Defstan 00 970
Defstan 00 970
Section 3
Issue 15 Date: 13 Jul 2015
_______________________________________
Design and Airworthiness Requirements
for Service Aircraft
Section 3: Structure
_______________________________________
DEF STAN 00-970 PART 1/15
SECTION 3
REVISION NOTE
This standard is raised to Issue 15 to update its content.
HISTORICAL RECORD
This standard supersedes the following:
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 14 dated 30 January 2015
Design and Airworthiness Requirements for Service Aircraft
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 13 dated 11 July 2014
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 12 dated 10 January 2014
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 11 dated 05 July 2013
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 10 dated 07 January 2013
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 9 dated 06 July 2012
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 8 dated 31 October 2011
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 7 dated 31 January 2011
Defence Standard (Def Stan) 00-970 Part 1 Section 3 Issue 6 dated 05 February 2010
Defence Standard (Def Stan) 00-970 Part 1 Issue 5 dated January 2007
Defence Standard (Def Stan) 00-970 Part 1 Issue 4 dated January 2006
Defence Standard (Def Stan) 00-970 Issue 3 dated October 2003
Defence Standard (Def Stan) 00-970 Issue 2 dated 01 December 1999
Defence Standard (Def Stan) 00-970 Issue 1 dated 12 December 1983
2
DEF STAN 00-970 PART 1/15
SECTION 3
CONTENTS
Description Page No
Preface 4
SECTION 3 - STRUCTURE
3.0 Introduction 6
3.1 Static Strength and Deformation 7
3.2 Fatigue 21
3.3 Symmetric Manoeuvres 32
3.4 Asymmetric Manoeuvres 36
3.5 Gust Loads 44
3.6 High Lift Devices and Airbrakes 49
3.7 Pressure Cabins 57
3.8 Spinning and Spin Recovery 68
3.9 Control Systems 70
3.10 Active Control Systems 83
3.11 Strength for Ground Handling 108
3.12 Picketing 112
3.13 Lightning Protection 115
Tables 116
Figures 120
SUPPLEMENTS
Leaflets 0 to 45 127
3
DEF STAN 00-970 PART 1/15
SECTION 3
PREFACE
(a) This Part of the Defence Standard provides requirements for Airworthiness and
Design Certification for the design, development and testing of Fixed Wing Aircraft for UK
Military operation. The requirements stated herein shall be applied by the Ministry of Defence
(MOD) and the contractor as agreed and defined in the contract.
(b) This document has been produced on behalf of the Military Aviation Authority
Executive Board (MEB) by the Military Aviation Authority (MAA), MAA Technical Group, MOD
Abbey Wood.
(c) The appropriate Parts of this document are to be used, when called up in the
Contract, for all future designs, and whenever practicable for amendments to existing
designs. If any difficulty arises which prevents application of this document, DSA-MAA-Cert-
ADS1 shall be informed so that a remedy may be sought: e-mail: DSA-MAA-Cert-
ADSGroup@mod.uk
(d) Where the requirements of other Standards are considered applicable, the relevant
chapters and/or clauses are cross-referenced by this Part of the Defence Standard.
(f) Please address any enquiries regarding this standard, whether in relation to an
invitation to tender or to a contract in which it is incorporated, to the responsible technical or
supervising authority named in the invitation to tender or contract.
(g) Compliance with this Defence Standard shall not in itself relieve any person from any
legal obligations imposed upon them. Project Leaders are to ensure that equipment procured
from outside of the European Union (EU) meets or exceeds those legal requirements
mandated within the EU (See MAA 01 Chapter 1 and the RA1000 Series).
(h) This standard has been devised solely for the use of the Ministry of Defence (MOD)
and its contractors in the execution of contracts for the MOD. To the extent permitted by law,
the MOD hereby excludes all liability whatsoever and howsoever arising (including, but
without limitation, liability resulting from negligence) for any loss or damage however caused
when the standard is used for any other purpose.
4
DEF STAN 00-970 PART 1/15
SECTION 3
WARNING
(i) The Ministry of Defence (MOD), like its contractors, is subject to both United
Kingdom and European laws regarding Health and Safety at Work. Many Defence Standards
set out processes and procedures that could be injurious to health if adequate precautions
are not taken. Adherence to those processes and procedures in no way absolves users from
complying with legal requirements relating to Health and Safety at Work.
Note: Where a design to the requirements of this document may result in an adverse
environmental impact the MOD PTL shall be advised.
5
DEF STAN 00-970 PART 1/15
SECTION 3
SECTION 3 - STRUCTURE
3.0 INTRODUCTION
3.0.1 This section specifies the requirements for structural strength of the aircraft. Requirements are provided to cover the following aspects:
3.0.2 The section is intended to provide information similar to that contained in CS 25 Section C.
6
DEF STAN 00-970 PART 1/15
SECTION 3
3.1.1 This clause states the requirements for static strength and freedom from deformation which might adversely affect airworthiness,
operational capability or supportability. These requirements apply to all structures and mechanical components, including mechanical systems
such as undercarriages. Additionally, they shall apply to all materials and methods of construction. Compliance with these requirements shall be
demonstrated by the procedures recommended in the accompanying leaflets, unless prior agreement has been obtained from the PTL. This
clause contains information similar to EASA CS 25.301 – 25.307
7
DEF STAN 00-970 PART 1/15
SECTION 3
8
DEF STAN 00-970 PART 1/15
SECTION 3
9
DEF STAN 00-970 PART 1/15
SECTION 3
10
DEF STAN 00-970 PART 1/15
SECTION 3
11
DEF STAN 00-970 PART 1/15
SECTION 3
12
DEF STAN 00-970 PART 1/15
SECTION 3
13
DEF STAN 00-970 PART 1/15
SECTION 3
14
DEF STAN 00-970 PART 1/15
SECTION 3
15
DEF STAN 00-970 PART 1/15
SECTION 3
(a) shear, bending moment and To this end, the first objective of the tests is to
torque at the wing roots and at one or obtain measured values of the loads in flight and
more stations in the wing, on the ground in various manoeuvres. The
accuracy of the calculated loads on which the
(b) shear, bending moment and design of the aircraft was based should be
torque at the roots of the fin and tailplane, checked on the basis of these measurements.
Information on the magnitude and variation of
(c) shear and bending moment at loads to be applied in the fatigue tests should be
one or more fuselage stations, obtained. Any loading conditions that were not
foreseen in the design, but which are detected
(d) hinge moments on control during measurement, should be accounted for.
surfaces,
The number and position of the measurement
(e) load distributions over particular stations will depend on the project concerned, on
surfaces where special investigations (for the quantities being measured and on the
example, of buffet loads) is required, practical aspect of accessibility.
16
DEF STAN 00-970 PART 1/15
SECTION 3
(b) altitude,
17
DEF STAN 00-970 PART 1/15
SECTION 3
18
DEF STAN 00-970 PART 1/15
SECTION 3
19
DEF STAN 00-970 PART 1/15
SECTION 3
20
DEF STAN 00-970 PART 1/15
SECTION 3
3.2 FATIGUE
3.2.1 This information states the fatigue requirements that apply to all structures and mechanical components which affect safety, operational capability
and supportability (Grade A parts), including undercarriages and systems. In some circumstances special requirements apply to fibre-composite
components. These are summarised in Leaflet 40. Guidance on the additional considerations which apply when service lives are extended is provided in
Leaflet 39. Terms specific to this clause are highlighted in bold at first use. These terms are defined in Part 0.
21
DEF STAN 00-970 PART 1/15
SECTION 3
22
DEF STAN 00-970 PART 1/15
SECTION 3
23
DEF STAN 00-970 PART 1/15
SECTION 3
24
DEF STAN 00-970 PART 1/15
SECTION 3
25
DEF STAN 00-970 PART 1/15
SECTION 3
26
DEF STAN 00-970 PART 1/15
SECTION 3
27
DEF STAN 00-970 PART 1/15
SECTION 3
28
DEF STAN 00-970 PART 1/15
SECTION 3
29
DEF STAN 00-970 PART 1/15
SECTION 3
30
DEF STAN 00-970 PART 1/15
SECTION 3
31
DEF STAN 00-970 PART 1/15
SECTION 3
3.3.1 The requirements of this clause apply to all aeroplane types and to the aeroplane as a whole in each case. The requirements are expressed in
terms of a flight envelope, the co-ordinates of which are the forward speed and normal acceleration. Refer to Part 0 for definitions.
32
DEF STAN 00-970 PART 1/15
SECTION 3
33
DEF STAN 00-970 PART 1/15
SECTION 3
34
DEF STAN 00-970 PART 1/15
SECTION 3
35
DEF STAN 00-970 PART 1/15
SECTION 3
3.4.1 The requirements of this clause are applicable to all aeroplanes and to the aeroplane as a whole unless otherwise specified. The asymmetric flight-
conditions associated with side-slipping and rolling are considered, spinning being dealt with in Clause 3.8
36
DEF STAN 00-970 PART 1/15
SECTION 3
37
DEF STAN 00-970 PART 1/15
SECTION 3
38
DEF STAN 00-970 PART 1/15
SECTION 3
39
DEF STAN 00-970 PART 1/15
SECTION 3
40
DEF STAN 00-970 PART 1/15
SECTION 3
(2) where the motivators are driven (4) With the aeroplane in the steady
solely by the deflection of the pilot's stick rate of roll condition referred to in (2)
or wheel, that corresponding to a control above, the loads which arise when the roll
force of 267N (60 lbf) for a stick control of motivator(s) are moved to a deflected
222N (50 lbf) applied to the rim of a wheel position of equal magnitude of that in (1)
of diameter Dm (Din), resulting in a couple above, but of opposite direction.
of magnitude 222D (Nm) or 50D (lbf in)
COMBINED ROLLING AND PITCHING
3.4.14 The loads arising from the combined For the definition of normal acceleration see
application of the pitch and roll motivators shall be Clause 3.3.3 and Part 0.
considered for all speeds up to the design diving
speed, VD, and all altitudes up to the maximum
operating altitudes for the following conditions:
41
DEF STAN 00-970 PART 1/15
SECTION 3
42
DEF STAN 00-970 PART 1/15
SECTION 3
43
DEF STAN 00-970 PART 1/15
SECTION 3
3.5.1 The requirements of this clause are applicable to all aeroplanes and to the strength of the complete structure, when the aeroplane encounters gusts
normal to the flight path both in the plane of symmetry (vertical gusts) and perpendicular to the plane of symmetry (lateral gusts).
44
DEF STAN 00-970 PART 1/15
SECTION 3
45
DEF STAN 00-970 PART 1/15
SECTION 3
46
DEF STAN 00-970 PART 1/15
SECTION 3
47
DEF STAN 00-970 PART 1/15
SECTION 3
48
DEF STAN 00-970 PART 1/15
SECTION 3
3.6.1 The requirements of this clause are applicable to all aeroplanes and to the strength of high lift devices and airbrakes, their operating mechanisms
and all parts of the complete airframe affected by the load cases specified. Attention is drawn to the fatigue and damage tolerance requirements of Clause
3.2 and to the aeroelasticity requirements of Part 1, Section 4, Clause 4.8
49
DEF STAN 00-970 PART 1/15
SECTION 3
50
DEF STAN 00-970 PART 1/15
SECTION 3
(a) For the high lift devices in the take- (a) If an automatic positioning or load
off position, V TO which shall be the greater limiting system is used the speeds and
of, 1.15 times the equivalent airspeed during corresponding positions programmed or allowed
take-off at the maximum design mass before by the system may be used.
the devices can be fully retracted, or 1.6
times the equivalent stalling speed with the (b) When high lift devices are used en-route
flaps in the take-off position, zero engine or in combat condition appropriate design
power and maximum design mass. speeds shall be established which are
equivalent to speeds V A , V C , V D , V G and V H as
(b) For the high lift devices in the specified in Part 0 except that allowance may be
intermediate position, V BL , which shall be the made for the position of the high lift devices.
greater of, the equivalent airspeed attained in
a baulked landing at maximum design
landing mass before the devices can be
retracted to the intermediate position, or 1.8
times the equivalent stalling speed with the
devices in the intermediate position, zero
engine power and maximum design landing
mass.
51
DEF STAN 00-970 PART 1/15
SECTION 3
52
DEF STAN 00-970 PART 1/15
SECTION 3
53
DEF STAN 00-970 PART 1/15
SECTION 3
54
DEF STAN 00-970 PART 1/15
SECTION 3
55
DEF STAN 00-970 PART 1/15
SECTION 3
56
DEF STAN 00-970 PART 1/15
SECTION 3
3.7.1 The requirements of this clause apply to all pressurised cockpits or cabins designed to maintain an internal pressure in excess of the external
atmospheric pressure.
57
DEF STAN 00-970 PART 1/15
SECTION 3
58
DEF STAN 00-970 PART 1/15
SECTION 3
59
DEF STAN 00-970 PART 1/15
SECTION 3
60
DEF STAN 00-970 PART 1/15
SECTION 3
61
DEF STAN 00-970 PART 1/15
SECTION 3
62
DEF STAN 00-970 PART 1/15
SECTION 3
63
DEF STAN 00-970 PART 1/15
SECTION 3
64
DEF STAN 00-970 PART 1/15
SECTION 3
65
DEF STAN 00-970 PART 1/15
SECTION 3
66
DEF STAN 00-970 PART 1/15
SECTION 3
67
DEF STAN 00-970 PART 1/15
SECTION 3
3.8.1 This clause states design and strength requirements for spinning, including entry to and recovery from the steady condition. It is applicable only to
those aeroplanes where recovery from a post stall gyration or spin is called for in the Aeroplane Specification and it applies whether the entry to the post
stall gyration or spin is intentional or inadvertent If spin prevention devices having sufficiently low probability of failure are used the spinning requirements
which then apply shall be agreed with the PTL. Part 1, Section 2, Clauses 2.24.11 to 2.24.13, specifies the aeroplane handling requirements for post stall
gyration and spins. Guidance on Design Criteria for Spin Resistance and Spin Recovery is given in Leaflet 18.
68
DEF STAN 00-970 PART 1/15
SECTION 3
69
DEF STAN 00-970 PART 1/15
SECTION 3
3.9.1 The requirements of this clause apply to all control systems, including engine controls, in which the position of the motivators is determined solely by
the pilot through a direct connection from the inceptors. It also applies to those parts of the flight control system which are not covered by the requirements
of Clause 3.10 for active control systems, or of Part 1, Section 6, Clause 6.5 for autopilot systems.
The requirements relate mainly to mechanical components, but powered controls, power operated trimmers and devices such as Q-feel and variable stops
are also included. Autostabilisers, auto-trim and similar devices which can move the motivators independently of inceptor inputs are not included, but are
covered by Clauses 3.10 and Part 1, Section 6, Clause 6.5
The requirements shall be considered in conjunction with other relevant clauses including:
70
DEF STAN 00-970 PART 1/15
SECTION 3
71
DEF STAN 00-970 PART 1/15
SECTION 3
(b) Pilots acting in opposition. Note: This case provides a design condition for
72
DEF STAN 00-970 PART 1/15
SECTION 3
73
DEF STAN 00-970 PART 1/15
SECTION 3
74
DEF STAN 00-970 PART 1/15
SECTION 3
75
DEF STAN 00-970 PART 1/15
SECTION 3
76
DEF STAN 00-970 PART 1/15
SECTION 3
77
DEF STAN 00-970 PART 1/15
SECTION 3
78
DEF STAN 00-970 PART 1/15
SECTION 3
79
DEF STAN 00-970 PART 1/15
SECTION 3
80
DEF STAN 00-970 PART 1/15
SECTION 3
81
DEF STAN 00-970 PART 1/15
SECTION 3
82
DEF STAN 00-970 PART 1/15
SECTION 3
3.10.1 This information contains the requirements relating to flight with active controls. Active Control Systems (ACS) are systems in which commands to
the motivators are computed from sensor inputs with or without inceptor inputs. Thus, although the term is normally applied to systems which must operate
continuously and without whose correct operation safe flight cannot be maintained, it also includes all forms of autostabiliser and other systems which can
command motivator position independently of inceptor units. Definitions of terms are given in Part 0.
The requirements of this clause are concerned primarily with the most critical systems, those which operate continuously (full-time). The extent to which
they apply to other flight control systems will be stated in the Aeroplane Specification. All flight control systems, including ACS, must satisfy the
requirements of Clause 3.9 Requirements for autopilots are stated in Part 1, Section 6, Clause 6.5 For piloting aspects, handling qualities are dealt with in
Section 2, and, in particular, the reaction of the pilot to mode changes and failures, in Part 1, Section 4, Clause 4.10
ACS may be provided for purposes other than flight control. In all cases the requirements of this clause will apply if maintenance of safe flight is involved.
Fig 4 shows the fundamental features which constitute an ACS. The Categories A and B outlined define those functions which are respectively purely
automatic and those which involve manual inputs. Table 4 lists representative control functions in Categories A and B. See also Leaflet 25.
83
DEF STAN 00-970 PART 1/15
SECTION 3
84
DEF STAN 00-970 PART 1/15
SECTION 3
85
DEF STAN 00-970 PART 1/15
SECTION 3
86
DEF STAN 00-970 PART 1/15
SECTION 3
87
DEF STAN 00-970 PART 1/15
SECTION 3
88
DEF STAN 00-970 PART 1/15
SECTION 3
89
DEF STAN 00-970 PART 1/15
SECTION 3
90
DEF STAN 00-970 PART 1/15
SECTION 3
91
DEF STAN 00-970 PART 1/15
SECTION 3
92
DEF STAN 00-970 PART 1/15
SECTION 3
93
DEF STAN 00-970 PART 1/15
SECTION 3
94
DEF STAN 00-970 PART 1/15
SECTION 3
95
DEF STAN 00-970 PART 1/15
SECTION 3
96
DEF STAN 00-970 PART 1/15
SECTION 3
97
DEF STAN 00-970 PART 1/15
SECTION 3
98
DEF STAN 00-970 PART 1/15
SECTION 3
99
DEF STAN 00-970 PART 1/15
SECTION 3
100
DEF STAN 00-970 PART 1/15
SECTION 3
101
DEF STAN 00-970 PART 1/15
SECTION 3
102
DEF STAN 00-970 PART 1/15
SECTION 3
103
DEF STAN 00-970 PART 1/15
SECTION 3
104
DEF STAN 00-970 PART 1/15
SECTION 3
105
DEF STAN 00-970 PART 1/15
SECTION 3
106
DEF STAN 00-970 PART 1/15
SECTION 3
107
DEF STAN 00-970 PART 1/15
SECTION 3
3.11.1 The requirements should be read in conjunction with the design, operational and standardisation requirements for transport, handling and storage
given in Part 1, Section 4, Clauses 4.4.8 to 4.4.24
108
DEF STAN 00-970 PART 1/15
SECTION 3
109
DEF STAN 00-970 PART 1/15
SECTION 3
110
DEF STAN 00-970 PART 1/15
SECTION 3
111
DEF STAN 00-970 PART 1/15
SECTION 3
3.12 PICKETING
3.12.1 The requirements aim to ensure that provision is made on all aeroplanes for picketing.
112
DEF STAN 00-970 PART 1/15
SECTION 3
113
DEF STAN 00-970 PART 1/15
SECTION 3
114
DEF STAN 00-970 PART 1/15
SECTION 3
3.13.1 All information currently held within Part 1, Section 4, Clauses 4.27.23 to 4.27.40 and Def Stan 59-113.
115
DEF STAN 00-970 PART 1/15
SECTION 3
Estimated Sample size of coupon tests n, used to estimate population cv in observed failure mode
population cv
from n coupon 15 30 100 or more
tests
i.e. Number of element tests N
1* 2 3 5 10 1* 2 3 5 10 1* 2 3 5 10
Characteristic 3% or less 1.14 1.11 1.10 1.09 1.08 1.12 1.10 1.09 1.08 1.07 1.10 1.09 1.08 1.07 1.06
cv in observed 5% 1.26 1.20 1.18 1.16 1.14 1.20 1.17 1.16 1.14 1.13 1.18 1.15 1.14 1.12 1.11
failure mode 7.5% 1.37 1.31 1.28 1.26 1.23 1.32 1.27 1.24 1.22 1.20 1.28 1.23 1.21 1.19 1.17
10% 1.53 1.44 1.40 1.36 1.32 1.45 1.37 1.34 1.31 1.27 1.39 1.32 1.29 1.27 1.24
15% 1.89 1.74 1.68 1.61 1.55 1.75 1.62 1.57 1.51 1.46 1.63 1.53 1.48 1.44 1.39
TABLE 1
FACTORS BY WHICH THE MEAN STRENGTH OF DETAILS OR ELEMENTS MUST BE REDUCED TO OBTAIN A 'B' ALLOWABLE VALUE:
WHERE APPROPRIATE ENVIRONMENTAL DEGRADATION MUST BE INCLUDED IN THE TESTS:
THESE FACTORS APPLY TO ALL GRADE A DETAILS AND TO ALL MATERIALS
*Normally at least 2 tests must be done so that any gross inconstancy is likely to be revealed.
116
DEF STAN 00-970 PART 1/15
SECTION 3
Estimated cv for type of Number of tests upon which the mean value
complete is estimated
item under test* 1 2
5% or less 1.00 1.00
8% 1.10 1.05
10% 1.15 1.10
15% 1.33 1.25
TABLE - 2
TEST FACTORS BY WHICH THE MEASURED STRENGTH OF COMPLETE STRUCTURES FOR COMPONENTS MUST EXCEED THE DESIGN ULTIMATE LOAD IN
ORDER TO ALLOW FOR THE INFLUENCE OF VARIABILITY - WHERE APPROPRIATE A SEPARATE ALLOWANCE MUST BE MADE FOR ENVIRONMENTAL
DEGRADATION
*To be agreed with the PTL. It is customary to use a figure of 5% for conventional metal structures.
Wherever practical at least 2 tests must be done so that any gross inconsistency is likely to be revealed.
117
DEF STAN 00-970 PART 1/15
SECTION 3
TABLE 3
118
DEF STAN 00-970 PART 1/15
SECTION 3
TABLE 4
ACTIVE CONTROLS
Categories of function
119
DEF STAN 00-970 PART 1/15
SECTION 3
VALUES SPECIFIED FOR PARTICULAR AEROPLANE TYPES (see Part 0 Annex F for definitions).
120
DEF STAN 00-970 PART 1/15
SECTION 3
DEPENDENT VALUES
(Unless otherwise specified in the Aeroplane Specification).
VA is the manoeuvring speed that is that minimum flying speed at which n1 can be achieved.
VH shall be the maximum speed attainable in level flight with the aeroplane flying at the basic design mass, with no external stores at the maximum
continuous cruise engine condition.
n2 = 1-0.3n1
n3 = -0.6(n1 -1)
NOTE:
(a) Aeroplanes where speed V H is equivalent to a Mach number of less than unity, other than weapon system aeroplanes; V G shall be either the speed
determined by the intersection of the line representing the maximum lift coefficient and the 20 m/s (66 ft/sec) gust line on the n-V diagram or V S n G
where V S is the stalling speed in level flight and n G is the maximum load factor resulting from a 15.2 m/s (50 ft/sec) gust at speed V H .
(b) Weapon system aeroplanes and others where speed V H is equivalent to a Mach number of unity or greater; V G shall be determined by mission
requirements, the permissibility of reducing speed and the slow-down speeds attainable and shall be agreed with the Project Team Leader, but need not be
greater than V H .
