3


Floor load
P = 6 000 N m2
(f) Material and Equipment Winding Loads

Floor loads Cl 5.4


Static load Cl 5.4.1.1
M = 10 000 x 9,8 = 98 000 N

Impact loads (assume two equally loaded axles, and no spring buffers)
98 000
Cv = 2x = 98 000 N
2
CH = 0,5 x 98 000 = 49 000 N

(g) Emergency Loads Cl 5.6


Er = 1 760 000 N

3.3 Cage Element Design

3.3.1 Top Transom and Bottom Transom

The design of these elements is the same procedure as used for skips, so it
is not reproduced here.

3.3.2 Bridle Hangers

The bridle hangers are taken to consist of the main vertical member and two
diagonals at each deck level.

Design for Emergency Conditions

128
July 01

1760 000
Load on main vertical, Tu = 1,05 = 924 000 N
2
Try 250 x 16 P

SABS 0162 Cl 13.2(a)


Tr = 1,1 x 250 x 16 x 300 = 1 320 000 N > Tu OK
Tr = 0,85 x 0,9 x (250 x 16 - 2 x 26 x 16) x 450 N
= 1 090 584 N > Tu OK SABS 0162 Cl 13.2(b)

Load on Diagonals

Assume that each pair carries 1 3 of the total load.


924 000
Tu = = 195 429 N
2 x 3 x sin 52o

Try 100 x 10P

SABS 0162 Cl 13.2(a)


Tr = 1,1 x 100 x 10 x 300 = 330 000 N > Tu OK
SABS 0162 Cl 13.2(b)
Tr = 0,85 x 0,9 x (100 x 10 - 26 x 10) x 450 = 254 745 N > Tu OK

All lateral forces are resisted by the guide channels and not the bridle
hangers.

3.3.3 Bridle Channels


Two load cases are considered here, which may occur concurrently.
10 086 x 2,65
(a) Bridle bending moment = 1,6 N.m
4
= 10 690 N.m
(This is derived as shown on the sketch below.)

HS

HS x 2,30
BM = 4

129
July 01

98 000 x 0,10
(b) Load eccentricity bending moment = 1,6 x = 1 960 N.m
2x 4
(This assumes an eccentricity of 100 mm for the equipment load, and a
distribution of bending moments on each guide channel as shown for
collision forces in figure 29 of the clause commentary.)
Muy = 10 690 + 1 960 = 12 650 N.m
Try 240 x 85 PFC

SABS 0162 Cl 13.5


Mry = 0,9 x 46 900 x 30 x 10-3 = 12 663 N.m > Muy OK

3.3.4 Kep Attachment Beam


227 850 x 2, 0
Mux = 1,6 x 182 280 N.m
4

Try 305 x 165 x 46 I

Beam is laterally supported by the roof plate.

SABS 0162 Cl 13.5


Mrx= 0,9 x 722 x 300 = 194 940 N.m > Mux OK

Fatigue check

Assume 2 000 trips month with equipment load and using keps. This is
48 000 trips within a 2 year life of the cage.

182 280 x 103


Stress is σa = = 253MPa
722 x 103

With a class B detail, SABS 0162 Annex K gives a fatigue life of 250 000
cycles. This exceeds 48 000 OK

3.3.5 Floor Beams

Assume floor beams to be spaced at 0,9 m centres between the guide


channels and at each end of the cage.

(a) Personnel loads


6 000 x 0, 9 x 2, 0 2
Mux = 1,6 = 4 320 N.m
8

130
July 01

(b) Equipment loads


Assume each axle applies 2 equal loads to rails mounted 0,765 m
apart.
98000 x 1 (2,0 − 0,765 )
Mux = 1,6 2 = 48 412 N.m
2
Design beams to withstand equipment loads.
Try 200 x 75 PFC

Beams are laterally supported by the floor plating.

SABS 0162 Cl 13.5


Mrx = 0,9 x 191 x 300 = 51 570 N.m > Mux OK

Design check for emergency load


If the emergency load is to be applied as a distributed load on the
bottom deck (which may be done at the discretion of the Engineer – see
clause 4.5.3.1) we have:
1760 000
Pressure = = 352 000 N m2
2 x 2, 5
352000 x 0, 9 x 2, 0 2
Mux = 1,05 = 166 320 N.m
8
Mrx = 1,1 x 191 x 300 = 63 030 N.m < Mux No good

Try 300 x 100 PFC

Mrx = 1,1 x 534 x 300 = 176 220 N.m > Mux OK

These larger beams are required for the lower deck only.

The equipment load is only applied to the lower deck, so the upper
decks are only required to resist the personnel load.

Try 100 x 50 PFC

Mrx = 0,9 x 41,1 x 300 = 11 097 N.m > Mux OK

131

Anda mungkin juga menyukai