121
DEF STAN 00-970 PART 1/15
SECTION 3
122
DEF STAN 00-970 PART 1/15
SECTION 3
FIGURE 3
123
DEF STAN 00-970 PART 1/15
SECTION 3
Categories:
124
DEF STAN 00-970 PART 1/15
SECTION 3
125
DEF STAN 00-970 PART 1/15
SECTION 3
126
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 0
REFERENCES
Each set of references within this leaflet is divided according to the reference number within
Section 3.
No additional references.
ARC Reports
R and M 3327 Simplified loading formulae for pull out manoeuvres of tailed aircraft
No additional references.
ARC Reports
R & M 2308 An analysis of the lift slope of aerofoils of small aspect ratio, including
fins, with design charts for aerofoils and control surfaces
CP324 The variation of gust frequency with gust velocity and altitude
NACA Reports
Page 1 of 5
DEF STAN 00-970 PART 1/15
SECTION 3
No additional references.
MOD Specifications
Defence Standards
RAE Reports
ARC Reports
Page 2 of 5
DEF STAN 00-970 PART 1/15
SECTION 3
ARC Reports
AGARD Proceedings
Other Papers
RAE Reports
Aero 1972 Spring tab controls - Notes on development to date, with special
reference to design aspects
Aero 2232 Papers and proceedings of a meeting to discuss powered flying
controls
Mech.Eng.12 Transitional friction effects in powered controls with particular
reference to hydraulic jacks
Page 3 of 5
DEF STAN 00-970 PART 1/15
SECTION 3
RAE Specifications
ARC Papers
Page 4 of 5
DEF STAN 00-970 PART 1/15
SECTION 3
RAeS Journal
Interim DEF STAN 00-31(Obsolescent) The development of safety critical software for
airborne systems.
No additional references.
3.11 PICKETING
No additional references.
3.12 FATIGUE
No additional references.
No additional references.
Page 5 of 5
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 01
UNDERLYING PRINCIPLES
1 INTRODUCTION
1.1 The purpose of this Leaflet is to outline the basic principles underlying the strength and
deformation requirements in this publication and, in particular, to show why these requirements are
sometimes expressed in terms of proof loads, sometimes ultimate load, and sometimes both.
2.1 There are several causes of uncertainty that must be taken into account in designing
aircraft to operate within the required envelope.
2.2 One of these is the variability in the strength of materials and structural details. This is
accounted for in three separate ways.
2.2.1 First, the designer is required to use 'B' allowable values of stress in design (see
Part 1, Section 4, Clauses 4.1 and 4.5) because they are founded on a standardized test
procedure and give a uniform statistical assurance of the strength of materials. Exceptionally,
however, Specification 'S' values may be used where no 'B' value is available.
2.2.2 Second, any test of a structural detail or structural element is subject to a test factor,
as required by Table 1 of Clause 3.1. These are related to the measured variability in the
strength of the materials from which the details or elements are made; data on the variability
of metallic materials can be found in ESDU 00932 Metallic Materials Handbook. The test
factors are calculated to give assurance, with 95% confidence, that 90% of similar items will
reach or exceed the required strength. It is emphasised that separate allowance must be
made for any uncertainties arising from differences between elements and the structural
details they represent.
2.2.3 Finally, the designer must make separate allowance for environmental degradation
and demonstrate that the allowable values (or 'S' values) of ultimate stress are not exceeded
under the DUL. The most appropriate allowance for degradation can be obtained by
conducting tests on structural elements under the most adverse combinations of moisture
content and temperature to which they will be exposed in service (see Part 1, Section 3,
Clauses 3.1.23 – 27). It should be noted that cold/dry conditions may also have an adverse
effect on the strength of the material, and should therefore be considered in addition to
'hot/wet' conditions. Furthermore, the effects of exposure to ultra-violet light and other non-
ionising radiation should be considered.
3.1 To perform a check stress is a condition of compliance with the static strength
requirements. A reference analysis shall be done, using definitive loads and masses, together with
a structural analysis model, all fully representative of the final design. The check stress should
ideally be completed by an independent team.
Page 1 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
4.1 Each aircraft is subject to an enormous variety of combinations of loads and other
conditions, and it is obviously impracticable and unnecessary to consider them all. The most
adverse of these are identified in the form of design cases, which must be used in demonstrating
compliance with the requirements.
4.2 In the majority of design cases, the severity of the loads that may arise in service is not
limited naturally, and limiting conditions have to be chosen for design purposes. A classic example
is provided by the symmetric flight case. For aircraft without an active control system, there are
usually no means (other than the pilot) to prevent the exceedance of either the speeds or the
normal accelerations that are permissible for an aircraft in symmetrical flight. In choosing the
limiting conditions for design, use is made of past experience of the speeds and accelerations
needed for the operational duty of the type, of the predicted flight characteristics and performance
of the aircraft, and of the way structure mass is likely to depend on the severity of the chosen
conditions.
4.3 Every aircraft in service is subject to flying limitations that prohibit its use beyond the
limiting conditions chosen for design (see Part 1, Section 4, Clause 4.9).
4.4 In some situations, limit load cases are defined-such that it is impossible (extremely
improbable, or so unlikely as not to occur during the entire operational life of all aircraft of one type)
or extremely remote (unlikely to occur during the entire operational life of all aircraft of one type,
but nonetheless still possible) for-the loads to be exceeded. These cases are normally chosen
either to reduce the number of combinations to be considered (ease of analysis), where there are
limiting mechanisms, or where the loads are difficult to define rationally.
5.1 Though the design ultimate strength defines the margin of strength that must be provided, it
is not sufficient on its own to ensure that there is also a margin of airworthiness. This margin of
airworthiness is defined by the design proof requirement.
5.2 The proof requirement defines the minimum conditions up to which the airworthiness and
serviceability of the aircraft can be depended upon. The designer should ensure, by analysis and
tests as appropriate, that loads up to the design proof load will neither impair safety nor
functionality nor necessitate inspection or rectification other than to predetermined and defined
schedules acceptable to the operator. Thus, under the design proof load and after its removal
there should be no loosening or pulling of fasteners, structural deformation which might cause
unacceptable redistribution of aerodynamic loads, jamming or undue slackness of controls, or any
other sign of structural distress which might cause the operator to have doubts about the
serviceability of the aircraft. Beyond the proof strength, catastrophe could result from structurally
induced causes such as jamming of controls or unstable aerodynamic loading.
5.3 There are several causes of uncertainty that must be taken into account in designing the
aircraft to withstand the pertinent limiting conditions. For example, the methods used in the
evaluation of the loads and in the stress analysis are inexact.
Page 2 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
5.4 Furthermore, when the aircraft enters service, the most severe conditions under which it is
actually used may differ from the design cases. For all these reasons the aircraft must be designed
to have, in each design case, an ultimate strength that is more than would be needed just to
sustain the limit load.
6 REDUCTIONS IN FACTORS
6.1 In some circumstances, load factors may be reduced, where it can be shown that
exceedances of limiting conditions are impossible (extremely improbable) or extremely remote, as
defined in paragraph 4.4. In these cases there is no proof requirement above limit load.
6.2 A reduced ultimate factor of 1.4 may be used in conjunction with these cases, provided it
can be shown that all underlying rational cases, multiplied by the conventional ultimate factor of
1.5, are covered. For example, this may apply to aircraft equipped with active flight control
systems, where it can be demonstrated that such systems are load-limiting and sufficiently reliable.
Other examples may include load limiting mechanisms, such as pressure relief valves, provided
that the upper tolerance is used.
6.3 To justify the use of a reduced ultimate load factor deterministic methods have traditionally
been used and the amount of the reduction has been based on engineering judgement. However,
such an approach does not rigorously account for systems effectiveness (in limiting the loads) and
reliability.
6.4 In order to take a probabilistic approach, it is necessary to define the structural factors of
safety as a function of systems reliability. Such an approach has been proposed for inclusion in
JAR 25 (NPA 25C-199, dated 24 April 1996 incorporated into CS-25 amendment 1), and a similar
methodology would be considered acceptable.
7.1 The traditional, semi-deterministic approach involves designing the structure to withstand a
factored maximum load to satisfy the required limit load criterion. In some cases, however, it may
be acceptable to use a probabilistic approach to estimate the required limit or proof loads having
an acceptably low probability of occurrence. In essence, this is achieved by modelling the variables
determining the loading envelope and then deriving a suitable value for the limit (or proof) load,
based on the distribution.
7.2.2 There may be different consequences of exceedance of the chosen limit load, with
consequences ranging from the catastrophic to the merely inconvenient, and this will be
reflected in the probability of failure which is considered acceptable. More information on
determining the level of risk that will be acceptable is given in Def Stan 00-56.
Page 3 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
7.2.3 It should be remembered that the calculated probability of occurrence takes into
account stochastic variability only. It does not take into account the possibility of human
errors on the part of the user; in practice, failures due to non-quantifiable sources of
unreliability may well predominate.
7.2.4 It should also be noted that, as with deterministic analyses, changes in usage may
change the loading envelope. Thus any limit load definition, based on assumptions from a
probabilistic analysis, may be invalidated by changes in usage. Therefore, it is essential that
usage data are compared with any assumptions made, as part of the mandatory aircraft
monitoring programmes.
7.3 UNCERTAINTY
7.3.2 Firstly, there is the unavoidable uncertainty associated with the values of quantities
which exhibit scatter. It is this inherent stochastic variability which provides the motivation for
the probabilistic design method, and which must be modelled when using this method. In the
absence of such variability, the probability of failure associated with a particular design of
structure would always be either 0 or 1.
7.3.3 Secondly, there is the uncertainty that arises from incomplete knowledge of the
probability distributions of variables which affect structural integrity. Given complete
knowledge of the form and parameters of each such distribution (and the deterministic
relationships between them), the loading envelope could, in principle, be accurately
determined. However, uncertainty in the designer's knowledge of these distributions, due to
lack of data, leads to uncertainty in the calculated probability of occurrence. It is essential
that this be taken into account: this can be done by calculating confidence limits. Determining
loading distributions to the accuracy required to have sufficient confidence in what may occur
in at the extremes of the distribution is no easier than defining strength distributions.
7.3.4 Finally, only as complex an analysis and test programme as will have a significant
effect on the resulting design should be undertaken. The degree to which the design of a
component can be optimized should depend on the overall contribution which it makes to the
reliability, performance and cost of a structure.
7.4.1 The quantities whose inherent variability may have a significant effect on the loads
must be identified and modelled by probability distributions. These quantities might include
coefficients, of friction, temperature, manoeuvre acceleration, gust loads, internal pressure,
vibration and shock loads.
Page 4 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
7.4.2 Ideally, the applied load (or stress) envelope would be modelled for each design
case, but in many cases this will not be feasible, especially at the initial design phase. In
these circumstances, a more indirect approach is necessary, involving the probabilistic
modelling of quantities on which loads depend. In such cases, an analytical solution it
unlikely to be obtainable and a simulation method should be used to sample from the
appropriate probability distributions.
7.4.3 It is the variability in the maximum load relevant to each critical design case which
should be modelled, not the variability of the loading over the life of the structure.
7.4.4 A sensitivity analysis should be carried out at an early stage in the project to
identify, on the one hand, the variables whose scatter has no significant effect on the failure
probability, and which can therefore be treated as constants and, on the other hand, the
variables whose scatter is highly significant. Effort can then be concentrated on more
accurate modelling of the latter.
7.5.1 The choice of a suitable form of probability distribution to model each of the
significant variables may involve:
- physical considerations
- previous experience
- examination of relevant test data.
7.5.3 Alternatively, rather than estimating the form and parameters of the probability
distribution of a variable as described above, a non-parametric approach may be used to
analyse available data. This involves making no assumptions about the form of the
distribution, but instead constructing an empirical cumulative distribution from the data
themselves. This approach can be useful where there is a lack of previous relevant
experience to suggest a suitable form of probability distribution, and when the data do not
obviously come from one of the common-distributions. Its applicability is limited, however,
since it is frequently necessary to extrapolate- into the tails of a distribution from data which
all lie within a few standard deviations of the mean. In order to do this, it is necessary to have
knowledge of the distributional form (or to assume a form which will result in an over-
estimate of failure probability rather than an underestimate).
8 SPECIAL CASES
8.1 In most design cases both strength and deformation are regarded as important. It seldom
happens that both requirements can be satisfied exactly; usually a margin above the standard
required by one requirement must be provided in order to comply with the other. Nevertheless
when both requirements are imposed, the designer is expected to demonstrate compliance with
them both.
Page 5 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
8.2 In some exceptional design cases, deformation is all that matters and a design ultimate
requirement is superfluous. The principal example is that of ejection seats and their surrounding
structure. Under the worst possible combinations of flight loads and ejection gun thrust, the seats
and the airframe must not distort so much that they prevent an ejection, or cause a seat to eject
too slowly or in the wrong direction, or make the ejection of one seat interfere with the ejection of
another.
8.3 In a few other cases, strength is critical but deformation is not important, and a design
ultimate requirement is imposed without a design proof requirement. An example is that of seats
and harnesses at crash stations, and their supporting structures. In a crash landing that would do
serious damage to the aircraft, some distortion of the seat and harness attachments would be
trivial, (and may be beneficial in absorbing the energy of impact), but the crew must remain safely
restrained within reasonable physiological limits. Accordingly an ultimate factor of 1.0 is required in
the crash landing case but there is no proof requirement (see Part 1, Section 4, Clause 4.22).
8.4 There are a few other special cases in which the respective values of the proof and ultimate
factors are lower than the usual 1.125 and 1.5. These cases represent rare events such as
malfunctions of systems and operation of systems under exceptional conditions. Provision of the
full strength and stiffness that would be necessary to achieve the usual factors in these cases
would add mass to the aircraft that would penalise it in normal operations. In emergencies that are
infrequent and of short duration, and in which the aircraft may be at additional risk for other
reasons, the higher risks to the aircraft and its crew that are associated with the lower factors are
regarded as acceptable. Examples are malfunctions of automatic flight control systems and the
use of anti-spin parachutes at exceptionally high speeds (Part 13, Section 3, Clause 3.7).
8.5 Finally there are several cases in which both the proof and ultimate factors are higher than
the usual values. They represent the normal operations of systems and installations in which the
loads are particularly difficult to estimate. The difficulties are caused principally by transient
pressures in hydraulic, pneumatic and fuel systems, by variability in the performance of pressure
relief devices, and by variability in the loading on air brakes, flaps and undercarriages. Since these
cases represent events that happen frequently, the greater risks associated with the greater
uncertainty in estimation of the loads must be compensated by higher factors. Examples are flap
systems (Clause 3.6), undercarriage retraction and lowering (Part 1, Section 4, Clause 4.11) and
refuelling and defuelling systems (Part 1, Section 5, Clause 5.2).
8.6 The foregoing paragraphs indicate the diverse nature of the strength requirements and the
reasons why emphasis is sometimes placed on a proof requirement, sometimes on an ultimate
strength requirement, and why often both are specified. It is clear that the same factors, and the
same ratio of proof factor to ultimate factor, cannot be used in all cases without serious loss in
structural integrity. This is a consideration that far outweighs the convenience that uniform factors
might bring.
Page 6 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 02
1 INTRODUCTION
1.1 The test procedure adopted should provide the maximum practicable information on
the structure. In particular, the tests should be conducted with the following aims:
a. to verify the structural analysis by means of strain, deflection and other appropriate
measurements under all test conditions and thereby demonstrate, for each Grade A
structural detail, that the allowable values of stress or strain are not exceeded at the
design ultimate load.
b. to establish, for selected design cases, the highest load at which the proof
requirement is satisfied.
c. to demonstrate, in the critical design case(s) established by earlier testing and
analysis, a strength that is sufficient to show, as far as practicable, that all Grade A
structural details have been identified.
3 COUPON TESTS
3.1 Coupon tests can be used to determine the materials allowables on a ‘B’ basis for
use in design. Since the ‘B’ allowable is a material property, tests must be undertaken for all
potential failure modes. Statistical methods for the derivation of ‘B’ basis values are given in
Reference 1.
Page 1 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
3.2 In determining the ‘B’ value, allowance must be made for the degradation by
moisture absorption, exposure to extreme temperatures, ultra-violet (UV) light and non-
ionising radiation. The scope, range and combination of environmental parameters
considered should be appropriate to the material and its application. Examples may include,
but are not limited to, the effects of extremes of temperature on material properties such as
strength, stiffness and fracture toughness.
3.3 Tests should include, but not be limited to:
tension,
compression,
shear,
bearing,
holes and notches,
Young’s Modulus.
3.4 When considering the scope of the coupon test programme necessary to obtain
materials design data (materials qualification issues are addressed in Part 1, Section 4,
Clause 4.1), limitations of the applicability of these data to the structural level should be
considered. Such limitations may include:
scale effects,
differences in variability and properties between the as-received material and
the as-manufactured component.
Page 2 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
and additionally, for example, for fibre composite structures, this may include:
tension angles,
shear angles,
flange / skin attachments,
impact damage, e.g. bird strike, tool drop, etc.,
rain / hail erosion.
4.4 The number of tests necessary at this level will be dependent on past experience.
Where a new design is based on existing structures, or confidence in the analytical methods
to be used is already established, the number of tests may be reduced.
5 SUB-COMPONENT TESTS
5.1 Sub-component tests should be representative of large portions of the aircraft
structure and should be used to validate the analysis of structural features and confirm
design allowables. The tests represent a significant part of the test programme and have the
following objectives with respect to the structural qualification of the aircraft:
to demonstrate that data derived from coupon and element tests can be
successfully used to predict the performance of the structure;
to verify analysis techniques (including FEM where applicable);
to obtain load/strain relationships for the sub-components. In the case of
composites, tests should be carried out under both room temperature / as
received (RT/AR) and degraded conditions for read-across to the major static
tests (see note below).
to substantiate the fatigue life and residual strength of large scale critical features
(see Clause 3.2).
Note: Where major tests involving composites are undertaken under RT/AR conditions,
the results of the sub-component tests under degraded conditions are required to obtain read
across from the major test results to the actual service environment.
5.3 It may also be necessary to carry out comparability tests (for components
manufactured at two or more locations) at the sub-component level, for components that
undergo a complex manufacturing route, for example mouldings, forgings and
superplastically-formed, diffusion bonded structures.
Page 3 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
6 COMPONENT TESTS
6.1 It is a requirement to carry out at least one test of the entire structure: it is
permissible, and may be more manageable, to test the structure in a series of component
tests. The loading on these test articles should be representative of the operating conditions
acting upon the component. The objective of the major test(s) is to substantiate the structural
integrity of all primary structure, including all major load paths and to validate the structural
analysis in terms of the load paths, strains and deflections.
6.2 The test articles shall be fitted with such control runs and systems as are necessary
to demonstrate compliance with the proof requirement.
6.3 Consideration should be given to the instrumentation of the test article, especially to
the placing of strain gauges (and other instrumentation) and the placement of loading fixtures
and reaction points to ensure that sufficient data are obtained to demonstrate compliance
with the requirements.
6.4 The following test procedure is recommended:
a. Load in increments up to the design proof load in each critical loading case. If
measured flight loads data are available, these should be used. Check that there is no
structural deformation which might cause unacceptable redistribution of aerodynamic
loads. After removal of the load, examine and report on the condition of the specimen,
noting with photographs, all structural distress, which, if it occurred on an aeroplane in
service, might cause it to be declared unserviceable or not airworthy (for example, skin
distortion, loose rivets, serious permanent deformation, etc.).
b. Consideration should be given to loading the test article, in each test case, to an
intermediate level between design proof load and design ultimate load. This serves to
maximise data obtained in the event of premature failure.
c. Demonstrate an appropriate test factor from Table 2 within Section 3 on the design
ultimate load for static tests, or the required residual strength. If, during these tests,
serious damage to the specimen occurs unexpectedly or appears imminent, a decision
should be made whether to repair the specimen or to proceed to (d).
d. Repeat stage (c) for each of the other selected loading cases.
6.5 For mixed metal and composite structures, the use of test factors to account for
variability and environmental effects can cause problems. Loading beyond the design
ultimate load may over-stress components with low static reserve factors. Hence,
consideration may be given to the use of dummy metallic structures and where these are
used, it will be necessary to provide alternative test substantiation for the real parts.
6,6 Where enhanced loading is not used, the only practicable method of ensuring that
the structure is capable of carrying ultimate load is by comparing the strain gauge results
obtained from the major static test with the results obtained from tests on smaller boxes and
elements containing representative defects and tested under the correct environmental
conditions. These results are used to validate the analysis.
Page 4 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
7 FUNCTIONALITY TESTS
7.1 Testing is done to validate analysis of the functionality of major control and
mechanical systems up to the proof load. The objective of such testing is to show that these
systems have full range and freedom of movement under critical loading conditions.
7.2 Such tests may be done on the static test article or it may be necessary, or
expedient, to use separate test installations.
8 DEFINITIONS
8.1 B basis data – a (one sided) lower 95 per cent confidence limit for the value above
which 90 per cent of the population lies. This means that there is a 95 per cent probability
that at least 90 per cent of the material released will exceed the B value.
9 REFERENCES
Metallic Materials Data Handbook ESDU 00932, ESDU International plc, ISSN 0261-2402.
Page 5 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
Details, e.g.
Skin/stringer panels
Major joints
Spar or rib box test Development Tests
Proof of concept
Elements, e.g.
Bolted joint elements Design data generation
Compression panels
Impact damage toleranc e
Coupons, e.g.
Materials & processes
Design data generation
Environment
Page 6 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
tension
compression
shear
Coupons holes & notches
defects
Obtain design ‘B’ allowables Young’s modulus
environmental effects
joints
Details & elements compression panels
shear panels
Detail stress analysis skin cut-outs & reinforcements
flange to skin attachments
Components
Page 7 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 03
1 INTRODUCTION
1.1 In Clause 3.1, it was stipulated that structural qualification should be achieved by
means of a staged, integrated programme of structural analysis, structural tests and flight
load measurements. The staged approach allows the use of clearance factors to provide an
aircraft release at intermediate qualification levels, thereby enabling preliminary test flights.
1.2 As each stage of the qualification programme is successfully completed, these
factors are then individually removed to allow progression to full clearance; that is, at the
100% strength envelope capability.
1.3 This approach may also be used to provide a staged release to service.
2 CLEARANCE FACTORS
2.1 Clearance factors provide a means of restricting the flight envelope, such that the
loads experienced by the structure are appropriately limited for the stage reached in the
qualification programme. These factors apply to the appropriate minimum reserve factors
shown by analysis. The flight envelope is then restricted to respect this allowable load
envelope. It should be noted that the same procedure can be used for flight and ground
loading actions, although this Leaflet refers mainly to flight loads.
2.2 The following defines the range of clearance elements that should be considered to
provide this staged approach:
2.2.1 The use of unconventional structures. Aircraft projects have seen an
increased use of new and complex and/or unconventional materials and manufacturing
processes, for example carbon fibre composites and superplastically formed, diffusion
bonded titanium. Until the properties of these unconventional structures have been fully
explored, in the full aircraft test programme, a restriction should be applied for early
clearances. A factor of 0.8 – 1.0 should be agreed for this requirement, dependant on
the extent and nature of unconventional structure employed and the test evidence
available.
2.2.2 Absence of Ultimate Load structural tests. Full aircraft static tests are
normally undertaken using a staged approach, initially exploring main load cases to
limit load conditions, before proceeding to the ultimate cases. As the test programme is
likely to continue over an extensive period, early clearance requirements will again
require an envelope restriction until the ultimate load cases have been completed.
Factors to be applied are given in Table 1.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
2.2.3 Absence of flight and ground loads measurement. The flight and ground
load surveys are usually undertaken late in the development programme when most of
the structural testing has been completed and a reasonable envelope can thus be
cleared for this stage of the flight testing. A factor of 0.9 should normally be applied to
account for the absence of aircraft loads data and validation of loads models used in
the flight clearance process.
2.2.4 Absence of qualification of any load limiting aircraft system. Aircraft
programmes have begun to employ “carefree handling” techniques where control of the
aircraft is vested in the flight control system and the development design process will
aim to ensure that the control system “respects” the strength envelope of the structure.
In some cases, the Design Organisation may seek to reduce the Ultimate Factor, on
the basis that the control system offers better protection from structural overloads than
would normally be the case for conventional aircraft. In such cases a factor based on
the ratio of ultimate factors should be considered for this element and the final level will
depend on development progress, where the following aspects will require
consideration:
limited carefree handling flights and evidence of structural overshoot levels;
Flight Control System (FCS) software development progress;
other load control systems; for example, pressure relief valves (PRVs). In these
cases, the upper tolerance level shall be used.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
3.1.5 Final Clearance for Development Flying. The final clearance stage
provides a release to 100% only when the load limiting capabilities of the flight control
system have been fully proven.
3.1.6 Release to Service. For this programme, the PTL agreed that the initial
Release to Service should be restricted at 93% (based on the ratio of 1.4/1.5) until it
had been proven that the flight control system fully respected the strength envelope
under service conditions. Confidence in the ability of the FCS to provide carefree
handling under service conditions would be obtained from the Structural Health
Monitoring System (event monitor) and should eventually lead to the removal of the
final factor.
3.2 It should be noted that individual components had different levels of load clearance,
allowing maximum exploitation of the flight envelope. For example, some flying control
surfaces had been fully qualified before the wing and fuselage.
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 04
1 INTRODUCTION
1.1 This Leaflet recommends the use of methods and design factors to ensure that
significant thermal effects are properly accounted for in the design and substantiation of
aircraft structures.
1.2 When a Design Organisation wishes to use a different method, he should seek the
agreement of the PTL.
1.3 Changes in structural temperature can produce the following effects:
a) stresses and strains caused by differential expansion within the structure;
b) a temporary loss in material properties if temperatures pass a critical threshold;
c) permanent deterioration in material properties due to prolonged exposure to
temperatures beyond a critical threshold;
d) creep due to prolonged exposure to temperature and stress;
e) for composite materials, stresses and strains due to local anisotropy.
Page 1 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
3.2 This Leaflet does not call for the provision of a safety margin by an arbitrary
increase in the structural temperature, either directly by factoring the temperature or
indirectly by factoring the design speed. Thus flight in excess of the design speed could lead
to a temperature higher than that considered in design. Although cases of exceeding the
design speed are rare, it is recommended that, as far as practicable, materials should be
used which do not suffer a substantial reduction in properties at a temperature slightly higher
than that corresponding to the design speed.
Page 2 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
5 DESIGN CASES
5.1 The Clauses of Part 1, Section 2 define the manoeuvres to be considered in design,
specifying aircraft speeds in each case, but they do not define those conditions giving rise to
thermal stress or strain (i.e., rate of change of speed and time spent at any given speed). For
each type of aircraft, therefore, it is necessary to examine all flight conditions in which the
combination of applied loads and thermal strains are likely to produce critical structural loads.
Normally, these loading cases are based upon the combination of full manoeuvre loads with
the most adverse thermal effects. Where such cases are thought to be improbable,
appropriate alternative cases should be formulated and these should be discussed with the
PTL.
5.2 The effect of thermally induced stresses either alone or in combination with other
loading actions should also be considered.
5.3 It may also be necessary to consider residual strength of vital controls and structure
during specified emergency conditions such as fire.
Page 3 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
6.5 In both methods the thermal strain, rather than the thermal stress, is factored. This
means that the ratio of stress at limit conditions to that at ultimate conditions is usually higher
for thermally induced loads than that for applied loads. The ratio is therefore also higher for
the combined loading case. This decision was made on the grounds that an ultimate factor
on thermal stress is not justified as far as static strength is concerned, because thermal
stresses are relieved when plastic deformation occurs. Consequently, a thorough
investigation of all fatigue and creep aspects is necessary, including creep buckling and
progressive distortion under cyclic temperature and loading conditions.
6.6 Cases where the load and thermal effects are opposite in sign may not be critical
because the most severe cases may well be when either one or other of these affects is
absent. If there is a critical case when both are present and the load effect is the greater it
may be desirable to disregard any alleviating thermal effects. When the thermal effect is the
greater, it appears reasonable to combine it with the load effect using the appropriate factors.
In either case, care should be taken to ensure that an adequate net load is used for design
purposes, particularly when the opposing effects are comparable in magnitude.
7 GROUND TESTS
7.1 When it is necessary to include thermal effects in strength tests in order to
demonstrate compliance with the requirements, the factored loads and, if practicable, the
factored thermal strains should be simulated. The method of simulating the factors on
thermal strain should be discussed with the PTL.
7.2 It should be noted that, in order to obtain realistic conditions in any structural
member, it will usually be necessary to test it as part of the complete structure so that all
balancing loads can be developed. When this is not done, results of the tests should be
interpreted bearing in mind any differences between the actual and the applied balancing
loads.
8 INSTABILITY FAILURES
8.1 For structures in which failure by instability is possible, thermal strain may be high
enough to produce structural collapse under low external loads. It is essential, therefore, that
this effect be taken into account in all design cases when considering the combination of
thermal strain and applied load.
9 REFERENCES
1. Metallic Materials Data Handbook ESDU 00932, ESDU International plc, ISSN
0261-2402.
Page 4 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 5
The content of this leaflet has been removed JAC Paper 1348 refers.
Page 1 of 1
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 6
STRUCTURES
GENERAL INFORMATION
1 INTRODUCTION
1.1 This Leaflet amplifies the requirements of Clause 3.1 and explains more fully the purpose
of the tests.
2 BASIC AIM
2.1 The basic aim of the tests, in conjunction with the static and fatigue strength tests, is to
ensure that the structure achieves the required strength factors both static and fatigue, under the
loads that are actually applied to it during flight and ground manoeuvres. This is particularly
important for novel designs for which there is no basis of comparison and on which calculated
loads could be seriously in error. Refer to Leaflet 3 Phase Flight Release.
2.2 To this end the first objective of the tests is to obtain measured values of the loads in flight
and on the ground in various manoeuvres and thus check the accuracy of the calculated loads on
which the design of the aeroplane was based and also provide information on the magnitude and
variation of the loads to be applied in the fatigue tests. It may also bring to light any loading
condition not foreseen in the theoretical assessment. In particular cases where structural
temperatures are expected to be significant temperature measurements can also be made.
2.3 In some cases the design loads may be found to have been overestimated and an
unexpected reserve of strength becomes available or, alternatively, an alteration thought
necessary because of a failure during strength tests may be found to be unnecessary.
3.1 Any strength tests done before the flight measurements are available can only be based on
calculated loads and are therefore only a check on the internal load distribution and the structural
strength under these calculated loads. However such tests are valuable in that they show up any
unforeseen weakness in the structure and this is borne out by experience which has shown that in
many cases the first design of a structure has needed strengthening alternations in order to carry
the design loads. For this reason, until the static strength tests have shown that the structure has a
satisfactory level of strength, flying must be restricted as required by Clause 3.1 Refer to Leaflet 3
Phase Flight Release.
3.2 When flight load measurements are available from the early restricted flight tests the early
strength tests can be re-assessed and the measured loads can be used in further static tests up to
factored conditions to ensure safety in flight tests up to higher loading conditions. An ideal
procedure is set out in Para 4.
4 PROCEDURE
4.1 The flight load measurement programme and the static strength test programme should be
planned together early in the design stage of the aeroplane. The static strength test procedure of
Leaflet 2 should be used but it should be phased in with the flight load measurement programme.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
4.2 The manoeuvres in which loads are to be measured in flight, as required by Clause 3.1,
cover a wide variety of loading conditions and it may not be practical to cover the corresponding
cases in the static strength tests in the time available. It may therefore be necessary to perform
some manoeuvres in flight without having first done ground tests. This has always been standard
practice however and the measurement of the flight loads will itself contribute to safety since by
progressively increasing the severity of a particular manoeuvre and comparing the measured loads
with the design loads, precautions can be taken should the measured loads prematurely reach the
design values.
5.1 The response of the aeroplane in terms of the loads, their distribution and their time history
during a particular manoeuvre or gust can be used in determining the representative loading cycle
to be applied to the structure in the fatigue test. The frequency of application of this loading cycle
at any particular level of severity must be obtained from Service records. Reliance entirely on a
calculated load cycle may produce large errors in fatigue test results and life estimation.
6.1 Two current methods of measuring loads in flight, which may be combined if required, are:
(a) strain gauging - A technique of flight load measurement using this method has been
developed in the United States and is described in Ref 1. This technique requires the
interpretation of the gauge responses in flight in terms of a static calibration on the ground,
(b) pressure plotting - This method has also been developed in the United States and
requires the installation of several hundred pressure points with corresponding pressure
transducers either for continuous plotting of pressure at each point or for use with a scanning
mechanism by which a number of points are scanned over a period of 2 or 3 seconds. A
comprehensive system of accelerometers is also necessary for the measurement of inertia
loads.
REFERENCES
No AUTHORS TITLE
1 Skopinski, T.H Calibration of strain gauge installation in aircraft structures for the
Huston, W. B measurement of flight loads N.A.C.A. Report 1178, August 1952.
Aiken, Jr. W.S
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 7
SYMMETRIC MANOEUVRES
1 INTRODUCTION
1.1 Clause 3.3 requires the use of a rational representation of the motion of the pitch
control motivator to be used to evaluate the loads which arise during a pitching manoeuvre.
The control systems of many aeroplanes are complex and it is not possible to present a
detailed analytical procedure for the load evaluation which will cover all cases. In the more
general case it is possible to give a broad outline of the stages of the analysis required to
derive the loads. For these cases see Leaflet 8.
1.2 Where there is effectively a direct link between the pilot's inceptor and the
deflections of the pitch motivator it is possible to give an analytical technique. Examples of
such closed algebraic forms may be found in R and M 3001 and 3327.
2 MANOEUVRE CONDITIONS
2.1 The complete pitching manoeuvre involves the transition of the aeroplane from
steady level flight at a particular speed and altitude to a condition of steady normal
acceleration, usually at some point on the boundary of the flight envelope; and then a
return to level flight. As the manoeuvre is primarily the response of the aeroplane to control
input in the short period longitudinal mode it is rapid and it is adequate to assume that the
forward speed does not vary during the manoeuvre.
2.2 The aim of the analysis is to specify the control input which results in a maximum
increment in the normal acceleration corresponding to the desired value at the appropriate
point on the boundary of, or within, the flight envelope. It should be noted that unless the
short period response is overdamped the motion of the aeroplane will be oscillatory, in
which case the maximum incremental value of normal acceleration will exceed the steady
state value appropriate to that control input. Whilst the steady state value is simply related
to the dynamic characteristics of the aeroplane for a given motivator deflection, the
maximum value is a function of the shape of the motivator movement versus time profile.
2.3 The final phase of the manoeuvre is conventionally analysed by assuming that the
aeroplane has achieved the steady state value of the normal acceleration and then
applying the pitch inceptor input required to return it to level flight.
2.4 From the foregoing description it will be realised that the important parameter is the
way in which the movement of the pitch motivator varies with time rather than the motion of
the pilot's inceptor. In a complex control system, such as one using active control
techniques, it is necessary to consider the effect of all the inceptors which contribute inputs
to the longitudinal control system. Refer to Leaflet 8 and Clause 3.10
Page 1 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
3 PILOT INPUT
3.1 For those cases where the pilot's inceptor is either directly linked mechanically to
the longitudinal control motivator, or through a power control unit having no other inputs it is
possible to specify the motivator movement in terms of the input of the pilots inceptor. Two
separate forms of motion have been used, in the past:-
(a) Unchecked; where the inceptor is moved to the appropriate deflection and
held there whilst the aeroplane responds.
(b) Checked; where after the initial movement of the inceptor, it is moved back
towards its initial setting, sometimes after a time delay.
3.2 The variation of the inceptor motion with respect to time generally takes a form
closely described by a ramp function, although for analytical purposes the unchecked
motion has often been represented by an exponential relationship.
3.3 The U.S. Mil-A-008861A (Inactive) Para 3.19.2 identifies three specific movements
of the pitch inceptor, all based on ramp functions. In each case the inceptor is assumed to
be deflected through an angle, such as to give the desired normal acceleration at the end
of the aeroplane response time. The type of aeroplane effects the time, t1, assumed to
achieve the angle, , it being 0.2 seconds for highly manoeuvrable aeroplanes. The three
conditions relate to the movement of the inceptor after the angle, has been achieved.
(a) Movement of the inceptor back to the original position in the same time as
the application, unless inceptor angle limitation requires a dwell time, t2, at angle
to enable the specified normal acceleration to be achieved.
(b) Dwell at the angle, , for a time t3 determined so that the specified normal
acceleration is just achieved in a total time (2t1 + t3), , being determined
accordingly.
(c) Dwell at the angle, , for a time, t4, and then movement of the inceptor to an
angle, /2, in the reverse direction in a time 1.5t1. Both t4 and are determined to
enable the specified normal acceleration to be just achieved in a total time (2.5t1 +
t4).
4.1 The first stage in the actual calculation of the loads which arise in a pitching
manoeuvre is to determine the response of the aeroplane to the specified pilot and/or other
inceptor inputs.
Page 2 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
4.2 This involves setting up the equations of motion of the aeroplane. In accord with the
assumptions of Para 2 these consist merely of the Z-force equation and the pitching
moment equation (as in the approximation to the short period motion of the aeroplane).
This completes the description of the dynamic system for the aeroplane, for which is a
prescribed function of time.
4.3 In the more general case, it is necessary to add all the equations governing the
dynamic behaviour of the control system.
4.4 To proceed to the calculation of a control surface loading, the effective incremental
angle of incidence at the control position is required. One of the response quantities that
can be obtained from the response calculation is the incremental angle of incidence of the
aeroplane (i.e. at the cg). This quantity may then be used to determine the incremental
angle of incidence at the longitudinal control/stabiliser surface, making due allowance for
pitching velocity and rates of change of downwash with time and wing incidence where
appropriate.
4.5 In addition the variation of motivator deflection during the manoeuvre will either
have been specified or will have to be determined as a response quantity.
5.1 At any time during the manoeuvre the total load is the algebraic sum of the load on
a control/stabiliser surface due to its incidence at that time and the incremental load due to
the deflection of any part of it at that same time.
5.2 When the aeroplane having a conventional wing-tail layout is pitched from steady
level flight towards the positive normal acceleration boundary the total surface load will
pass through a maximum download in the early stages of the manoeuvre, then through a
maximum upload before reaching the steady circling flight value. It finally passes through
another maximum in the upload sense when during the second stage of manoeuvre the
input is reversed to return the aeroplane to steady level flight. In the case of an aeroplane
having a foreplane-wing layout the foreplane load passes through a maximum upload
during the first part of the manoeuvre as the steady circling flight value is approached. A
maximum download condition may result as the input is reversed to return the aeroplane to
steady level flight. It should be noted that in this second configuration the critical design
loads conditions are reversed. That is the critical foreplane load is likely to arise in a steady
state condition whilst the transient case is likely to give the wing design conditions.
6.1 The moment about the hinge line of the aerodynamic load acting on a control
surface is also dependent on the incidence of the control/stabiliser where appropriate and
the control surface deflection. Again this passes through maximum positive and negative
values as does the load, when the aeroplane is of conventional wing-tall layout.
Page 3 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
7 GENERAL REMARKS
(a) variation of forward speed and the gravity component during the motion is
neglected, already mentioned.
(b) the contribution of the control motivator force to the total lift is omitted.
(d) the equations of motion are linearised and the aerodynamic derivatives
assumed constant throughout the manoeuvre.
7.2 Should it be considered necessary in a given case to account for any of these
effects, the equations of motion of the aeroplane must be amended accordingly.
7.3 If the data required under Clause 3.3 are not forthcoming from wind-tunnel or other
tests, it is acceptable that the effects under Clause 3.3 shall be estimated using appropriate
theories, in combination with the best available structural data. Whenever possible, the
effects under Clause 3.3 shall be based on the data from model and full-scale tests of a
closely similar design.
7.4 For aeroplanes with unswept wings and of conventional structural design flying at
low Mach numbers, in the absence of data from any of the investigations detailed above,
some allowances should be made. Changes in Cmo of ± 0.0075 and a shift of ±2.5% of the
mean chord in the position of the aerodynamic centre should be considered.
Page 4 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 8
SYMMETRIC MANOEUVRES
1.1 The application of the techniques described in Leaflet 7 is inappropriate for those
aeroplanes which use an active control system incorporating load limiting devices. In these
aeroplanes the automatic system can override pilot input and result in a so called "carefree"
manoeuvring situation. For the purposes of structural load evaluation it is usually necessary
to consider cases where motion of the aeroplane occurs in six degrees of freedom,
although constant forward speed may be assumed in some circumstances. The analysis
may be further complicated by the integration of flight and propulsion controls, as in the
case of vectored-thrust aeroplanes. Reference in all these cases should be made to Clause
3.10, which deals with the airframe aspects of active control systems design.
1.2 The use of a simple expression to specify the movement of the control motivators is
not appropriate for advanced flight control systems. The specification of the flight envelope
itself is relevant since the implied accelerations are one of the common design features of
the airframe and control system.
2 MANOEUVRE PARAMETERS
2.1 One way of proceeding with the design of both the airframe and control system is to
establish a set of "manoeuvre parameters". These include aeroplane velocities as well as
accelerations. They should be derived after consideration of the handling and agility
requirements of the aeroplane and in consultation with the Project Team Leader.
2.2 Design loads are calculated by balancing the aeroplane at selected points within
and on the boundaries using an assumed, appropriate control usage.
2.3 Subsequently, as the design of the flight control system becomes more fully defined
it is necessary to investigate the effect of control response on the manoeuvre parameters
and to check the load calculations in order to obtain flight clearance.
2.4 Reference should be made to Leaflet 28 for a further discussion of the structural
implications of active control systems.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
.
Modified pitching acceleration = ( q - pr)
.
where q is the pitching acceleration
A typical form of the envelope for a supersonic aeroplane is shown at Fig 1. As would be
expected maximum negative, nose down, pitching acceleration coincides with all positive
values of normal acceleration. Equally there is a reduction in design positive, nose up,
pitching acceleration as the maximum normal acceleration is approached.
3.2 The value of maximum normal acceleration is defined in Clause 3.2 and the
maximum pitch acceleration by aeroplane handling and agility considerations.
4.2 It is also necessary, to specify the combinations of pitching velocity, q, with normal
acceleration. The relationship shown in Leaflet 10, Fig 2 for asymmetric manoeuvre
parameters is also appropriate for combinations of q/q MAX and n/n1.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 9
ASYMMETRIC MANOEUVRES
1 INTRODUCTION
1.1 This leaflet gives explanatory notes on the requirements of Clause 3.3 and outlines
the methods of determining the critical stressing conditions arising in the course of the
manoeuvres specified therein. It is particularly applicable to those aeroplanes where the
movement of the control motivators is directly related to the movement of the pilot's
inceptors.
1.2 As in the case of the symmetric manoeuvres of Clause 3.2 and Leaflet 7, present-
day developments in aeroplane control systems can render past practices inapplicable in
the general case. Whilst in the past there always existed a direct relationship between a
given manoeuvring task and a control device, this is no longer the case. The demands of
improved performance in a number of directions have resulted in an increasing trend to
combine the effects of a number of control devices for a given task. Simple examples of this
trend are provided by the use of ailerons supplemented by differentially deflected tail
panels. When this is coupled with the increased application of automatic control systems
and unusual aerodynamic configurations the consequence can be an extremely complex
situation. However, there does exist a type of conventional aeroplane, namely those for
which the pilot's control is connected directly, or via a power unit only, to the rudder and
aileron control surfaces that can be catered for by simplified methods as hitherto (RAE
Report Structures 76).
2.1 Whether a yawing moment arises from the application of the yaw motivator(s) or by
a sudden loss of engine thrust on one side of a multi-engine aeroplane, the resulting motion
has the same character, being primarily a combination of yawing, sideslipping and rolling.
2.2 For the simple type aeroplane mentioned in Para. 1.2 above the first stage in
satisfying the requirements of Clause 3.4 is the calculation of the response of the aeroplane
as the solution of the lateral/directional equations of motion. In the more general case it is
necessary to include the equations which govern the dynamics of the control system when
determining the overall response.
2.3 The inputs referred to in Para 2.1 above principally give rise to yawing moments but
in certain cases there may be significant induced rolling moments and sideforces.
2.4 It has been usual for the linearised equations of motion to be used in calculating the
response of the aeroplane. Often the effect of the sideforce may be neglected in the
calculation of the overall response of the aeroplane but it does have a significant effect
upon the lateral and yawing accelerations. If the wing of the aeroplane is unswept and of
moderate to high aspect ratio it may be acceptable to neglect the rolling content of the
lateral motion, at least over part of the range of operating conditions.
When this is so for the simple aeroplane referred to, the response of the aeroplane is
derived from a solution of the two equations which describe the lateral and yawing motions.
Page 1 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
2.5 For the simple aeroplane it is adequate to assume that there is no input from the roll
or pitch motivators during the execution of the complete yawing described in Clause 3.4.
However, when inertial or aerodynamic cross-coupling is present it is admissible to
introduce corrective action both in the rolling and longitudinal senses. In order to ensure
that adequate allowance is made for the loads which may arise before this corrective action
is effective, Clause 3.4 specifies minimum values of bank angle, coupled with dynamic
sideslip angle and nominal acceleration which must be catered for, unless it can be
demonstrated that the flight control system ensures lower limited values.
3.1 The inputs into the yaw channel are generally specified such that the deflection(s) of
the appropriate motivator(s) may be derived. For the simple aeroplane with a more direct
link between the pilot and the yaw motivator the input may be specified directly or indirectly
in terms of applied pilot force or inceptor movement. See Clause 3.4
3.2 When the yaw motivator(s) is (are) sensitive or powerful the application of a specific
deflection may result in unrealistically high loads for the directional stabilising/control
surface. It may then be desirable to change the characteristics of the motivator to reduce
the load. Alternatively there may be strong reasons for retaining the sensitivity or power of
the control, in which case some reduction of the severity of the loading case might be
reasonable if there is an aeroplane characteristic (such as high lateral acceleration on the
pilot) which would act as a natural deterrent against high pilot inceptor inputs. In such
cases the Project Team Leader should be consulted.
3.3 If calculations indicate that sudden loss of lift on the vertical stabilising/control
surface is likely to occur in the specified manoeuvres, then compliance with the
requirement of Part 1, Section 2, Clause 2.22 is in doubt. It may be found necessary to
adjust the aeroplane characteristics in order to satisfy this requirement.
4.1 The stressing case specified in Clause 3.4 to cover the loads arising following the
loss of thrust or power of an engine envisages that corrective action will be taken by the
pilot. The case therefore is concerned with the response of the aeroplane to the sudden
application of a constant yawing movement. It is usual for the loads which arise in a yawing
manoeuvre case, Clause 3.4, to be higher than those consequent upon an engine failure.
However, should the aeroplane characteristics be such that the engine failure case gives
more severe loads, then the implication of the corrective action specified in Clause 3.4
needs to be carefully considered.
5.1 The solution of the overall equations which govern the motion of the aeroplane, as
described in Para 2 above, yields the response parameters which are inquired to determine
the loads acting on the vertical (directional) stabiliser/control surface(s). These are:
Page 2 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
From these the effective angle of incidence of a location on the vertical surface(s) may be
calculated. The corresponding deflection of the yaw motivator(s) is (are) also required to
enable the total loads to be evaluated.
6.1 Clause 3.4 specifies the loading conditions to be applied to the longitudinal
stabiliser/control surface in yawed flight.
6.2 It is required that the magnitude of the pitching moment coefficient, Cmo, be
increased to allow for the effect of sideslip angle. The required value should be derived by
the best available method and verified in the flight testing, but in the absence of more
precise information it should be assumed that the low speed value of Cmo is increased by -
0.0015 per degree of sideslip.
6.3 It is also required that the resulting trim (balancing) load be distributed
asymmetrically. The effective rolling moment on the longitudinal stabiliser/control surface is
defined as:
p V2 St bt Kß
2
K is the slope of the curve of longitudinal surface rolling moment coefficient versus
sideslip angle at the appropriate Mach number.
6.4 The value of the derivative K should, wherever possible, be verified by actual tests.
It depends upon the configuration of the aeroplane, and for instance in a conventional
aeroplane layout it is significantly affected by such factors as tailplane dihedral, sweepback
and the position of the tailplane on the fin and also to a lesser extent by other aerodynamic
components. However, experience has indicated that, for unswept, relatively low down
tailplanes having no dihedral, a value of K of -0.0025 per degree of sideslip is conservative.
Page 3 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
6.5 The provision of strength to meet the effects of asymmetric buffeting which may
occur near the minimum permissible flight speed has apparently been covered incidentally
on many past aeroplanes by the use of K = -0.0025 in the asymmetric flight cases. It is
important however to ensure adequate strength on aeroplanes which are designed for
much smaller numerical values of K. It will probably be satisfactory to cater for a difference
in lift coefficient between the two halves of the horizontal stabiliser of 1.0 or CL max,
whichever is smaller. As an overriding minimum it is suggested that the horizontal stabiliser
shear strength inboard of the pick-up points should not be less than the strength outboard.
6.6 During asymmetric flight the drag as well as the lift on the longitudinal surface will
usually be distributed asymmetrically. This asymmetry of drag loads should be considered
in the design, although the provision of extra structure to meet it should not normally be
necessary.
7.1.1 The roll input conditions to be used for stressing purposes are defined
according to the requirements of Clause 3.4 That is expressed indirectly in terms of
the specified minimum rate of roll. Since designers will usually ensure that this rate
of roll will be exceeded by some margin, a factor of one and a third has been
applied. To determine the deflection of single roll motivator (or the deflections of
individual motivators where a combination of these devices is used) that
corresponds to the rate of roll which results after the factor has been applied, it is
sufficient to use a modified single degree-of-freedom roll equation. For an aeroplane
employing wing-mounted flap-type motivators, such as ailerons, flaperons or
segmented flaps, the modification consists of a factor applied to the rolling moment
due to the roll motivator embodying the reversal speed, to account for the elasticity
of the wing structure. The one and a third factor takes some account of the indirect
effect of the yawing and sideslipping motions that must occur.
Page 4 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
7.2.4 The roll motivator power on some aeroplane designs may be so great that
the application of the requirements of Clause 3.4 results in unduly high values of the
rate of roll. In this case it is suggested that for the ground attack role, the maximum
rate of roll that need be considered to develop before the specified deflection of the
roll motivator (or the deflections of each roll-motivator, where a number of these are
used in combination) is removed, is the maximum that can be achieved in rolling
from a steady pull-out with zero bank into a 90 degree banked turn, when the
specified roll input is applied as rapidly as possible, or 200 degrees/sec, whichever
is less. For air fighting and evasive manoeuvres, the maximum rate of roll that need
be considered to develop before the specified deflection of the roll motivator (or the
deflections of each roll motivator, where a number of these are used in combination)
is removed, is the maximum that can be achieved in rolling from a 90 degree
banked turn in one direction to a 90 degree banked turn in the opposite direction
when the specified roll input is applied as rapidly as possible, or 200 degrees/sec,
whichever is less. The value of 200 degrees/sec is tentatively suggested as an
operationally usable maximum for fighters. Corresponding rates of roll for other
aeroplanes are not suggested as it is possible that on these aeroplanes the
maximum rate of roll corresponding to the specified roll input may be an adequate
criterion.
7.2.5 Any of the four stages of the manoeuvre, specified in Clause 3.4 (i.e., the
initiation, the steady roll, the arresting of the rolling motion and reversed roll), may
give critical loads in some parts of the structure. If instantaneous roll motivator
movement is assumed, it will probably be satisfactory to consider only the four
separate cases. If instantaneous roll input is not assumed, or if the maximum rate of
roll considered is less than the steady rate of roll corresponding to the applied roll
input, then examination of the response of the aeroplane throughout the manoeuvre
becomes advisable.
7.3.1 Solution of the appropriate equations for the overall aeroplane system, yields
the response quantities necessary for the calculation of the loads acting on the
wings, their attachments to the fuselage and the directional stabiliser/control
surface(s).
Page 5 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
(b) the distributed effective angle of attack resulting from the rate of roll,
(c) the corresponding quantity associated with the rate of pitch, and
8 GENERAL REMARKS
8.1 Although the effects of the sideslip which occurs during asymmetric manoeuvres are
most easily calculated for the fin and rudder loads, there are noticeable effects on other
parts of the airframe which are not so amenable to calculation. Such items include
undercarriage doors, cabin hoods, radar scanning housings and externally carried stores,
in general. Special wind-tunnel tests to investigate the magnitude of these effects are
recommended whenever it is not possible to base an estimate on the results of previous
tests of a closely similar design.
8.2 The yawing velocities which occur during asymmetric manoeuvres cause a
gyroscopic pitching couple on the engine and propeller which should be considered in the
design of the engine mounting. The effect of rate of yaw on other rotating parts within the
aeroplane should not be overlooked. See Clause 3.1 and Leaflet 3.
Page 6 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 10
ASYMMETRIC MANOEUVRES
1 INTRODUCTION
1.1 The application of the techniques described in Leaflet 9 is inappropriate for those
aeroplanes which use an active control system incorporating load limiting devices. This
different design situation is discussed in Leaflet 8 to which reference should be made, as
well as in Clause 3.10 and Leaflet 28.
1.2 One way of dealing with the design is to make use of Manoeuvre Parameters, see
Leaflet 8, Para. 2. These should be determined after consideration of the handling and
agility requirements of the aeroplane and in conjunction with the Project Team Leader.
2.1 Clause 3.4 specifies the yawing manoeuvre load cases for aeroplanes with
conventional controls. In the case of aeroplanes with active control systems it is useful to
consider an envelope which describes the interaction of lateral acceleration and yawing
acceleration. In practice it is convenient to use a 'modified' yawing acceleration defined as:
.
Modified yawing acceleration = ( r +pq)
.
where r is the yawing acceleration
p is the rate of roll
q is the pitch rate
2.2 The maximum value of the yawing acceleration is determined by the control system
characteristics required to meet the aeroplane handling and agility requirements. The
maximum lateral acceleration is likely to be limited by consideration of crew tolerance.
3.1 Clause 3.4 specifies the design load cases for an aeroplane with conventional
controls in combined pitching and rolling manoeuvres. The situation is more complex in the
case of aeroplanes designed to possess 'carefree' manoeuvring characteristics due to the
need to consider all six degrees of freedom at the same time.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
3.2 Fig 2 is an example of an envelope which specifies the relative values of lateral
manoeuvre parameters in terms of the normal acceleration. The lateral manoeuvre
parameters covered are:-
The diagram shows that for values of each of these parameters of up to 25% of the
maximum value there is coincidence with the maximum normal acceleration. There is then
a linear variation so that 80% of the maximum value coincides with 80% of the normal
acceleration. Further linear variation of the parameters occurs so that the maximum values
coincide with 50% of the normal acceleration.
3.3 The maximum values of the roll and yaw rates and acceleration are determined by
consideration of the handling and agility requirements of the aeroplane.
4.1 The combination of maximum angular acceleration with angular rate for both the roll
and yaw motions is a further consideration. A typical diagram showing this relation can be
found in Leaflet 8, Para. 4.1 and Fig 2.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 11
GUST LOADS
1 INTRODUCTION
1.1 To enable Service flying to be undertaken in all weather, pilots must be advised of
the best speed at which to fly in severe turbulence. This Leaflet outlines the type of
investigation which should be made to determine the conditions of speed and altitude at
which there is least risk of structural damage or loss of control resulting from excessive
incidence in a gust. The outcome of this investigation together with practical flying
experience will form the basis of advice given to pilots.
2.1 When an aeroplane meets a severe gust at low altitude, the upper limit of the safe
speed range is the speed at which the gust load can just be withstood without structural
damage. The lower limit is the speed at which the incidence during the passage of the gust
just reaches that at which loss of control would occur as a result of stalling or instability, or
at which severe buffeting would endanger the aeroplane. Within these limits lies the safest
speeds at which the severest gust can just be withstood without either structural damage or
loss of control. At high altitudes, the load developed in a gust may be insufficient to cause
structural damage throughout the speed range of the aeroplane, and loss of control or
severe buffeting is then the only danger.
2.2 Safe speed ranges at various altitudes can be shown on a speed-altitude envelope.
A safe speed-altitude envelope and the safest speeds within the envelope should be
determined for gusts of 20 m/s (66 ft/sec) EAS below 6100 in (20,000 ft) altitude thereafter
decreasing linearly to 11.6 m/s (38 ft/sec) at 15,200 in (50,000 ft) altitude, with the
aeroplane at the most adverse weight for the altitude considered. Fig 1 shows examples for
hypothetical subsonic and supersonic fighter aeroplanes.
3 STRENGTH BOUNDARY
3.1 The strength boundary should correspond to the achieved strength, unfactored,
resulting from the gust or manoeuvring strength requirements, and should be determined
for upward, downward and sideways gusts.
4.1 The loss of control boundary should be regarded only as an indication of the
conditions in which control difficulties might be expected to occur in severe turbulence.
4.2 On low speed, low altitude aeroplanes, the loss of control boundary should be
assumed to correspond to the stalling C L max. On high speed, high altitude aeroplanes,
flight tests will be necessary to establish the C L - Mach No. boundary that can be achieved
without loss of control, dangerous instability or severe buffeting. It will usually be
impracticable for this to be obtained from flight tests in severe turbulence, so the boundary
obtained in manoeuvring tests will have to be used.
Page 1 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
4.3 It will generally be unnecessary to allow for the transient increase in C L max or
buffet C L that occurs during the rapid change in incidence produced by a gust, as this
effect, which is difficult to determine, provides additional safety from stalling or buffeting
beyond the boundary obtained assuming static C L max or C L max determined from
manoeuvring flight tests.
4.4 In addition to the effects of excessive incidence, control difficulties may arise in
continuous turbulence at high altitudes and supersonic speeds, as a result of deterioration
in damping and control effectiveness. Control difficulties may also arise in turbulence at low
altitudes and high subsonic speeds because of over-effectiveness of the elevator or aileron
control. These may be particularly dangerous when attempting to fly at high speed near the
ground. These effects cannot be determined from considerations of the aeroplane meeting
an isolated gust. Advice on the flight procedure to be adopted where they are present will
have to be based on experience gained during flight in turbulence.
5.1.1 For aeroplanes on which compressibility effects can be neglected the safest
speed is a constant E.A.S. at low altitudes, and is limited at higher altitudes by the
maximum speed of the aeroplane.
5.2.1 The variation of safest speed with altitude for a high speed subsonic
aeroplane can generally be represented by a constant E.A.S. at low altitudes where
the upper limit of the safe speed depends on strength, and a constant Mach No at
higher altitudes where the upper limit depends on the deterioration in maximum safe
C L as transonic speeds are approached.
5.3.1 Supersonic aeroplanes are safer both at subsonic and supersonic speeds
than in the transonic range, where the load produced by a gust reaches a
maximum, and the C L at which loss of control or buffeting occurs deteriorates to a
minimum. The conditions, under which it is safer to fly supersonically than
subsonically, and vice versa, are an important consideration.
5.3.3 Loss of control as a result of excessive incidence in a gust will set only a
lower limit to the safe speed in supersonic flight. At any given altitude an aeroplane
is safer in this respect at supersonic speeds, once it is through the transonic range,
than at the safest subsonic speed.
Page 2 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
5.3.4 The safest flight procedure for a supersonic aeroplane will generally consist
of a constant subsonic E.A.S. up to a point A on Fig 1 corresponding to the altitude
above which the safest subsonic speed is determined entirely by control difficulties
or buffeting. This will be followed by transition to supersonic speed at a safe
altitude. On some aeroplanes with low gust or manoeuvre load factors the control
and strength boundaries may intersect in the transonic range, indicating no clearly
defined safe speed for the gust specified. In these circumstances the recommended
speeds should be chosen on the safe side of the strength boundary.
6 OPERATIONAL CONSIDERATIONS
6.1 For operational reasons, it is not always practicable to fly at the safest speed
determined from the foregoing considerations, and a compromise between this and the
normal operating speeds for best rate of climb and descent, or maximum range then
becomes necessary. Consideration of the speed band within which 95% of the maximum
range could be obtained might assist in reaching a compromise. On high altitude
aeroplanes, particular attention should be given to climbing and descent where turbulence
is most likely to be met. Any special limitations which might arise from the use of flaps or
airbrakes during descent should be considered. For crew comfort and efficiency, a speed
somewhat lower than the safest speed may have to be chosen, particularly with supersonic
aeroplanes in which excessive pitching may be troublesome at supersonic speeds.
6.2 It is important that the speed finally recommended in Pilots' Notes should be in the
simplest terms, so that it is easily remembered. A range of speeds should be given
covering all stages of flight, and lying within the strength boundary of the safe speed-
altitude envelope as near as possible to the safest speed. If a single speed can be chosen
that is a reasonable compromise for all stages of flight, this should be quoted for simplicity.
Complicated variations of speed with altitude should be avoided as far as possible.
Page 3 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
Page 4 of 4
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 12
GUST LOADS
1 INTRODUCTION
1.1 This leaflet presents a method of calculating gust loads for conventional aeroplanes in the
absence of a rational investigation of the aeroplanes response to gusts. It applies only to
aeroplanes with straight or sweptback wings and tailplane behind the wings. It may be used as a
preliminary design aid to more detailed full response calculations. The full response calculations
should always be undertaken for aeroplanes of unconventional layout or abnormal stability or
when structural flexibility is likely to be significant.
2 GENERAL
2.1 The speed and attitude of the aeroplane should be assumed to be unchanged during the
passage of the gust.
2.2 The maximum change in load on an aerofoil produced by the gust should be assumed to
equal that resulting from a change in incidence of:
tan-1 FU
V
2.3 Allowance should be made for the extent to which the lift coefficient of an aerofoil, where
effective angle of incidence is increasing rapidly, may exceed the maximum value at steady flow.
In the absence of better information the maximum lift coefficient assumed should be at least 1.25
times the maximum static value corresponding to the appropriate Mach number.
2.4 In calculating the loads on the horizontal tail surface the conditions of Clause 3.5 are to
be applied. Allowance should be made for the change in downwash produced by the gust on the
mainplane.
3.1.1 Mainplane - The alleviating factor for vertical gusts is given by:
0.88
Subsonic flight: - F =
5.3
103
.
Supersonic flight: - F=
6.95 103
.
2M
where µ - is a mass parameter as: - µ =
Sca
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
3.1.2 Horizontal tail service - The alleviating factor for vertical gusts acting on the
horizontal tail surface in subsonic flight should be that derived for the mainplane in
Para 3.1.1. In the case of supersonic flight it should be taken as unity.
0 .88 L
Subsonic flight: -
5. 3 L
2
2M k c
µL =
St ct a, t t
a, t - is the slope of the vertical tail lift coefficient/incidence curve at the appropriate Mach
number (per radian)
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 13
PRESSURE CABINS
STRENGTH TESTING
1 INTRODUCTION
1.1 This Leaflet amplifies the strength test requirements of Clause 3.6 and makes
recommendations on testing methods.
2 LOADING CONDITIONS
2.1 The requirements of Clause 3.7 call for tests of the pressure cabin under the
loading conditions either of Clause 3.7.37 or of Clause 3.7.38, or of both, as agreed with
the Project Team Leader. The loading cases and the method of testing to be used should
be discussed with the Project Team Leader.
3 METHOD OF TEST
3.1 The best method of testing a cabin under internal pressure, whether static or fatigue
loading is needed, is to use water as the pressurising medium and so avoid the risk of
explosive failure which can occur if air is used (see Para 4). The specimen is placed in an
empty tank and both specimen and tank are filled with water until the specimen is
completely submerged; water pressure is then applied internally. During all filling and
emptying processes the water level in the specimen should be the same as that in the tank
thus avoiding any loading of the fuselage due to static head effect.
3.2 With this method it is possible to encounter a premature failure, identify a weak
spot, repair and strengthen, and then continue the test with the minimum of delay. In cases
where the failure occurs at or near the design pressure, it is possible to continue the tests
to higher pressures after repairing and strengthening the specimen so that data may be
obtained for the design of future projects.
4.1 Calculations of the energy stored in compressed air show that tests using air
pressure can be exceedingly dangerous since if failure occurs, the sudden release of
energy may lead to widespread destruction. This has been borne out by experience which
has shown that where failure originates in the metal structure violent explosion occurs with
excessive tearing and with bodily movement of the test specimen. Where the failure occurs
in a window the damage is usually limited to the window, but danger to personnel is still
present from flying fragments and from movement of the specimen.
4.2 Special precautions must therefore be taken to protect personnel and adjacent
equipment. There have been cases where protective steel netting and cables have been
broken and shelter doors sucked out of their frames by blast. Observers should therefore
remain at a safe distance and should not approach the specimen until the pressure is
reduced to a known safe value. The danger can be reduced if the quantity of air in the
cabin is reduced to a minimum by filling the cabin with some suitable material of low
density, but internal loading due to the weight of the material should be avoided.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
4.3 Apart from the danger aspect, explosive failure causes so much secondary damage
that the determination of the primary cause of failure is extremely difficult and even if it can
be identified, repair and strengthening of the weak spot may be impossible. The structure
may then be useless for further tests, and in a case of premature failure it would be
necessary to build another specimen before the tests could be continued and, for a large
fuselage, this would mean great expense and long delay. The advantages of using water
as the pressurising medium are therefore overwhelming.
5.1 From the foregoing it is evident, in the interests of safety, that all tests in which the
use of air pressure cannot be avoided, (e.g., functioning tests, leakage tests and the
proving pressure test as required by Clause 3.7) should be deferred until the static test, as
required by Clause 3.7 has been satisfactorily completed.
5.2 There seems to be no practical way of avoiding the use of compressed air for these
proving tests and they will therefore be attended by some risk of explosion through failure
of the pressure shell. This risk will be reduced as much as possible by the previous
strength tests on the test specimen. However, it is not always practical to do the water
pressure test on the whole of the pressure shell. Hence although all known design
difficulties may have been represented, some risk may still arise from some unknown
design weakness not included in the strength test specimens, or through some design
modification, works' concession, bad workmanship, service defect or repair causing a
reduction in the strength of the pressure shell.
5.3 Although the risk of explosion may be slight it is recommended that during all air
pressure tests in which the maximum differential pressure for normal operation is to be
exceeded, the following safety precautions be taken:
(a) the test should be made in the open (the use of blast walls for tests made
within a building tends to increase the risk of roof damage and falling fragments),
(b) the test cabin should be well away from other aeroplanes or any other
objects which could be damaged, and
(c) all personnel should remain at a safe distance and only approach for close
observations, such as leakage tracing, after the pressure has been reduced to the
normal working pressure.
6.1 In routine tests of the pressure cabin system on aeroplanes in service it will not be
necessary, and indeed is undesirable, that the maximum differential pressure for normal
operation (i.e., the pressure "p" of Clause 3.6) should be exceeded. No special precautions
are therefore needed.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 14
PRESSURE CABINS
1 INTRODUCTION
1.1 This Leaflet gives recommendations amplifying the requirements for warning and
pressure controlling devices, for the prevention of damage due to discharge or leakage
from the pressure supply system, and for the operation of cockpit hood inflatable seals.
2 WARNING DEVICES
2.1 It is essential that the warning device should operate only when the pressure loss is
sufficient to constitute a state of emergency, which the pilot must quickly do something
about; too frequently premature warnings seriously reduce the effectiveness of the warning.
2.2 To prevent unnecessary operation of the warning device due to the rapid change in
differential pressure during a dive, it is recommended that the warning device should
operate only when the differential pressure falls by more than 7 kPa below the value set by
the pressure control unit coupled to the discharge valve. There is, however, no need for the
warning to operate at altitudes below 7,600 m (25,000 ft), since even sudden pressure loss
would not be serious enough to warrant a warning.
3.1 GENERAL
3.1.1 The requirements of Clause 3.6 have, for convenience, been presented in
terms of various types of valves but in each case any device which ensures
compliance with the basic requirement will be termed a valve for the purpose of the
requirement.
3.2.1 The requirements call for a safety valve as a safeguard against failure of the
discharge valve. Acceptable means of meeting the requirement are to duplicate the
discharge valve (see also Para 3.2.3) or to incorporate the safety valve and the
discharge valve in a single unit (see also Para 3.2.4 and 3.2.5).
3.2.2 In any case, the inlet to the discharge and/or safety valves should not be
protected by a fine mesh filter as blockage of such a protective mesh has been the
direct cause of cabin failures.
3.2.3 If any part of the pressure control system should fail, this should cause the
discharge valve(s) actuated by the control system to maintain a safe differential
pressure in the cabin.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
3.2.4 Although it is desirable to have the discharge and safety valves completely
separate, consideration of icing prevention, space, etc., may dictate that the valves
be incorporated in a single unit. The flow of warm air through the discharge valve
tends to keep the safety valve free from ice.
3.2.5 If the discharge and safety valves are mounted together with a common
inlet, one of the following alternative arrangements is recommended:
(b) the flow of air to the common inlet should be ducted through more
than one inlet; each inlet in the ducting should be suitably protected by a
coarse mesh debris guard.
3.2.6 The safety valve should open when the cabin pressure exceeds the
maximum working differential pressure and be fully operative as soon as
practicable, but in any case before the pressure has reached 7 kPa above the
maximum working differential pressure (see Clause 3.7).
3.3.1 In complying with the requirements for the inwards relief valve, consideration
should be given to the conditions arising when owing to enemy action or to failure of
the air supply, the cabin pressure drops to atmospheric while the aeroplane is at its
operational ceiling following which it dives steeply to a low altitude.
4.1 Ductings will contain air at high temperature and at high pressure and hence
adequate precautions should be taken to ensure that the discharge of air from relief valves
or leakage from components cannot damage adjacent structure, equipment or services.
5.1 On aeroplanes fitted with sliding hoods and inflatable seals, the last movement of
the hood in closing should cause the seal to inflate automatically and the first movement of
the control for opening the hood should cause the seal to deflate.
6.1 The maximum local working temperature and pressure should be measured at all
critical points in the prototype system to check that the temperature and pressures used in
the design are not exceeded. The same measured temperatures and pressures should be
used in the fatigue test of Clause 3.7
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 15
PRESSURE CABINS
1 INTRODUCTION
1.1 The requirements of Clause 3.6 state that every possible precaution shall be taken
to avoid contamination of the air in the pilot's cabin. Experience has shown that serious
contamination, and a fire hazard, can readily arise when air tapped from the engine
compressor is used both for pressurising the fuel tanks and for air conditioning the cabin.
This Leaflet is concerned with the design precautions needed to prevent contamination in
such a case.
2 BASIC PRINCIPLES
2.1 There are two main ways by which the possibility of contamination can be greatly
reduced. They are:
(a) by arranging that the tapping which supplies air for cabin conditioning is
positioned well above, and remote from, that which supplies air to the fuel tanks,
and
(b) by using a reliable non-return valve in the air pipe from the engine to the fuel
tanks.
2.2 On many aeroplanes where only one of those two alternatives has been employed
cases of cabin air contamination have arisen, either due to the malfunctioning of the non-
return valve or because the two engine tappings, although separate and a few centimetres
apart, were nevertheless so placed that fuel coming back from the tanks into the engine
could enter the air pipe going to the cabin. It is therefore strongly recommended that both
courses of Para 2.1(a) and (b) should be adopted.
3.1 It is obviously desirable that the two tappings be as far apart as is reasonably
practicable, and that the tapping for the fuel tank air supply should be below that for the
cabin air, thus ensuring that any fuel coming back from the tanks cannot drip into the cabin
air supply tapping. The tapping for the fuel tank air supply, being the one where danger
arises, should be placed so that the flow of air through the engine will tend to carry fumes
away from the cabin air tapping when the engine is running. It will usually be satisfactory to
take any subsidiary air supplies, (e.g., for gun heating, equipment cooling or pilot's
ventilated suit), from the cabin air tapping. In other words, as the danger arises from the
pipe going to the fuel system, that is the one to keep apart from others.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
4.1 Two points need study here - the kind of valve used and its location.
4.2 The valve should be one which is absolutely leakproof when in the nominally closed
position and which can be relied upon to function with the necessary reliability in a pipe
passing air, or air and fuel vapour mixed.
Note: Advice on non-return valves already under development for this purpose can
be obtained from the Project Team Leader.
4.3 As regards location of the non-return valve, the designer will probably wish to make
one valve serve for several fuel tanks, if not for the whole fuel system and will thus wish to
place the valve fairly close to the engine. A further advantage of doing this is that, by
incorporating a testing connection near this non-return valve, the whole of the fuel system
can be subjected to a pressure test as a convenient servicing operation when desired.
However this location of the non-return valve close to the engine may subject it to high
temperatures and the aeroplane designer should establish that the valve he chooses for
the job can withstand the temperatures to which it will be subjected in the particular
installation. Duplication of the non-return valve (i.e., two of them in series) may be justified
in some cases.
5.1 Some aeroplanes do not take the cabin air from an engine tapping, but have a
separate blower or, possibly, bottled air supply. Even in such cases the fuel system needs
careful study as an air/fuel mixture might be a serious fire hazard if it passed into, say, an
electrical generator cooling system.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 16
PRESSURE CABINS
1 INTRODUCTION
1.1 Clause 3.6 requires that the time needed for an emergency descent from the
maximum operating altitude down to a safe altitude shall be established by flight tests so as
to confirm the calculations made in complying with the pressure cabin sealing
requirements. This Leaflet lays down the features which should be considered when
making the design calculations and flight tests.
1.2 In addition, the flight tests are required to provide information on the adequacy of
the choice of the pressurised clothing and equipment specified in the Aircrew Equipment
Assembly (AEA).
2 STARTING CONDITIONS
2.1 The time required may vary markedly depending on the initial flight conditions. On a
fighter, for example, the time for descent from a given operating altitude will probably be
quite different for initial conditions of say M = 0.7 and M = 2.0. It may also vary appreciably
depending on whether the aeroplane is initially in a steep climb, in level flight or in a dive.
Other factors that will affect the time taken are initial engine conditions (e.g., throttle
position, reheat on or off) and possibly the position of the air brake, nose flaps, etc.
External stores or weapons may affect the time required by imposing speed or handling
limitations. On a bomber, although there is likely to be less variation in initial conditions,
there can be a large variation in operating altitude due to the large change of weight during
a sortie.
3 TEST TECHNIQUE
3.1 It is for the Contractor to propose for each aeroplane the conditions under which
descent measurements should be made. In this connection, it must be remembered that in
a real emergency, the pilot will probably have suffered sudden decompression immediately
prior to descending. Any descent procedure adopted should therefore be straightforward,
simple to do and not necessitate precise control of the aeroplane. For this reason, it is
suggested that the tests should be made at what are likely to be the normal Service limits
for the aeroplane. Obviously on a real emergency descent, it may be expected that normal
Service limits will on occasions be exceeded and this is acceptable provided that a
condition is not reached where the aeroplane would be seriously endangered. However if
design speeds are not to be grossly exceeded, this will probably mean that the
recommended technique for the descent should be the Normal Service limit and that tests
should therefore be made at this condition.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
3.3 On an emergency descent, an appreciable part of the time may be spent in losing
engine thrust and reaching steady diving conditions with air brakes out. This time should be
included in the design estimates and in the measured times for descent.
3.4 The flight condition after reaching the safe altitude is unimportant provided the flying
limitations are not exceeded and it can be assumed that the aeroplane continues through
the safe altitude in the steady dive condition.
4.1 It is obviously desirable that the tests should be made as soon as practicable after
the aeroplane has been cleared for high altitude flying, so that the adequacy of the
pressure cabin sealing can be checked and the final decision on the type of pressurised
clothing and breathing equipment to be fitted can be made.
4.2 However, it is also important that the development state of the aeroplane should be
such that the results that will be obtained on a production aeroplane can be reliably
estimated, since too high a complexity of pressure clothing is as undesirable as too low a
standard of safety.
4.3 The following list gives examples of the development stage that should ideally have
been reached before making these tests:
(a) that the likely Service limiting speed and Mach Number are known and that
the aeroplane has been cleared for flight at those speeds with the appropriate
normal accelerations,
(b) that the airbrake design has been finalised and any limitations on use have
been established,
(c) that the engine and full system have been cleared for maximum 'g' push
overs, or alternatively if the dive is to be entered by a half roll and pull through, that
the aileron limitations have been established over the speed range, and
(d) that the engine idling thrust is representative of production engines and that
any engine handling difficulties have been resolved.
5 REPORTING
5.1 Copies of the flight test report should be sent, through the MOD Resident Technical
Officer (RTO) to the Project Team Leader.
5.2 The choice of pressurised clothing and equipment for service use in the aeroplane
will then be reviewed. If any changes are necessary, they will be conveyed to all concerned
by a re-issue of the AEA.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 17
1 INTRODUCTION
1.1 Clause 3.8 prescribes that the spinning conditions used for strength calculations shall
be those which give the most adverse combination of the yaw, pitch and roll components of
the rotation, with adequate allowance for unsteadiness. This leaflet outlines the means which
may be used to derive the design rate of rotation in the spin.
2.1 The design rate of rotation may be derived by one or more of the methods outlined
below.
18 . 3 V 1
S 2 rads / sec
K Z2 K X 2
K Z and K X - are the radii of gyration about the Z and X aeroplane axes of stability,
respectively, m.
S is the steady rate of rotation about a space vertical associated with an aeroplane
incidence of 60° as defined by the inclination of the aeroplane longitudinal axis to the space
vertical with the aeroplane lateral axis horizontal. The equation has been derived from a
consideration of the pitching moment equation in the steady spin. The value of the constant
was derived by consideration of measured rates of rotation on a series of models with a
standard yawing moment applied. Good agreement was obtained between the measured
model rates of rotation and the prediction of the formula and there was fairly good agreement
with the average rates of rotation measured full scale during incipient spinning tests. The
formula applies only to aeroplanes of conventional layout and moderate aspect ratio. It is
suggested that a factor of at least 1.15 be applied to S to allow for unsteadiness during
spin.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.3.1. Various model test techniques are available (see for example AGARD CP 199
Paper 13). Model scaling is extremely important and it is especially necessary for the
model Reynolds number to be as representative of full scale as possible. Where the
steady rate of rotation has been evaluated from model tests it is necessary to
compare it with that given by the formula of Para 2.2 above and if less the procedure
to be adopted shall be discussed with the Project Team Leader. The factor of 1.15
should be applied to the agreed steady rate of rotation to cover unsteadiness.
2.3.2 Model test techniques which will be acceptable for providing evidence of spin
characteristics include:-
(b) Vertical wind tunnel tests which can be used to obtain data on the
steady developed spin and recovery, but usually require careful interpretation
because of the disparity in the Reynolds number.
(c) Free-flight tests with model dropped from a helicopter, with either pre-
programmed control settings or radio control.
(d) Wind tunnel rotary balance tests, which may be used to obtain the
basic aerodynamic characteristics of the aeroplane whilst in the spinning
mode and hence the rates of rotation by solving the equations of motion of the
aeroplane (see for example AIAA Paper 80-1564).
2.4.1 Simulator studies demand the availability of a valid mathematical model of the
aeroplane and for this reason the application to the spin regime may be limited.
However, the technique has value in investigating the conditions from which a spin
may develop.
3.1 In the absence of other information a rate of rotation of 5 rad/sec about the
longitudinal axis, assumed to be vertical, should be used to cover the condition of rolling
during recovery from a spin or post stall gyration.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 18
1 GENERAL
1.1 In the case of training aeroplanes which are to be certificated as Part 1 designs (see
Part 1, Section 2, Clause 2.1) it is necessary to establish the spinning and spin recovery
characteristics of the aeroplane (see Part 1, Section 2, Clause 2.24 and Section 2 Leaflet
53).
1.2 The Aeroplane Specification may state that aeroplanes in other categories,
especially highly manoeuvrable combat types, have to be designed to meet spin conditions,
although they should also be resistant to departure from controlled flight and to post stall
gyrations and spins in accordance with Part 1, Section 2, Clause 2.24 However, if an
automatic spin prevention device with a sufficiently low probability of failure is used Clause
3.8 states that the spinning requirements to be applied must be agreed with the Project
Team Leader. If such a device is part of an automatic flight control system reference should
be made to Clause 3.9
2 SPIN RESISTANCE
(see for example AGARD CP-199 Paper 5, and CP-235 Paper 19).
2.1 The aim of spin resistance is to incorporate characteristics in the aeroplane which
result in resistance to departure into spin type motions subsequent to a stall.
2.2 Design criteria suitable for use during the design stage have been developed, but
must be substantiated by subsequent model and full scale testing as prescribed by Clause
3.8.3. The full attainable incidence range should be investigated.
I
N v DYN N cos z L v sin
v Ix
This is derived from a consideration of the lateral directional equation of motion of the
aeroplane. It is a measure of the stability of the aeroplane about the flight path and predicts
departure from longitudinal control inputs. Essentially N v
DYN should be greater than
zero in all conditions to avoid the possibility of departure in yaw.
Page 1 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
Various versions of this criteria have been proposed. They have been developed from a
consideration of the simplified rolling and yawing moment equations assuming the
aeroplane to be laterally and directionally trimmed. Essentially the criterion predicts the
point where the rolling moment due to sideslip, resulting from adverse yaw, overcomes the
rolling effect of the ailerons and, where appropriate, rudder.
N KN
LCDP = N v - L v
L KL
where K = and represents the aileron/rudder interconnect
and L v o
Nv Ix
where - tan 1
Lv Iz
N KN
tan 1 Ix
L KL Iz
(e) Some evidence suggests that the value of the pitching moment due
to sideslip C m v should be zero or
Page 2 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
3 SPIN RECOVERY
3.1 The design of an aeroplane for spin recovery is complicated by the number of
design parameters which influence the problem. The most important of these are:-
(a) The relative moments of inertia of the aeroplane about its three reference
axes in all appropriate configurations including with stores.
(b) The wing geometry, particularly on the outer parts of the mainplane.
(c) The fuselage shape, especially in cross section and nose shape.
(d) The relative position of the horizontal and vertical stabilising surfaces,
including the degree to which the fin and rudder are shielded by the tailplane in spin
attitudes.
(e) The effectiveness of the motivators and their cross axis effects.
3.2 A relatively simple criterion known as the Tail Damping Power Factor (TDPF) was
used for many years. This basically relates the product of a so called tail damping ratio and
the 'unshielded' rudder volume coefficient to the difference between the moments of inertia
in roll and yaw. More recent work has shown clearly that it is not a satisfactory criterion as it
does not describe either a necessary or sufficient characteristic for satisfactory spin
recovery. This is not surprising in view of the limited parameters covered.
3.3 A more comprehensive criterion is that given in ARC CP No. 195. This method
predicts an overall unbalanced (anti-spin) rolling moment coefficient (URMC) and relates it
to the moment of inertia parameter
1 Iy
Ix
The unbalanced rolling moment coefficient is the sum of three separate quantities:-
(a) that due to the body damping coefficient, L B , which makes allowance
for body shape - Fig 1a. The effect depends upon the width of the strakes, if
present, which are horizontal extensions to the fuselage forward and in the
plane of the tailplane.
(b) that due to the unshielded rudder volume coefficient, L RV Fig 1b. The
elevator position used should be that which results in the minimum
unshielded rudder area.
Page 3 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
The boundaries for satisfactory spin recovery are shown in Fig 3. The technique has shown
good correlation with flight tests for aeroplanes of conventional layout with wings of
moderate aspect ratio. The parameter:-
1 Iy
Ix
represents a simplification which effectively assumes that the second moment of mass in
the z direction is negligible. This may introduce errors when I x is approximately equal to I y .
However, the method considers only fin/rudder recovery effect and is thus limited to
aeroplanes of conventional layout. AGARD CP 199 Paper 7, Fig 10 suggests a criterion for
roll control which may be dominant on slender configurations.
4 NOTATION
b Wing span
bl I z I x / [S (2b) ,]
Cm v pitching moment derivative due to sideslip
I x ,I y ,I z moments of inertia about body axes in roll, pitch and yaw respectively
S Wing area
Page 4 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
angle of attack
N I
tan 1 v x
Lv I z
N KN I
x
tan 1
L KL I z
1
1. 3 2 b
AR b1 2V
air density
Page 5 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
Page 6 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
Page 7 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
Page 8 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 19
CONTROL SYSTEMS
1 INTRODUCTION
1.1 Basic safety requirements for powered flying controls are specified in Part 1,
Section 4, Clause 4.10 This Leaflet deals with the strength of the powered flying control
circuit, mounting the control unit into the aeroplane and the control characteristics when the
controls are operated in conjunction with an automatic stabilisation device.
2.1 The design load for a mechanical input circuit of a powered flying control system is
that applied by the pilot in the presence of a feed-back or artificial feel mechanism: the
powered control unit valve(s) input may, therefore, be loaded by the pilot's maximum effort,
less whatever is absorbed by the feel mechanism. Hence there is no significant difference
between a mechanical input circuit of a powered control and a manual control circuit. This
part of the circuit should meet the strength requirements of Clause 3.8 which call for an
ultimate factor of at least 1.5 on the maximum effort which a pilot is capable of applying,
with a jam of the powered control unit input. The pilot's nominal effort, however, is only a
small fraction of his attainable maximum and so this part of the circuit will usually meet the
fatigue strength requirements of Clause 3.12 (see Para 2.4 below), when designed to the
strength requirements of Clause 3.8 Reaction of the maximum effort in both directions
should be assumed at all possible parts of the circuit including an extreme position at the
power unit. As this loading implies a failure of the powered control unit or its input valve to
respond to normal effort, its application must be considered in both directions.
2.2 Account should be taken of loads that may be induced in the parked condition when
surfaces may deflect downwards under weight or upwards and downwards due to wind
loads. Circuit loads will depend on the location of input circuit stops, possibly artificial feel
forces, and damping devices if fitted.
2.3 The design load for the output circuit should be the maximum load which the power
unit is capable of applying increased by the pilot's effort through the feed-back or artificial
feel mechanism where this is possible. The contribution of the powered control unit is,
however, much larger than that due to the pilot and in general the working load in this
circuit is much closer to the maximum attainable load than is the case in the input circuit.
Consideration should also be given to conditions arising in the circuit during and after
change-over to any emergency system of control. High transient peak loads may occur
when the system is brought to rest rapidly e.g., by sudden closure of a valve after
movement at maximum velocity, owing to the kinetic energy to be dissipated in a small
displacement. It is important therefore in hydraulic systems to stress for the loads then
arising as these may be greater than those due to the maximum working pressure.
Care should also be taken to avoid stress concentrations; testing to an approval schedule
will be necessary. The maximum load may be greater than that needed for aeroplane
control: for instance, each part of a duplicated power system may be capable of three
quarters of all of the necessary force.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.4 The fatigue strength of mechanical powered flying control circuits should be
established in accordance with the principle outlined in Clause 3.12 (Account should be
taken of the loading actions referred to above).
3.1 When the powered flying controls are to be operated by an automatic flight control
or in conjunction with an auto-stabiliser the control characteristics under powered flying
controls should be satisfactory both with and without the automatic flight control and/or the
auto-stabiliser in operation (see Part 1, Section 4 Clause 4.10).
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 20
CONTROL SYSTEMS
1 INTRODUCTION
1.1 In this Leaflet, information and explanation are given regarding the tests of powered
flying controls called for in Clause 3.9
(a) the powered control unit by itself - part of the Type Approval of the Unit, and
(b) the powered control installation for a particular aeroplane covering the
complete system from pilot's input to the control surface.
2.1 The test rig for the powered control unit should be capable of applying loads to the
output member of the unit varying at the required relationship to the displacement and
should operate the input member through the required duty cycle at any required load,
speed, frequency, and for any desired endurance period. Means should be provided for
measuring all characteristics of the control such as loadings, friction, rates of operation,
elasticity, mass, inertia, resonance frequencies, stability, and temperatures and pressures,
where applicable, at critical points in the control. The need to apply assisting loading and
not merely resistive, when this may occur naturally in flight, should not be overlooked. A
chamber should be available for checking the operation of the unit at high and low
temperatures as required by Part 1, Section 7, Clause 7.2
2.2 In the case of the complete control system the test rig should represent, as closely
as possible, the conditions of the aeroplane in which the powered control is to be flight
tested. All loads should be applied to the actual control surface or its test rig equivalent and
operation of the input member should be through the pilot's input. For check testing of the
complete system when installed in the aeroplane, some form of control surface loading rig
should be available.
3.1.1 During the development stages of a new design of control unit some testing
of the prototype unit will be necessary before the unit reaches the test stage of Para
3.2. In order to ensure a reasonable standard of safety in flight, before any flight
testing is done the prototype control unit should be submitted to a Flight Clearance
Test in a ground rig as required by Clause 3.8 This test may be either separate
from, or included in, the Flight Clearance Test of the installation in which the unit is
to be flight tested, described in Para 4.1 below. An endurance test, followed by an
inspection and performance check, should be carried out before flight.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
3.2.1 The development of a new design of control unit to the stage when it is
ready for production may involve much rig and flight testing. Before production is
started, however, approval of the type should be obtained and for this purpose a
prototype/typical production unit should be subjected to the Unit Design Clearance
Test called for in Clause 3.8 This should be an extended test covering all aspects of
performance, endurance, stability and functioning of safety devices under all
reasonably foreseeable types of simulated failures. During the test the unit should
demonstrate that it gives its declared performance, that it functions within its
declared temperature conditions and also when subjected to negative 'g' and
sustained inverted flight.
3.2.2 For endurance testing, the input member should be operated through a 'duty
cycle' repeating it continuously for an adequate total running time. The composition
of this duty cycle should be decided on before the start of the test to suit the
particular control and the type of aeroplane for which the control is intended, in
accordance with Leaflet 21.
3.2.3 The aim of the complete test should be to establish that the performance of
the unit is maintained over the specified flying hours between scheduled removals.
Should unit design clearance - testing be included in Installation Design Clearance
testing only sufficient testing to give confidence for development flying need be
done.
3.2.4 No modifications or major adjustments should be made during the test and
on completion there should be a complete strip and examination of the unit. The unit
will not be deemed to have passed the test unless the strip and examination is
satisfactory, the standard to be that its condition is such that the unit would still be
satisfactory for safe use in the aeroplane.
4.1.1 Before any powered control unit is flight tested either in an established type
of aeroplane or in a new prototype, the complete powered control system, including
control runs and/or circuitry and control surface actuators is required by Clause 3.8
to undergo a Flight Clearance Test, partly in a ground test rig and partly when
installed in the aeroplane in which flight tests are to be made.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
4.1.2 The rig test should be carried out on a prototype of the complete system to a
test schedule which should be representative of the test flying intended. The test
schedule should call for an endurance period but this need not be extensive if the
unit is already cleared for flight since the aeroplane controls themselves do not
usually require long endurance testing. If the control unit is not already cleared for
flight by either of the tests at Para's 3.1 or 3.2 adequate endurance testing should
be included in the schedule. All other characteristics of the system such as
performance, stability, resonance and functioning of safety devices under all
required conditions should be thoroughly investigated. (For resonance testing, see
Part 1, Section 4, Clause 4.8 and Section 4 Leaflet 30).
4.1.3 Tests should be made to demonstrate that the safety requirements of Clause
4.10 are met and that the operation or malfunctioning of other powered systems in
the aeroplane has no adverse effects on the powered control system under both
normal and emergency conditions.
4.1.4 The aeroplane ground test may be made on the actual installation which is
to be flight tested. This installation should be identical with that tested in the rig. If
both rig and aeroplane ground tests are satisfactory and meet the standards
declared in the pre-test declarations, flight testing may proceed.
4.1.5 It is possible that the rig test may reveal that flying time should be restricted
until further ground tests have proved the reliability of the system.
4.2.2 In cases where the control unit employed in the installation has not already
been tested and approved separately, the Installation Design Clearance Test may
be extended to cover the Control Unit Design Clearance Test of Para 3.2
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 21
CONTROL SYSTEMS
1 INTRODUCTION
1.1 In this Leaflet, specimen test schedules for the testing of powered flying controls, as
called for in Clause 3.8 and as described in Leaflet 20, are given in detail. These schedules
are intended to show the general form that the testing procedure should take and should be
modified where necessary to suit any particular design of control.
2 PROTOTYPE CONTROL UNIT - FLIGHT CLEARANCE TEST (see Leaflet 20, Para 3.1)
2.1.1 The following, each at the appropriate power supply conditions, need to be
decided before commencement of the test:
(a) the composition of the "duty cycle" appropriate to each portion of the
test,
(b) the maximum and minimum rates of operation with the corresponding
maximum loads at the control unit output member, and the maximum rate at
zero load,
(c) the maximum loads at the control unit output with the corresponding
maximum rate, and
2.2.2 Every 5 hours approximately, carry out full cycles of the input member (full
cycle comprises movement from the neutral to one extreme position, return through
neutral to opposite extreme and return to neutral) as follows:
Page 1 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
(d) operate safety device and allow 10 full cycles in the emergency
condition. Where a means of change-over from emergency control back to
normal is provided, check that this operates satisfactorily.
Note: In special cases where the change-over from the duty cycle to the full
cycle involves undue time and labour, the total of 200 full cycles may all be
carried out at the end of the endurance test.
2.2.3 The temperature rise at critical points in the control unit, (e.g., electric motor
cooling air, fluid in hydraulic pumps, motors and jacks, etc.), should be recorded
during the test.
2.2.4 At some suitable time during the test, check under low temperature
conditions sufficient to clear the unit for the experimental flying intended and under
negative 'g' conditions sufficient to clear the unit for operation under negative 'g' or
inverted flight as applicable.
3.1.1 The following, each at the appropriate power supply conditions, need to be
decided before commencement of the test:
(a) the composition of the "duty cycle" appropriate to each portion of the
test,
(b) the maximum and minimum rates of operation with the corresponding
maximum loads at the control unit output member, and the maximum rate at
zero load,
(c) the maximum loads at the control unit output with the corresponding
maximum rate, and
Page 2 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
3.2.1 This test, comprising about 20 cycles at varying loads and rates, should
precede all performance and endurance testing. Checks on the operation of the
safety device should be made by simulating a failure in various positions of the
control output member allowing a number of cycles in the emergency condition in
each case. This should be repeated for all agreed possible types of failure including
runaway, freeing or seizure of the control. Where a means of change-over from
emergency control back to normal is provided, this should be checked in each case.
The time to change over to the emergency condition, and back again, when
appropriate, should be measured in each case.
3.3.1 This test should precede the endurance test in order to establish the
performance of the unit in the new condition. It should include the following:
3.3.2 The above should be repeated over a range of power supply conditions from
the minimum emergency conditions to the maximum.
3.4.2 Every 10 hours approximately, carry out full cycles of the input member (full
cycle comprises movement from neutral to one extreme position, return through
neutral to the opposite extreme and return to neutral) as follows:
Page 3 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
Note: In special cases where the change-over from the duty cycle to the full
cycle involves undue time and labour, the 10 hour interval between the sets
of full cycles may be increased to not more than 50 hours with the number of
cycles at each interval adjusted to make the same total of 800.
3.4.3 The running times should be arranged to include at least one continuous
endurance period equivalent to the endurance of the particular aeroplane for which
the unit is designed, or in the case of a unit not associated with a particular
aeroplane, a period of twelve hours.
3.4.4 The temperature rise at the critical points in the control unit, (e.g., electric
motor cooling air, fluid in hydraulic jacks, etc.), should be recorded during the test.
3.4.5 At some suitable time during the test, check under the temperature
conditions of Clause 7.2, as appropriate to the particular case under consideration.
Record time taken to reach satisfactory functioning condition from starting. Check
safety device.
3.5.1 Repeat preliminary performance check at Para 3.3 and note changes.
4.1.1 The following, each at the appropriate power supply conditions, need to be
decided before commencement of the tests:
(a) the composition of the duty cycle appropriate to each portion of the
test,
(b) the maximum and minimum rates of operation with the corresponding
maximum loads at the control surface, and the maximum rate at zero load,
(c) the maximum loads at the control surface and the corresponding
maximum rate, and
Page 4 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
4.2.1 The complete powered control system should be installed in the ground test
rig under conditions similar to the installation in the test aeroplane. About 20 duty
cycles at varying loads and rates should be made. Checks on the operation of the
safety device should be made by simulating a failure in various positions of the
control surface allowing a number of cycles in the emergency condition of each
case. This should be repeated for all agreed possible types of failure including
runaway, freeing or seizure of the control. Where a means of change-over from
emergency control back to normal is provided, this should be checked in each case.
The time to change over to the emergency condition and back again, when
appropriate, should be measured in each case.
4.3.1 This test should be made with the complete control system in the test rig and
should precede the endurance test in order to establish the performance of the
system in the new condition. It should include:
(e) resonance tests (see Part 1, Section 4, Leaflet 40) applied to the test
rig.
4.3.2 The above should be repeated over a range of power supply conditions from
the minimum emergency conditions to the maximum, including that appropriate to
the all engine failure case. The tests using the emergency power supply appropriate
to the all engine failure case should check that the performance is adequate and
that the system will function for a sufficient length of time to meet the requirements
of Part 1, Section 1, Clause 1.1 appropriate to the particular aeroplane concerned.
4.3.3 Tests should be made to demonstrate that the safety requirements are met
and that operation of all other powered systems supplied from the same power
source, does not materially affect the performance of the powered control.
4.4.1 With the system in the test rig and with the power supply, if engine driven, at
cruising conditions, operate the pilot's control through the declared endurance duty
cycle for a total running time of:
Page 5 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
(b) 25 hours with the full cycles of Para 2.2 if the control unit is not
already cleared for flight.
4.4.2 Carry out 600 full cycles of the pilot's control (full cycle defined at Para 3.4.2)
as follows:
4.4.3 After every 50 cycles and at same load and speed, operate safety device as
in Para 4.2.1 and allow five cycles in the emergency conditions in each case.
4.5.1 Repeat preliminary performance check at Para 4.3 and note changes.
4.6.1 Strip and examine all parts not already tested as part of the control unit.
4.7.1 Install a complete powered control system identical to that tested as above,
in the test aeroplane, and carry out the performance tests of Para 4.3 as far as
practicable on the ground using a control surface loading device.
4.7.2 Carry out the resonance tests (see Part 1, Section 4 Leaflet 30).
Page 6 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
5.4.1 With the system in the test rig and with power supply, if engine driven, at
cruising conditions, operate the pilot's control through the declared endurance duty
cycle for a total running time of:
(a) 50 hours if the control unit employed is a type tested and approved
unit, or
(b) 200 hours if the control unit employed is not a type tested and
approved unit.
5.4.2 Every 10 hours carry out full cycles (full cycle defined at Para 3.4.2) making
a total of 3000 cycles, as follows:
(b) in a 200 hour test, the number of full cycles at (a), (b) and (c) above
should be reduced to 50 so that the same total of 3000 cycles is achieved in
the complete test. Alternatively, the time interval between the sets of full
cycles may be increased from 10 to a maximum of 50 hours with the number
of full cycles in each set adjusted accordingly.
5.4.3 After every 50 full cycles and at same load and rate, operate safety device
as at Para 4.2 and allow five cycles in the emergency condition in each case. When
a means of change-over from emergency control back to normal is provided, this
should be operated in each case.
Page 7 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
5.7.1 Install the same prototype powered control system, or an identical one, in
either a prototype or a production version of the aeroplane for which the system is
intended, and carry out the performance tests of Para 4.3 up to about 30%
maximum load, using a control surface loading device, and check results with the rig
test results.
5.7.2 Carry out resonance tests (see Part 1, Section 4 Leaflet 30).
Page 8 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 22
CONTROL SYSTEMS
1 INTRODUCTION
1.1 The purpose of this Leaflet is to draw attention to certain airworthiness aspects
which should be considered when bob-weights or similar mass elements are incorporated
in conjunction with springs in flying control systems.
1.2 Cases have occurred in the past where such devices have been introduced into
flying control circuits in order to improve the stick force/g characteristics or to achieve some
measure of g restriction. A device of this type is sensitive to the normal acceleration, the
angular acceleration in pitch, the longitudinal acceleration or any other response quantity of
the disturbed longitudinal motion of the aeroplane, or to a combination of such quantities; in
turn it moves the pitch control surface and so controls the normal acceleration of the
aeroplane.
1.3 It may sometimes happen that an element such as an unbalanced link may be
unintentionally included in a control system in such a way as to act as a bob-weight and so
introduce a coupling between pitching of the aeroplane and motion of the control circuit.
Such a case should be treated as if the mass-spring unit had been deliberately included,
and the following recommendations should be taken as applicable.
2 RECOMMENDATIONS
2.1 In all cases where a mass spring system of the general type described in Para 1.2 is
to be installed, a full analytical investigation should be made of the stability of the
aeroplane-control system combination.
2.2 For such investigations the following points should be borne in mind:
(a) in cases where power units are incorporated in pitch control circuits, the
dynamic properties of these units (i.e., their transfer functions and time constants)
should be taken into account,
(b) dry friction in the elevator control circuit may adversely affect the stability of
the g-restrictor-aeroplane combination, and should therefore be also taken into
account,
(c) neither the numerical value of the ratio of the natural frequency of the g-
restrictor unit to that of the short period mode of the aeroplane, nor the stability
characteristics of the g-restrictor and the aeroplane taken separately give any prior
indication as to the behaviour of the combined system, and
(d) the forward speed may be an important parameter, and its whole practical
range should be covered in the analytical investigation.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.3 The above recommendations apply also to any devices having dynamic properties
similar to those of the g-restrictors considered, and in particular to devices intended to
improve stick force/g characteristics.
2.4 Some g-restrictors of the general type considered, though efficient and otherwise
satisfactory, may induce excessive control surface loads. The magnitude of these loads
should be assessed by appropriate response calculations embracing both the initiation of
an unduly severe manoeuvre and the checking movement of the control surface as induced
by the g-restrictor. For this part of the investigation R.A.E. Technical Note No. SME 349
may be of assistance.
2.5 The effect of accelerations on unbalanced control circuits controlling other axes
should also be considered.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 23
CONTROL SYSTEMS
1 INTRODUCTION
1.1 This Leaflet contains details of the fatigue testing of hydraulic powered flying
controls required by Clause 3.8 The tests aim to ensure that hydraulic powered flying
controls have adequate fatigue life under conditions of varying pressure stress to which
they will be subjected in service. Included in the tests are screw jacks driven by hydraulic
motors and, where applicable, automatic flight control system actuators and artificial feel
simulators.
2.1 The duty cycle for the control surfaces should be specified by the airframe designer
who should state clearly whether the additional factor of Leaflet 34 Para. 6 has been taken
into account. In the absence of such a specification the standard duty cycle in Table 1
should be used, in which case the factor specified in Leaflet 34 Para. 6 is not required.
Tests should then be applied in accordance with Para's 3 and 4 below.
2.2 Due to the wide variation between the performance characteristics of various types
of hydraulic powered flying control, standardisation of test conditions is based on flight duty
cycles as applied to the control surfaces rather than the fluid pressures in the hydraulic
powered control circuits. In the case of small amplitude high frequency control surface
movements, the internal jack pressures in a throttled valve servo will be entirely different
from the pressures in a variable delivery servo pump. Similarly, for identical output
conditions, the internal pressure will depend upon the precise nature of valve overlaps or
underlaps, and in duplicated jacks, upon the inter-jack mechanical stiffness.
2.3 In addition to the standard duty cycle, which is based on aerodynamic hinge
moment, fatigue loading on a powered control component may arise from other sources,
and in many cases these may constitute the heaviest fatigue loading. These sources
include ground testing cases, operation of other surfaces both on the ground and in flight,
pump ripple and parking locks. The conditions controlling the above factors must be given
full consideration in addition to the effects of the standard duty cycles.
3.1 Although the external loads applied to hydraulic powered flying control jacks will be
broadly similar to those applied to the local aeroplane structure, the internal jack pressures
will be determined by other considerations. In each case, therefore, it will be necessary to
determine experimentally the internal pressure fluctuations at critical points throughout the
powered flying control system resulting from the duty cycles. The test should be carried out
on a test rig representing as closely as possible the aeroplane installation and special
instrumentation may be necessary at this stage to ensure that local peak pressures are
recorded.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
3.2 Factors influencing the pressure variations are given below, but it is emphasised
that this list does not necessarily include all the parameters significant to the performance
of a powered flying control:
(e) phasing of control valve ports (the smallest variation in port position can
cause large pressure differences particularly in duplicated jacks, thus the magnitude
of pressures applied in fatigue testing should be adjusted to include the most
adverse conditions),
4 METHOD OF TEST
4.1 The tests should be carried out under any one or a combination of the following
conditions, or by any other approved method:
(a) By simple pressure pulse applications to units with output members locked
or unlocked according to whether or not the load is transferred through these
members during the particular duty cycle case. The jacks should be stationary,
externally restrained and the output member locked in the position given in Table 1,
and the relevant fluid pressures applied to the various pressure chambers. This
method offers the advantages of allowing greater acceleration of the test cycle and
reduction of test time (this can be lengthy when, for instance, clearance for 1000
hours is being sought and a scatter factor of 5 is to be applied).
(b) By stroking the units in the normal manner (i.e., by an extension of the
normal endurance tests called for in Leaflet 20) with the jacks moving against
externally applied loads. In this condition the normal supply pressure should be
provided at the powered control inlet and the internal jack pressure win be
developed automatically. When using this method, any reduction in the test time by
acceleration of the flight duty cycle is liable to modify the internal jack pressures and
due allowance should be made for this.
(c) With the jack and fittings uncoupled, in particular cases where the simulated
internal pressure conditions could impose loads on the external attachments of the
jacks in excess of those experienced in flight. In this condition, pressures on
opposite sides of the jack piston should be suitably balanced.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
5 TEST FACTORS
5.1 Fatigue test factors should be derived in accordance with Clause 3.12
TABLE 1
Note: For control systems without autostabilisation the jack stroke may be insignificant in
the standard duty cycle at 10% hinge moment.
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 24
CONTROL SYSTEMS
DESIGN RECOMMENDATIONS
1 INTRODUCTION
1.1 The leaflet presents supplementary information with respect to the design of
mechanical components in control systems. The aim is to amplify the requirements of Clause
3.8 for particular detail design considerations. In many cases the adoption of the
recommendations of this leaflet will be found to be necessary to meet the specified
requirements.
2.1 GENERAL
2.1.1 Clause 3.9 requires that a control system shall be designed to tolerate the
possibility of jamming, chafing and interference from personnel, cargo, loose objects and
freezing of moisture. These overall considerations should be applied to each detail
component.
2.2.1 Each control run should be arranged with adequate local clearances to minimise the
possibility of jamming by loose objects. Control runs should not be too close to horizontal
surfaces, or have levers designed to operate in local pockets. Multiple levers on common
spindles should not have lightening holes, or inward facing flanges. Chains with sprockets
having a horizontal pivot are undesirable and pulley guards should be sufficiently close to
prevent the ingress of small parts.
(a) 3 mm between elements which move in relation to one another but which are
guided or connected to the same component.
(b) 6 mm between elements which move in relation to one another and which are
guided or connected to separate components.
(c) 12 mm between elements and aeroplane structure or equipment to which they are
not attached, unless structural flexibility requires a greater clearance to be provided.
2.3 JOINTS
2.3.1 All adjoining parts should be secured in a manner that will prevent loosening when
subjected to internal, external and vibration loads. All pins, etc., subjected to load or motion
should be positively secured and locked using slotted nuts and split pins together with
additional means of retention where possible. Clevis pins retained only by split pins are not
acceptable.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
2.3.2 Control system joints in push-pull systems that are subjected to angular motion,
except those in ball and roller bearing systems, should have an ultimate factor of safety of
not less than 3.33 with respect to the ultimate bearing strength of the softest material used as
a bearing. The approved ratings of ball or roller bearings may not be exceeded.
2.4.1 If it is assumed that any failure or disconnection is extremely remote then the design
should be such that no failure or disconnection could be foreseen under practical
circumstances, including errors of operation, assembly or maintenance, for such an
assumption to be acceptable. The following considerations should be given:
(a) Choice of materials to avoid undue notch sensitivity or stress corrosion cracking.
(b) Adequate robustness to cater for all conditions arising, including those due to errors
in operation, assembly or maintenance.
(c) Avoidance of design features which tend to give rise to fatigue effects, including
sensitivity to vibration.
2.5.1 The choice of pulley diameter for use with cables should be such as to preclude the
possibility of fatigue arising from bending of the cable and local rubbing together of individual
wires in the cable.
2.5.2 Pulleys should not be arranged so that a reverse bend in the cable arises as it
traverses from one pulley to another, and the pulleys should be in the plane of the cables so
that the cable does not rub against the pulley flange.
2.5.3 The wrap angle of the cable round the pulley should be adequate and pulleys and
sprocket should be guarded to prevent any cable or chain jamming or coming off when slack
and re-engaging correctly after slack has been taken up.
2.5.4 Each kind and size of pulley should correspond to the cable with which it is used, to
ensure that the minimum bend radius of the cable is squalled or exceeded.
2.5.5 Fairleads should be installed so that they do not cause a change in cable direction.
2.5.6 Where turnbuckles are attached to parts having angular motion they should be
arranged so that fouling is positively prevented throughout the range of travel.
2.6.1 Bearing installations should be arranged in such a manner that failure of the rollers
or bags will not result in a complete separation of the control. Where direct axial control force
application cannot be avoided a failsafe feature should be provided.
2.7.1 The tubes used in push-pull and torque control systems should be adequately
vented and drained, or completely sealed, to prevent the accumulation of condensed
vapours.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
2.7.2 Joints incorporated in torque tube control systems to enable axial movement should
have sufficient engagement to ensure that disengagement will not occur and positive drive
will be retained under the most adverse set of manufacturing and installation tolerances and
structural flexibility. Minimum engagement should not be less than one diameter. A means
for inspection of the amount of engagement should be provided.
2.7.3 All torque tubes should be mounted on anti-friction bearings with supported
couplers spaced at close enough intervals with sufficient misalignment capability to prevent
undesirable bending or whipping of the tubes.
2.7.4 Each torque tube in a link system should be removable and re-installable in the
aeroplane without having to disturb the support, component, or other interfacing system
elements at either end of the tube.
2.7.5 The rated operating speed of a torque tube system should be no greater than 75%
of the critical speed unless a supercritical design has been agreed with the Project Team
Leader.
2.7.6 Where a broken tube could cause damage to other components an adequate guard
should be provided.
3 CONSIDERATION OF FAILURES
3.1 The failures to be considered include those of mechanical, hydraulic and electrical
devices in the primary and secondary control systems. Failures of Q-actuated mechanisms
(including devices such as Mach-trim and Q-actuated gearing, if not covered by Clauses 3.9
& 6.5 in Part 1, Section 6) are also to be considered.
3.2 Failures in power operated trimming systems can be particularly serious. Creeping
or runaway can lead to a dangerous situation from which it may be difficult to recover once
the failure has been detected. Evidence that a pilot can override any out-of-trim force is not
sufficient to ensure safety under all conditions, in an instrument approach for example.
3.3 Any kind of interlock between main and standby trimmers should be avoided.
Systems which, when the console switch is used, automatically render the main system dead
until it is reset, should not be used.
3.4 When pitot and static pressures are used to operate mechanisms in the flight
control system the consequences of failure of the pitot-static system by blockage or
breakage shall be considered.
3.5 Control system sensors such as probes, vanes and other mechanical devices
should be installed at locations which minimize exposure to conditions which could result in
inaccurate output signals or failures.
3.6 When pilot intervention is necessary to recover from a failure in the control system
or to change to a reversionary mode, a realistic time delay should be allowed for. See Part 1,
Section 4 Leaflet 36.
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 25
1 INTRODUCTION
1.1 Where the aeroplane designer, in order to meet the performance requirements,
accepts reductions in the natural stability and control characteristics to an extent which
demands artificial augmentation in order to restore acceptable levels of airworthiness, then
he must provide such augmentation with integrity at least as high as that achieved with
previous generations of primary control systems.
1.2 Such an aeroplane design, which may be inherently unstable, demands restoration
of stability by means which are viable throughout the full flight envelope. To achieve this,
full control authorities may be required, indeed the absolute limits of control authority may
well need to be re-assessed in the light of the degree of instability.
1.3 To effect both stability augmentation and to provide full flight control to the pilot, it is
desirable to provide an efficient and airworthy integration of these functions in one
continuously active control system.
1.4 Having taken the decision to adopt active controls for a new aeroplane, the designer
is then free to develop interactively the airframe and active control system fully to optimise
performance.
2 APPLICATIONS
2.1.1 An ACS can offer the potential for providing facilities to support a number of
other airframe systems.
2.1.2 If the ACS comprises dedicated motion sensors, then various computed data
may readily be made available for instrument display purposes and as reference
data for other systems.
2.3 Other systems which may be expected to interface with an ACS include:
Page 1 of 1
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 26
SYSTEM REQUIREMENTS
1 INTEGRITY
1.1 Two of the most dominant problems in the development of an ACS are the selection
of the means to achieve a very low probability of catastrophic failure and to demonstrate
that this has been achieved.
1.2 Designers may be asked to aim for a risk of loss of the aeroplane due to ACS failure
in the order of 10-7 per flight hour. For this reason, systems employing redundancy are
necessary.
1.3 Single lanes using electronic control may exhibit failure rates exceeding 1 in 1000
flying hours, it follows that a simple cross-compared redundant system of this quality may
require four independent lanes of control.
1.5 Use of redundancy introduces the prospect of having to transfer data between
otherwise isolated lanes.
1.6 Where such intercommunication is utilised great care should be taken to avoid all
sources of common-mode failure.
1.7 Use of similar redundancy is attractive to the designer because of the ability to
thereby secure good identity of data. It does, however, present another source of common-
mode failure.
1.8 The natural distribution of random defects or failures amongst similar hardware
components results in a low probability of simultaneous failure, but there nevertheless
remains the possibility of an unidentified common susceptibility to undefined influences
arising from the inherent commonality of design and manufacture. For this reason great
care should be exercised in the selection of components for use in redundant systems.
2 SENSOR CONFIGURATIONS
2.1 Conventionally, in redundant systems, there are three choices of primary motion
feedback sensor configurations, which may be summarised as follows:
Measuring axis close to aeroplane body axis and packaged in packs of redundant
similar sensors.
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
Measuring axes close to body axes but with a group of, say three axis sensors in
each set, one set per lane.
2.2 The disadvantages of the skewed axis system lie in the extra computing required to
resolve the skewed sensor signals into the required axis and that necessary to handle the
failure management, together with the architectural complications associated with this
computing and its power supplies.
2.3 A trade-off exists between the minimum number of sensors and the magnitude of
failure transients and nuisance warnings.
2.4 There are inevitable compromises which arise from the co-located sensors and the
fact that there may need to be at least two redundant sets located apart.
2.5 Each sensor and each set may sense variable mixes of structural mode responses,
the eradication or resolving of which demands special attention.
2.6 The optimisation of management of a skewed axis sensor configuration can lead to
this being implemented as a discrete sub-system with a different level of redundancy to that
of the ACS proper. In this case the choice of method by which essential sensor data is
transferred to the control computers needs special care, in order to ensure that the required
levels of integrity are met.
3 AIR-DATA SENSORS
3.1 Air-data sensors (pitot-static probes and airstream direction sensors) present
special difficulties. The special separation needed to ensure that two or more sensors are
not damaged by the same event (e.g. bird strike) causes inevitable differences between
their signals, especially at low speeds or at extreme attitudes. To avoid unacceptably
frequent rejection of their signals when no fault has occurred, cross-monitoring has to be
compromised with the risk that some faults remain undetected.
3.2 In some systems (e.g. Q-feel) the air-data signal only affects the characteristics of
the inceptor in some way and does not affect the position of the motivator. In others the
signal is fed into the system and automatically applies motivator movement. Both cases
should be considered, but for the second type some form of "fail safe" in the design is
essential.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
4 CONTROL AUTHORITIES
(a) Motivator (or actuator) authority, which normally represents the maximum
usable geometric output.
(b) Mode authority, or internal signal authority. This is a flexible parameter which
may be fixed or variable, and which is usually established with an integrity of similar
order to that of the system itself.
4.2 Finite motivator authorities have a profound effect upon the behaviour of an
aeroplane which relies upon an ACS to provide essential performance. The choice of
displacement and rate authority for a given function has to be made in the light of both
specified performance and malfunction requirements and the nature of circumstances
expected to be met at and beyond control saturation.
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 27
1 MATHEMATICAL MODEL
1.1 In order that the overall aeroplane performance objectives may be addressed, it is
necessary to acquire a thorough understanding of all the aeroplane and system dynamics.
From these a mathematical model of the total aeroplane and system should be built up.
1.2 This model should be as complete as possible. Airframe structural modes and the
effects of variable mass configuration such as disposable and consumable stores, fuel and
its dynamics should be considered together with detailed non-linearities and time
dependent phenomena such as transport and iteration delays and quantisation effects.
1.3 The efficient design of control laws will depend upon the accuracy of the model.
2 COMMONALITY OF SOFTWARE
2.1 If the designer chooses to utilise one common software interpretation of control laws
for use in all lanes of a redundant system, then he should recognise the implications of this
in respect of common-mode failure susceptibility.
2.2 This element of commonality is difficult to assess in terms of reliability and attention
is drawn to various methods of establishing confidence in the design of such software. Ref
Interim Def Stan 00-31 (Obsolescent).
2.3 Confidence is currently established by ensuring that all software paths and
functional elements are adequately exercised during rig testing.
Page 1 of 1
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 28
1 INTRODUCTION
1.1 This Leaflet provides guidance on acceptable means of compliance with the
structural requirements of Clause 3.9 and related requirements such as those for static
strength (Clause 3.1), fatigue and damage tolerance (Clause 3.12) and aero-elasticity (Part
1, Section 4, Clause 4.8).
1.2 The procedures used for static and fatigue design of aeroplanes incorporating ACS
are similar to those for conventional aeroplanes. However, critical design cases may be
more difficult to determine, and the design process will necessitate integration of the
procedures used for structural, aerodynamic and active control system design.
2.1 From a structural point of view Active Control Systems (ACS) may be grouped into
two categories;
2.2 ACS in the first category might include, for example, carefree manoeuvring (to
prevent departure from controlled flight), ride improvement, weapon platform improvement,
relaxed stability. manoeuvre enhancement, direct lift control.
2.3 ACS in the second category might include, for example, gust load alleviation,
manoeuvre 'g' limiting, flutter margin improvement, flutter suppression.
2.4 For systems in the first category and, for those systems in the second category
which as a secondary effect enhance performance (e.g., flutter suppression), loading
actions are usually of a different form to those for conventional aeroplanes.
2.5 For ACS aeroplanes it is likely that design requirements alone are insufficient to
define all critical static loading cases. Therefore the intended aeroplane usage should be
probed for potentially critical or hazardous loading actions due to pilot input or particular
gust patterns, separately or in combination. The effects of augmentation system
disconnects, degrades and failures should be evaluated.
2.6 In particular, attention should be given to the flight and ground loads which arise
when the active control system is:
Page 1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
2.7 When identifying those loads which are likely to occur during flight in the degraded
state the designer should consider whether the pilot could, if he was aware of the
degradation, reduce the severity of the loads which are likely to be encountered by taking
appropriate measures such as:
If, to cope with a particular degradation of the ACS, the designer chooses to maintain the
structural integrity of the aeroplane by requiring the pilot to observe a flying limitation, then
the margin between the maximum load predicted to occur following the implementation of
the limitation and the limit load capability of the degraded aeroplane should comply with the
guidance given in Section 2 Leaflet 9.
2.9 ACS aeroplanes which feature either carefree manoeuvring or manoeuvre 'g'
limiting may fly to the extremes of their flight envelope more frequently than conventional
aeroplanes. Providing the ACS is sufficiently reliable and DLL is adequately defined, it is
unlikely that an ACS aeroplane will exceed DLL more frequently than an equivalent
conventional aeroplane in which the pilot must observe flight limitations to prevent
exceedances of DLL- Nevertheless, the use of 'g' limiting system does not guarantee that
critical loads will not be exceeded as it may only limit one of the components of load in a
particular structural item. Consequently, to enable the Service Release levels in respect of
static loads to be assessed, the proposed aeroplane usage should be probed to identify
possible exceedance of DLL and the ease with which DLL can be approached.
2.10 Aeroplanes incorporating ACS are likely to exhibit increased control activity, in
particular more high frequency small amplitude motions, and the designer should pay
special attention to the fatigue design of the actuators, the motivators, the mechanical
interface between the motivators and the actuators, and the associated support structure.
2.11 The ACS should be designed to avoid inertia instability of the control system
occurring when the aeroplane is undergoing system testing on the ground.
2.12 There can be no guarantee that the loads analysis can identify all critical
combination of possible loading actions. Consequently, Clause 3.10 contains requirements
for the in-service confirmation of the load levels deduced from the analysis.
Page 2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
2.13 Where an ACS is used to contain structural loads within limits, the system
capabilities and performance derived from calculation must be substantiated by direct strain
measurement during development flying.
2.14 Special structural implications arise from the application of Active Controls to
combat aeroplanes and transport aeroplanes and these are examined separately in the
following Para.
3 COMBAT AEROPLANES
3.1 Aeroplanes which feature either a carefree manoeuvring or a manoeuvre 'g' limiting
facility will probably experience more frequent medium to high 'g' manoeuvre than
conventional aeroplanes. Therefore, special attention should be given to the definition of
fatigue load spectra.
3.2 The calculated flutter speed of an aeroplane with a fully serviceable ACS should be
at least 1.15V D at any point in its flight envelope for any authorised mass/stores
configuration. Also, an aeroplane with ACS inoperative should be flutter free to V D at any
point in its flight envelope for any store/mass configuration. However, it may be acceptable
to design an aeroplane employing an active flutter control system to be flutter free with that
system inoperative to a speed less than V D when carrying external stores if:
(b) The pilot can be given the necessary warning to enable him to perform the
actions detailed in Para 3.2(a).
4 TRANSPORT AEROPLANES
4.1 It is permissible, providing the requirements specified in Clause 3.10 are met, to
take full advantage of the amount of load alleviation provided by a Gust Alleviation System
when designing the structure to the loads corresponding to the gust cases specified in
Clause 3.4.
4.2 The use of load alleviation systems to achieve lighter and smaller structures may
result in reduced structural stiffness and reduced flutter stability. Consequently, ACS
aeroplanes may feature unconventional sizing of structural items, for example wing skin
thicknesses may be greater at the tip than at the root to achieve maximum torsional
stiffness or vary across the chord to achieve an optimum position of the flexural axis.
4.3 Notwithstanding the use of flutter suppression systems, the flutter speed with the
flutter suppression system inoperative should never be less than V D .
Page 3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 29
1 INTRODUCTION
1.1 Experience has shown that agile military aeroplanes equipped with ACS may be
subject to increased manoeuvre frequency and magnitude in service, compared with
conventionally controlled combat aeroplanes. It is recognised that if high levels of
manoeuvring are frequent then reduced airframe life will ensue, if not allowed for in the
design.
(a) Design for maximum performance and allow for airframe life consumption by
investing in a more resistant airframe (at possible weight cost).
(b) Design for two levels of performance; maximum performance which could be
termed 'Wartime' Performance and a lower level, perhaps termed 'Peace-time'
Performance, either of which could be selected for use by the pilot. Carefully
regulated service use of these alternatives could ensure a reasonable rate of
consumption of airframe life.
The latter option has system design and training implications in that a lower level
performance has to be defined and appropriate mode selection provided, with possibly
instinctive override available to the higher level of performance.
1.3.1 Redundant Systems with voting incur a finite probability of nuisance failure
warnings, although with careful design these may be minimised. When apparently
'nuisance' failures occur the conditions leading to the event can, if known, enable
confirmation and/or rectification to be carried out to avoid repetition and
investigatory loss of aeroplane availability. It is therefore desirable for context data
to be recorded for each 'failure' incident.
1.3.2 Experience has shown that incidents of this nature prove difficult to resolve if
no specific information, other than aircrew reports, is available.
- Failed item
- Relative errors (redundant items)
- Control system status
- Dynamic data (3 axes)
- Flight case (speed, height, mach no.)
Page 1 of 1
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 30
1 INTRODUCTION
1.1 In order to develop the ACS to flight standard, a rig is required in which the
complete system hardware can be fully functioned. A comprehensive real-time flying
simulation which permits all operational performance to be assessed should form part of
the rig.
1.2 The rig simulation should also be utilised to enable test pilot assessments to be
carried out with particular reference to system management including the effects of failures.
The tests should also allow impromptu pilot investigations which explore off-design cases
and any other points of curiosity.
1.3 This facility should provide the basis for system performance proving prior to first
flight. A detailed schedule of failure cases should be assessed.
1.4 This testing should precede installed system testing in the aeroplane and enable
installed testing to be restricted to essential items.
1.5 Endurance testing on this rig may be restricted to whole system operating with a
pilot in the loop or may be in part automated. Sequences of start-up, pre take-off checks,
taxi, take-off, climb, through representative flight conditions and manoeuvres covering all
operational roles etc., to return to circuit, landing, post-flight checks and shut-down should
be included.
1.6 These tests are all to be completed to an agreed equivalent number of flying hours
before first flight.
1.7 The required number of hours before first flight are to be conducted with full flight
standard ACS equipment and are expected to exhibit defect free operation.
1.8 Subsequent to first flight, rig tests should continue to back-up flight development.
1.9 A separate life test programme is necessary for individual ACS items and these life
tests are conducted by the suppliers prior to delivery for installation in the aeroplane.
Page 1 of 1
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 31
1 INTRODUCTION
1.1 The provisions of Clause 3.9 are specifically for those control systems where correct
operation of the system is essential for continued safe flight, so called 'full-time' systems.
(Refer to Part 0 Definitions). A 'full-time' system may be understood to include those
systems where there is automatic reversion to an alternative active mode in appropriate
circumstances. When an active control system is employed in such a way that continued
safe flight is possible after an incorrect operation of the system by reversion to some form
of direct control by the pilot it may be referred to as a 'part-time' active control system. The
extent to which the provisions of Clause 3.9 then apply will be related to the required flight
envelope and has to be defined in conjunction with the Project Team Leader. This Leaflet is
intended as a guide in making this definition.
(a) Where there is provision for reversion to direct control input, the reversion
mode not being part of the full-time active control system, and resulting in degraded
performance.
(c) A dual control aeroplane where the operation of the ACS is limited to the
inceptors of a single pilot station: the second 'safety pilot' being provided with a
means of controlling the aeroplane which is not affected by failures of the ACS.
2.1 In the case of a part-time ACS system the application of the requirements of Clause
3.9 may be limited to those components of the system which affect overall safety of the
aeroplane.
2.2 In the event of an incorrect operation of the ACS being detected there should be an
automatic reversion to the back-up mode, and there should be an indication to the pilot(s)
of the reversion. The requirements of Clause 3.9 therefore apply to those components
which sense the failure condition and initiate reversion to the back-up mode.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.3 Whilst the ACS system is in operation, including any incorrect operations and
reversion to the back-up mode, it should not introduce any situation which is likely to put
safety of the aeroplane in jeopardy.
3.1 In general, some aspects of all the main paragraphs of Clause 3.9 may be
applicable to part-time ACS.
3.2 Specific aspects of Clause 3.5 which may need to be considered are:
(a) Reversion.
(c) Nuisance reversion to the back-up mode, and especially failure transients.
(e) Control laws and software generally insofar they affect safe flight whilst the
ACS is in operation, and during reversion to the back-up mode.
(g) Operational and piloting aspects generally in as much as the continued safe
operation in the system is involved.
(h) Aeroplane and system proving, again related to those components which
ensure continued safe flight.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 32
TOWING
1 INTRODUCTION
1.1 Minimum mandatory strength requirements for towing are given in Clause 3.11 This
leaflet gives background information and recommendations relating to those requirements.
1.2 Design, operational, and standardisation requirements are given for towing in Part
1, Section 4, Clause 4.4 Note particularly that load-limiting devices are not allowed on tow
bars provided for use on board ship.
2 GENERAL
(a) towing from a point at or near a main or nose unit which happens to be on a
surface where the co-efficient of friction is nearly zero (as for spilt fuel or wet mud
on a firm surface) and where the only reaction is supplied by the inertia of the
aeroplane,
2.2 The applied towing force would, in practice, be balanced by a combination of some
or all of the following:
2.3 So far as possible, the need for detailed consideration of all these possibilities on
every aeroplane has been eliminated by specifying detailed stressing cases which, while
not representing accurately any one particular condition, will normally cover all reasonable
conditions except where the aeroplane geometry is unconventional.
2.4 It is not intended that the requirements should be used to design major parts of the
undercarriage or structure of the aeroplane if satisfactory towing procedures can be
devised which will provide for all normal aeroplane movements and also the extreme cases
described in Para 2.1 It is the Aeroplane Designer's responsibility to agree these
procedures with the Project Team Leader and include them in the Operating Manual.
2.5 The tyre and shock absorber deflections may usually be assumed to be those for
the aeroplane standing on level ground at maximum design take-off mass (MT) and
associated CG position. However, appropriate allowances should be made if there are
considerable geometrical changes under towing loads.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.6 Where towing backwards by attachment of a bridle to other strong points (e.g. the
arresting hook or tail bumper) is required the effect of snatch loads on the undercarriage
should be considered.
3.1 The design loads illustrated for the main-undercarriage towing points in Section 3
Fig. 5 are arranged to provide for the use of a towing bridle on one main-undercarriage unit.
4.1 Two different nose-wheel towing attachment methods are represented in Section 3
Fig. 6. Fig. 6a illustrates the case when the towing arm, wire, chain, or rope, is attached to
a strong point on a castoring nose unit. Fig 6b illustrates the case when a towing arm is
attached to the nose wheel such that towing and steering occur at the same time. The
applied loads will be opposed by side, drag and torsional tyre forces.
4.2 Consideration should be given to variations in these loads and forces arising from:
(d) Practical combinations of towing speeds and swivel rates particularly the
case of swivelling with zero forward speed.
4.3 It is the designer's responsibility to ensure that the material chosen for the towing
points specified in Part 1, Section 4, Clause 4.4 is such that the strength requirements of
Clause 3.10 are met.
5 SHIPBORNE AEROPLANES
5.1 Shipborne aeroplanes are not allowed to have a relief device in the tow bar (see
Part 1, Section 4, Clause 4.4) and must be stressed for the maximum loads expected in the
worst sea state for which towing is required. A dynamic analysis is therefore required
5.2 For preliminary stressing a tow force of 0.8G.MT should be used to represent
sudden braking of the aeroplane or tractor on a non-slip surface and the nose wheel
vertical load may be calculated from:
where f is the moment arm of the braking force about the aeroplane centre of gravity. Then
f = h for tractor braking with nose-wheel tow at the nose-wheel axle and f = H for main-
wheel braking of the aeroplane with any form of towing.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 33
PICKETING
1 INTRODUCTION
1.1 This leaflet gives recommendations relating to the requirements of Clause 3.11
1.2 Chocks are normally used during picketing but their effectiveness cannot be relied
on and so no account is to be taken of any contribution they may make. Brakes are not
used during picketing.
2 FORCES TO BE CONSIDERED
2.1.1 Wherever possible test data is to be used for calculating the aerodynamic lift
on the aeroplane.
2.1.2 If no test data is available it is acceptable to use the normal aerodynamic lift
formula provided that the value of the angle of attack used is calculated as follows:
2.1.3 If the calculated angle of attack exceeds the stalling angle then the stalling
angle is to be used instead.
2.1.4 The formula in Para 2.1.2 assumes that any high-lift devices remain fully
retracted.
2.2.1 During picketing, side and drag forces due to winds will be present on the
aeroplane. The effects of loads due to ship's motion must also be considered. For
calculations the basic mass of the aeroplane should be used together with a range
of co-efficients of friction from 0.1 to 0.8
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.3.1 The most adverse combination of ship motion and wind forces should be
considered. It may be assumed that provided the flight deck securing requirements
are met, then the hangar case will be automatically met, since there are no wind
forces and generally less ship motion in the hangar. The values of ship motion and
wind forces contained in Clause 3.12 are considered to be acting in the most
adverse weather conditions. It is recommended that the force imposed on secured
aeroplanes by the relative wind-over-deck, be added to those resulting from ship
motions.
3 PICKETING POINTS
3.1 To meet the requirements of Clause 3.12, it is advisable to have picketing points on
the aeroplane structure immediately inboard and outboard of each main undercarriage unit.
Where this presents difficulties a single point at each main unit is acceptable provided that
picketing can still be achieved in accordance with the picketing plan. When the
undercarriage track is narrow, and if other considerations such as the direction of
undercarriage retraction permit, single points should be outboard of the main wheels.
3.2 Securing points should be either built in to the structure of the aeroplane or the
undercarriage or take the form of detachable screw-in or clamp-type fittings.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 34
FATIGUE
MATERIAL SELECTION
1 INTRODUCTION
1.1 This leaflet provides guidance on the selection of metallic materials for aircraft
structural components where fatigue life and residual strength are major considerations.
Guidance is given on fatigue performance, fracture toughness, resistance to impact and
crack growth behaviour. Special considerations relating to fibre-composite components are
discussed in Leaflet 40.
1.2 Sufficient material must have been tested to demonstrate that the properties used
for design reflect variability both within and between batches, see for example Def Stan
00-932 (Cancelled – replaced by EDSU 932 Metallic Materials Data Handbook). Otherwise a
programme of testing will be required to establish material behaviour or to confirm assumed
or estimated properties.
2 FATIGUE PERFORMANCE
2.2 General. The extent to which the intrinsic fatigue performance of a material can be
realised is strongly dependent upon the influence of stress-concentrating features such as
notches, cut-outs and fillet radii, and the extent to which crack initiation is accelerated by the
effects of fretting, such as may occur in joints. Fatigue performance is also affected in
varying degree by material quality, that is the presence in a material of inclusions, residual
stresses and discontinuities introduced by material processing and heat treatment, and the
surface quality, that is the introduction of imperfections into a surface during manufacturing
processes and subsequent surface processes and treatments. Material quality and surface
quality are unlikely to have a controlling effect on fatigue performance unless stress
concentration factors are low and there is no fretting.
2.3 Comparative testing. True comparison of materials for fatigue performance must be
done using coupons representative of component condition (including any significant batch
effects), loading and environment. Where changes are made to the material processing or
manufacturing route, comparative testing must be done under representative conditions to
ensure that fatigue performance has not been adversely affected.
2.4 Surface coatings and surface treatments. Surface layers of different structure or
properties to that of the bulk substrate may significantly affect fatigue performance. For
example, plating, cladding or thickened oxide layers may develop cracks which can then
grow into the substrate. Chemical surface treatments can produce surface pitting leading to
reduced fatigue performance. Part 1, Section 4 Leaflet 18 gives guidance on the effects of
surface finishing and protective treatments on fatigue properties.
Page 1 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
2.5 Temper and tensile strength. At high applied stress levels, that is low-cycle fatigue,
the fatigue strength can be strongly influenced by material tensile strength. At low applied
stress levels under constant amplitude loading there is, in some materials, a clear fatigue
limit, which can be related directly to tensile strength and below which no significant fatigue
cracking develops. Under variable amplitude loading this fatigue limit is suppressed. Fatigue
limits determined under constant amplitude loading cannot be assumed to apply to the same
cycle applied during a variable amplitude sequence. Where the existence of a fatigue limit or
a shallow slope on the fatigue curve in the high cycle region are essential for aircraft life then
the design curve must be substantiated for each material by tests under realistic loading on
coupons representative of the structural details concerned. (see also Para 2.3)
2.6 Operating stress levels. The selection of materials for high tensile strength to
minimise structural mass is seldom consistent with good fatigue performance. Lower strength
materials are often more suitable for structures in which tensile applied loading
predominates, for example lower wing skins, where it is often fatigue life rather than static
strength considerations that define allowable stress levels. In such cases the selection of a
high tensile strength material may offer little or no practical fatigue advantage; a lower yield
strength allows benefit to be obtained from residual compression stress fields produced by
tensile yielding at critical features, but can be detrimental in those cases where the
predominant loading is in compression; particular care must be taken in relation to fatigue-
enhancing systems (see Leaflet 35, Para 7.9)
3 FRACTURE TOUGHNESS
3.2 Section thickness. In thick sections the plane strain fracture toughness, K Ic , is most
commonly used to describe resistance to fracture; this is a material property independent of
geometry. In thin sections the toughness is influenced by component dimensions and cracks
may extend in a stable manner under high loads (particularly in low strength, high toughness
materials). In these cases a single value of toughness cannot adequately describe the
fracture resistance and a crack growth resistance curve (R curve) is required to define
failure. Material selection for sheet materials should be made on the basis of single-pull tests
on identical panels or using R curves.
3.3 Temper and tensile strength. For a given material type the toughness decreases
with increase in tensile strength. Peak age tempers (such as the T6 and T651 tempers in
aluminium alloys) usually have the lowest fracture toughness for a material and should be
avoided where possible. Relatively small reductions in tensile properties, for example by
over-ageing, can produce a significant increase in fracture toughness (and usually an
increase in resistance to environmental attack).
3.4 Product form and grain orientation. Toughness varies with product form and grain
orientation (i.e. with loading and cracking direction with respect to texture imparted by
material processing). In general, the fracture toughness is lowest if cracks are running
parallel to the short transverse grain, that is in ST or SL orientations, and highest if cracks
have to cross the grain, that is in LT or LS orientations.
Page 2 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
3.5 Toughness to yield ratio. To maintain a required level of damage tolerance, for
example a minimum acceptable critical crack size as employed in some aircraft design
philosophies, the ratio of fracture toughness to tensile strength must be maintained at an
acceptable level whereas the material property trend is for toughness to decrease with
strength in a given material. A measure of the damage tolerance capability of metallic
materials is the ratio of fracture toughness to yield strength (0.2% proof strength). This
parameter is termed the Toughness to Yield Ratio, TYR [1,2]. In fracture-critical components
it is beneficial to use a material with a high toughness to yield ratio. Experience on a range of
military aircraft has been used to define a rating, A - D, based on the TYR, which can be
used for any aerospace structural alloy for material selection, see Table 1 and Figure 1.
Materials with 'A' or 'B' rating should be selected where possible. Materials rated 'C' or 'D'
must not be used without the prior agreement of the Project Authority. Materials with a 'D'
rating shall be used only in exceptional circumstances.
3.6 Operating environment. For many materials the fracture toughness is largely
unaffected by environmental conditions. However, some material/environment combinations
exhibit rapid crack growth under stored or residual stress conditions due to stress corrosion
mechanisms, see Part 1, Section 4, Clause 4.3 and Section 4 Leaflet 7. For some materials,
particularly high strength steels, there exists a transition temperature below which toughness
can be dramatically reduced. It is important that materials do not exhibit such a transition
within the working temperature range of the component for which they are used.
4 IMPACT RESISTANCE
4.1 Definition. Impact resistance is defined as the energy required to cause cracking in
a given section thickness. Impact resistance cannot be determined solely by monotonic
toughness testing and ballistic tests. Where impact threats have been identified, confirmatory
tests of material performance should be done at appropriate impact energies with suitably
representative impactors and impact angles on coupons with representative restraint.
4.2 Problems arising from impact. Cracking occurring during impact can significantly
reduce residual strength and may subsequently extend due to fatigue loads. Materials should
not be used in product forms or section thicknesses which suffer significant cracking or
damage under the impact threats anticipated in component manufacture or peacetime
service. Where forward facing surfaces, particularly those ahead of intakes, are made of
metal, it is preferable to use ductile, high toughness materials. This increases the likelihood
of denting and ductile tearing rather than fragmentation. Materials with low impact resistance
are likely to experience cracking during manufacture.
4.3 Material properties. Materials with a low ductility and low toughness are likely to
suffer extensive damage from low speed - low energy impact. Some materials may exhibit a
variation in material properties through the section thickness. In such cases it is possible that
a low ductility region may be exposed by chemical or mechanical removal of material -
leading to a serious reduction in impact resistance [1].
Page 3 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
5.1 Definition. Crack growth rate for a given material, stress ratio (R ratio) and
environment is closely linked to applied stress and crack length through the range of stress
intensity factor, K. It is usual for crack growth life calculations to be based on constant
amplitude data which are generally available in terms of growth rate, da/dN, against the
applied stress intensity factor range, K.
5.2 Temper, tensile properties and modulus. Growth rate is strongly dependent on
modulus - the higher the modulus the lower is the growth rate. Materials with similar
modulus, for example aluminium alloys, exhibit very similar growth rates except in the near-
failure region; see Para 5.4
5.3 Low crack growth rates and the crack growth threshold. Unless the use of a crack
growth threshold can be substantiated under realistic service conditions (loading, crack size
and configuration, environment, etc.) then the crack growth curve should be estimated by
extending the steady state growth region on a straight line basis to very low growth rates,
see also Para 2.5
5.4 Product form and grain orientation. Crack growth rates in aerospace structural alloys
are usually independent of product form and grain orientation except in the near-failure crack
growth region. The fracture toughness, on the other hand, is highly dependent on grain
orientation, product form and temper.
5.5 Operating environment. In some materials the growth rate shows a marked
sensitivity to the operating environment. The effect of operating environment on growth rate
must be considered in damage growth analysis; guidance is given in [1]. Where an adverse
operating environment is known, then confirmatory crack growth tests should be done in that
environment. The peak-aged tempers in high strength 7000 series (aluminium - zinc) alloys
are particularly prone to increases in growth rate with humidity.
REFERENCES
Page 4 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
Page 5 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
TABLE 1
Values in the table are TYR values (m). LT values illustrated in Figure 1. Values of TYR are
determined from consideration of scatter in strength and toughness [2].
Page 6 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 35
FATIGUE
SAFE-LIFE SUBSTANTIATION
1 INTRODUCTION
1.1 For combat aeroplanes, the cornerstone of safe-life substantiation is a
full-scale test of a complete airframe to a factor of 5 times the service life. This
factor consists of an allowance of 3 1/3 for scatter in fatigue performance and a
further factor of 1.5 which is applied to those parts of the structure where the
loads are not monitored on a continuous basis. Where only one side of a
component has been tested, for example a port or starboard taileron, the
scatter factor is increased to 4. Exceptionally, individual structural elements can
be substantiated by calculation alone; here it must be shown that the
calculations are based upon conservative assumptions and that the calculated
safe life exceeds the service life by a factor of at least 2.0
1.2 For helicopters, the emphasis is on the rotating (dynamic)
components, where the stress levels are relatively low and compliance is
demonstrated using a factor on stress. Several examples of each component
are tested and in these circumstances a stress factor of about 1.5 is required to
allow for scatter. An additional stress factor of 1.2 is applied unless the design
loads are suitably conservative. A further stress factor may be used to
accelerate the test. Exceptionally, individual dynamic components can be
substantiated by calculation alone; here it must be shown that the calculations
are based upon conservative assumptions and that the calculated safe fatigue
strength exceeds that which would be required by test by a factor of at least 1.2
1.3 The scatter factor of 3 1/3 is insufficient to demonstrate that an
adequate safe life has been achieved for all fatigue-sensitive features and so
the test must be underpinned by evidence that suitable allowable stresses have
been achieved and that the test loading is of suitable severity.
1.4 Acceptable procedures for demonstrating compliance are summarised
below and described in the paragraphs which follow:
(a) construction and use of Safe S-N curves;
(b) derivation of usage spectra for design and substantiation;
(c) adaptation of usage spectra for test purposes;
(d) supplementary evidence for features with higher scatter than can be
accommodated in airframe tests;
(e) allowances for additional uncertainties when loads are unmonitored on
individual aircraft;
(f) allowances for additional uncertainties when components have not
been tested;
(g) checking for sensitivity to increases in loading severity;
(h) adjustment of lives under the Design Spectrum to equivalent lives
under the Service Spectrum.
1 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
1.5 Considerations underlying the test factors that are normally required to
allow for scatter in the fatigue performance of full-scale test specimens are
discussed in Paragraph 10.
1.6 The factors associated with the different aspects of safe-life
compliance are presented in Tables 1 to 4. It is emphasised that these factors
will be judged to be sufficient only in those circumstances where the Service
Spectrum is confirmed by regular sampling of the service loads by Operational
Loads Measurement (OLM) or Operational Data Recording (ODR) in
compliance with the requirements of Clause 3.2
1.7 Special considerations relating to the safe-life substantiation of fibre-
composite components are given in Leaflet 40. See also Part 1 section 3 leaflet
44 “Impact Damage Resistance of Composite Materials.”
2 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
2.5 The safe curve must take much the same shape as the mean curve
and retain the life factor as long as possible, consistent with a smooth transition
to the stress factor by an endurance of about 5 106 cycles. More precise
guidelines are unnecessary because no such constraints apply in choosing the
shape of the mean curve. However, the shape of the mean curve must suit the
feature under consideration. It must also be compatible with the material static
strength.
2.6 Figure 1 shows examples of Safe and mean S-N curves for an
aluminium alloy pinned-lug specimen. The curves were drawn using the basic
factors of Table 2 and a conventional four-parameter Weibull equation of the
form:
S = S inf (1 + A / (N + G)m)
where S = stress,
S inf = stress at infinite life,
A = numerical constant,
N = endurance,
G = numerical constant governing low endurances,
m = numerical exponent.
m A S inf G
Mean 0.604701 7072.70 7.25900 2820.00
Life-factored 0.604701 3794.80 7.25900 1007.00
Stress-factored 0.604701 7072.70 4.87181 2820.00
Safe 0.604701 5831.13 4.87181 1156.50
2.7 The factors of Table 2 are suitable for use in the design of components
manufactured from conventional aluminium alloys, titanium alloys and steels.
The factors to be used for structural features in advanced metallic materials
should be agreed with the PTL. Factors for the construction of Safe S-N curves
for fibre-composite materials are given in Leaflet 40. Factors of 10 on life and 2
on stress may be used for the construction of Safe S-N curves for castings, in
the absence of rationally-derived scatter factors in the development phase. In
all other cases, scatter factors shall be rationally derived, and the factors to be
used shall be agreed with the PT.
2.8 The use of Safe S-N curves in constructing test spectra is described in
Paragraph 4.
3 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
4 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
4.3.3 The Safe S-N curve should first be positioned so that Miner’s
rule gives the Specified Life under the Design Spectrum or, where
appropriate, the Service Life that is to be demonstrated under the Service
Spectrum.
4.3.4 When the test loading used to represent the Design (or Service)
Spectrum is of appropriate severity, the life calculated using the Safe and
mean S-N curves will differ by a factor of between 3.0 and 4.0.
4.3.5 If the calculated life factor is too high, the severity of the
spectrum can be increased by using the Safe S-N curve to express some
of the lower amplitudes as a smaller number of higher amplitudes giving
the same damage and then repeating the procedure.
5 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
5.3 Where the calculated life factor is more than 4.0, or more than 6.0 for
monitored features which are to be tested to a life factor of 5.0, supplementary
evidence should be provided in the form of a calculated allowable stress for the
required safe life under the test spectrum. The allowable stress should be
associated with Limit Load or exceptionally with the maximum load in the
spectrum if this is lower. Structural analysis, and/or strain gauge measurement
(on the test specimen), should be used to show that these allowables have not
been exceeded. Retrospective analysis introducing, the ‘no-test’ factors of
Table 4 will be required in those cases where the calculations are not
supported by relevant test data on structural elements.
6 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
7 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
8 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
9 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
10.3.5 Where more than one result has been obtained, the test life of
each feature is adjusted to an effective value where the residual strength
requirement is just satisfied. The appropriate test factor is then applied to
the log (geometric) mean of the failures.
10.3.6 The scatter factors for use in conjunction with the further factor
of 1.5 for unmonitored structure are summarised in Table 1(a).
10.3.7 It is emphasised that these test factors are insufficient to
demonstrate that an adequate safe life has been achieved for all fatigue-
sensitive features and so the test must be underpinned by evidence that
suitable allowable stresses have been achieved and that the test loading is
of suitable severity.
10 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
REFERENCES
450
Mean stress (net) = 55 MPa
400
Life-factored
Alternating stress (net), MPa
300
250
200 Mean
150
100
50
0
0 1 2 3 4 5 6 7
10 10 10 10 10 10 10 10
Life
FIG.1a - SAFE AND MEAN S-N CUVES FOR AN ALUMINIUM ALLOY PINNED-LUG
SPECIMEN - SHOWING COMPATIBILITY WITH SAFE VALUE OF STATIC STRENGTH
11 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
110
Mean stress (net) = 55 MPa
100
90 Life-factored
Alternating stress (net), MPa
Stress-factored
80
70
60 Mean
50
40
30 Safe
20
10
0
4 5 6 7
10 10 10 10
Life
FIG.1b - SAFE AND MEAN S-N CURVES FOR AN ALUMINIUM ALLOY PINNED-LUG
SPECIMEN - SHOWING TRANSITION BETWEEN LIFE AND STRESS FACTORS IN THE
REGION OF MOST INTEREST
12 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
These life factors have been derived using the customary statistical procedure
(Bullen’s method 2(iii) [3]), together with a standard deviation in Log 10 life of 0.1296;
the figures of 3.75 and 4.00 have been slightly rounded. The factors correspond to a
probability of failure of 1 in 1000 on the basis of 1 or 2 items per aircraft. This same
procedure may be used to calculate the corresponding factors for other numbers of
tests.
(b) Factors on stress - to be used in lieu of factors on life in circumstances where
most fatigue damage is done by large numbers of low-amplitude cycles.
These factors must be used in calculating safe lives when loads are unmonitored on
individual aircraft, and the loading spectrum is a best estimate confirmed by regular
sampling of service loads (see Paragraph 6).
Factor on life at low service lives Factor on stress at high service lives
1.5 1.2
These factors are applicable only when the life has been estimated using S-N data in
accordance with the recommendations of Paragraph 7. The additional factors to be
used in conjunction with other methods of life estimation must be supported by
relevant test evidence and must be acceptable to the PTL.
Factor on life at low service lives Factor on stress at high service lives
2.0 1.2
13 of 13
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 36
FATIGUE
INSPECTION-BASED SUBSTANTIATION
1 INTRODUCTION
1.1 This leaflet describes acceptable procedures for demonstrating compliance with the
requirements of Clause 3.2 It covers:
(a) conditions governing the use of inspection-based substantiation;
(b) detection of cracks;
(c) time to first inspection;
(d) conditions governing the determination of crack-growth curves;
(e) allowances for uncertainties in estimates of inspectable life;
(f) derivation of inspection intervals;
(g) check for the sensitivity of inspection intervals to increases in loading severity.
It is noteworthy that consideration must also be given to the monotonic and cyclic
variations in service temperature where these may have a significant deleterious effect on
performance.
Terms shown in bold type when they first occur are defined in Paragraph 9.
1.2 This leaflet deals primarily with the use of directed fatigue-related inspections to
support structural integrity in service. It is generally assumed that inspections will be defined
using a deterministic approach using fixed mean or design data with factors applied to
calculated life to reflect the known variability in input parameters. This approach centres on
the use of a factor of 3 to allow for scatter in crack propagation as well as the uncertainty in
estimating the crack size that is very unlikely to be missed under service conditions. As such,
it makes provision for inspection intervals to be increased by improvements in inspection
technique, but leaves no room for the wider benefits that could be obtained if sufficient data
could be provided to justify a probabilistic approach [1]; the application of such an approach
would require, in addition, the probability distributions of load, crack length at a given time
and detection capability (probability of detection).
1.4 The leaflet is aimed primarily at metallic structure; special considerations relating to
fibre-composite components are discussed in Leaflet 40.
2.1 As a general rule, inspection-based substantiation should not be used for compact
and inaccessible structure or for any structure in which fatigue damage is difficult to detect. It
must be used, however, for any component which is susceptible to defects or to damage in
manufacture or service. It may also provide a means for extending the life of selected safe-
life components for which the inspection penalties are acceptable.
Page 1 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
2.5 Normally the critical crack size should be at least 2.0 mm.
2.6 Multiple-site damage (MSD) can present a serious threat to the integrity of
structures and mechanical components that are the subject of inspection-based
substantiation. MSD can occur where nominally-identical features are uniformly loaded so
that simultaneous cracking occurs from adjacent sites, such as rivet holes. Small cracks can
suddenly link up to form a crack of critical size; the small size of cracks at the point at which
they link up is often at the limit of practical detectability. The problem is largely associated
with pressure-cabin lap joints on large transport aircraft. Normally the safety of components
can be assured without fatigue-related inspections by withdrawing them from service upon
completion of a safe life; this would need to be established by test in the case of components
susceptible to MSD. However, experience suggests that it may be impractical to protect
pressure-cabin lap-joints from a degree of local degradation due to corrosion and/or
disbonding/loss-of-friction - thereby invalidating the safe life. In such circumstances it must
be shown that any MSD arising from local degradation will be safely contained until it is
detected by the agreed programme of RCM inspections.
3 DETECTION OF CRACKS
3.1 The cornerstone of inspection-based compliance is the crack size that will be
detected reliably by inspection - the detectable crack size. If too large a crack size is chosen,
inspection intervals will be unnecessarily short. On the other hand, if the crack size is too
small the inspection interval could be much too long (the relationship is exponential) and
safety could be undermined.
3.2 As a general rule, the aim should be to choose a detectable crack size that is very
unlikely to be missed at the given location under service conditions. This choice must be
guided by experienced NDI operators using accumulated evidence for the technique in
question and taking account of the standards that have been achieved when special trials
have been done. Detection capability must be demonstrated on test articles representing the
component geometry, cracking, surface condition and structural configuration of the
component in question, with consideration given to lighting and access restrictions which
may apply in service.
3.3 Inspections must be carried out by trained operators using techniques that have
been demonstrated on representative test articles with representative defects. Consideration
must be given to the need for regular calibration and replacement of NDI probes.
Page 2 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
3.4 Attention is drawn to the need to avoid exposing operators to the risk of missing
cracks because they are accustomed to finding nothing.
3.6 Loads acting to force crack faces together or surface treatments such as machining
or shot peening can mask surface-breaking cracks. Where practicable the aircraft should be
loaded such that anticipated cracks will be open during an inspection.
3.7 Evidence of cracking from service inspections and data available from test and tear-
downs should be collated to increase confidence in the specified inspection regime.
3.9 Where limited disassembly is required for crack detection, for example fastener
removal, then the structure must afterwards be restored to the pre inspection condition,
including any fatigue enhancements previously incorporated. Disassembly and re-assembly
must not in themselves introduce damage atypical of the original build quality.
3.10 The method of inspection must reflect the position, orientation and mode of cracking
most likely to occur, taking into consideration processes such as cold working that can
transfer the crack initiation location away from that expected in untreated structures.
4.1 The time to the first inspection must be the safe life (Leaflet 35) except where an
item is made from a material which is susceptible to defects or to damage in manufacture or
service, including impact from runway debris, environmental degradation or may still contain
cracking or damage following a repair. In such cases the time to first inspection shall be
equal to the normal inspection interval.
4.2 Following fatigue arisings in service, the time to first inspection (safe life) for
unrepaired aircraft in the fleet must be assessed by recognised statistical techniques, such
as a maximum likelihood approach [2]. It cannot be assumed that a particular arising is
typical of fleet status.
4.3 Normally, the safe life of inspection-dependent structure shall be at least half the
Specified Life under the Design Spectrum. Exceptionally, a lower safe life may be acceptable
for easily-inspectable structure. Conversely, a higher safe life may be needed for buried
structure in order to meet the requirement for the inspection penalty to be acceptable on
operational and economic grounds.
Page 3 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
5.1 By definition, inspection-based substantiation must be based only upon that part of
the crack-growth curve that is detectable in service with NDI equipment that is acceptable to
the Project Authority.
5.2 A calculated crack growth curve for the detail in question is central to the definition
of inspection intervals. A best estimate of the mean crack growth curve must be produced as
described below. Appropriate allowance must be made for the uncertainties examined in
Paragraph 6.
5.3 The following are examples of the considerations which may influence the
maximum acceptable crack size:
the onset of rapid, unstable crack growth.
net section failure (tension or compression) including any allowance for loss of
area due to corrosion.
unacceptable leakage or loss of pressure.
maximum crack size that can be removed - this may be governed, for example,
by maximum bush or oversize-fastener dimensions.
maximum size allowing practical or economic repair.
any other failure mode, such as buckling, induced by crack growth.
5.4 Mean growth rate data, determined from constant amplitude tests, are normally
used in growth rate calculations. These data must include environmental and temperature
effects where appropriate. If the crack growth characteristics of the material are not known
then test data sufficient to define the growth behaviour, including environmental effects,
variability in growth rates within and between batches, etc., must be obtained.
5.5 It is usual to calculate crack growth using linear summation methods, making no
allowance for retardation of crack growth occurring as a result of peak loads in the spectrum.
However, as non-linear prediction routines are generally available and can afford a
significant increase in predicted life over linear summation predictions, which are inherently
conservative for tension dominated loading spectra, then consideration must be given to their
use. Any benefits over and above linear summation predictions must be shown to be
appropriate for the in-service loading, confirmed by representative tests and agreed by the
Project Authority.
5.6 The crack growth threshold observed in coupon tests with long, through-section
cracks under constant amplitude loading in laboratory air may not be observed under
variable amplitude loading in realistic service environments; much shorter cracks may
propagate under these conditions. It is usually conservative to assume a zero threshold and
extrapolate the steady-state growth rate behaviour down to very low stress intensities. Any
inspection period relying on the existence of a crack growth threshold must be substantiated
by appropriate representative testing.
Page 4 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
5.7 The variation of stress intensity factor coefficient during crack development may be
estimated from standard texts, from finite element analysis or from observations of crack
behaviour made in tests and tear-downs. A record of solutions used must be included in the
Fatigue Type Record.
5.8 Cracks should be considered to exist at the most unfavourable location in a load
path. Where adjacent load paths or details are equally loaded then identical cracks should be
considered in each.
5.9 Where cracking reaches a boundary (an edge, cut-out or adjacent hole) a
secondary crack must be assumed to exist at the most unfavourable position in relation to
this boundary to allow consideration of continuing damage.
5.10 Inspection-based approaches must take no account of the life following arrest of
rapid, unstable crack growth unless arrest capability has been demonstrated in
representative tests and the damage will be evident in service.
6.1 The following uncertainties must be taken into account in estimating the mean crack
growth curve which determines the inspectable life. The calculations should be supported,
where practicable, by test evidence from element tests or the structure itself.
6.2 The geometrical correction factor used in the stress intensity factor solution, the
assumed loading spectrum and the data used in the crack growth analysis should err on the
side of conservatism.
6.4 Uncertainties in fracture toughness: The inspectable life is terminated when the
crack reaches the maximum acceptable size. Where the maximum size is governed by
fracture toughness, the critical crack length should be based upon Minimum Specification or
‘B’ allowable values of fracture toughness. Where only mean toughness data are available
these should be reduced by a factor of 1.3
Page 5 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
7.1 The inspection interval is obtained by dividing the inspectable life by a factor of 3.0
to allow for scatter in crack propagation and uncertainty in estimating the detectable crack
size. The derivation of inspection intervals is shown schematically in Figure 1. It is
noteworthy that the factor of 3.0 is in addition to the factor of 2.0 that is applied to estimates
of inspectable life which are unsupported by relevant test evidence (Para 6.3).
7.2 The inspectable life must be divided by a further factor of 1.5 when loads are
unmonitored (Leaflet 35). Higher factors may be required for helicopter dynamic components
and other items which experience large numbers of damaging low-amplitude stresses.
7.3 The smaller the detectable crack size the longer will be the inspectable life and,
potentially, the inspection interval. However, to find smaller cracks a more sensitive
inspection technique will be required which may incur economic and downtime penalties that
can offset any benefits from an increased inspection interval.
7.5 Where inspection intervals have been specified in terms of flying hours under the
Design Spectrum, they must be expressed as equivalent flying hours under the Service
Spectrum. This complication is avoided in those cases where it is appropriate to specify
intervals in terms of Fatigue Index or some other appropriate parameter.
7.6 If there are any significant changes in usage or in the aircraft structure itself then the
inspection periods of those items which are affected must be re-assessed.
7.7 It may sometimes be practicable to permit continued operation after a crack has
been found. In these circumstances the desired maximum crack size to be left in the
structure after inspection would be balanced against the desired inspection interval.
Specifically, the maximum crack size to be left in the structure would be treated as if it were
the ‘detectable crack size’ and the inspection interval would be obtained in the usual way.
8.1 As noted in Leaflet 35, in the longer term, the loading spectrum experienced in
service seldom corresponds to the Design Spectrum. For wing-bending, for example,
increases in aircraft mass and in the severity of the normal acceleration spectrum can
increase the severity of the loading spectrum by a factor of 1.2 or more.
8.2 At the design stage, those features with exceptionally high sensitivity to increases in
loading severity must be identified and their stresses reduced so that they do not impose
restrictions on inspection intervals for the aircraft as a whole.
Page 6 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
8.3 The inspection intervals of individual structural features must be estimated with the
stresses under the Design Spectrum elevated by a suitable factor. This factor, which will
typically be 1.2, must be agreed with the Project Authority. The inspection interval under the
enhanced loading will normally be estimated by a calculated adjustment of the interval
established under the Design Spectrum. In making the calculation, the extreme stresses
under the elevated loading should be clipped/truncated to the extreme stresses under the
Design Spectrum (normally the Limit Loads).
8.4 The inspection interval required under the elevated Design Spectrum must be
acceptable to the Project Authority and reported in the Fatigue Type Record.
9 DEFINITIONS
9.1 Detectable crack size: The size of crack that is very unlikely to be missed by a
given technique for a particular combination of material, component, access and
environment.
9.2 Inspectable life: The average period of usage throughout which a crack is of
detectable size. The period begins when the crack reaches the detectable size and
terminates when the crack reaches the maximum acceptable size.
9.3 Time to first inspection: The period from entry into service to the point at which the
first inspection must be made.
9.4 Inspection interval: The maximum interval between inspections under the Design
Spectrum.
9.5 Equivalent inspection interval: The maximum interval between inspections under
the Service Spectrum.
9.6 Critical crack size: The crack size associated with the onset of rapid, unstable
crack growth under the load corresponding to the residual strength requirement.
9.7 Maximum acceptable crack size: The Critical crack size or a lesser size
associated with other constraints such as those listed in Para 5.3
REFERENCES
Page 7 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
Page 8 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 37
FATIGUE
TESTING
1 INTRODUCTION
1.1 This leaflet describes the scope of testing which must be done to satisfy the safe-life
substantiation requirements of Clause 3.2.12 It covers:
(a) development tests;
(b) airframe fatigue tests;
(c) representation of high-frequency loading;
(d) test accuracy;
(e) documentation of tests;
(f) residual strength tests and tear-down inspections.
1.2 For components which experience significant fatigue loads at high cycles, such as
rotating components on helicopters, fatigue testing should be designed to substantiate both
the low and high frequency load cycles and to establish accurately the endurance limit.
2 DEVELOPMENT TESTS
2.1 Where appropriate, design development tests must be done to support the selection
of materials, to establish basic data for design and to support the methods of analysis to be
used in substantiation.
2.2 Generic data on fatigue performance and crack growth may need to be
supplemented by tests on specimens representing particular features of the structure. In
general, these specimens should be tested under constant amplitude loading and under
spectra which will reveal any interaction effects that may need to be taken into account.
Where it is necessary to use a specimen representing a major sub-assembly, it will usually
be sufficient to test this under realistic loading.
Page 1 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
3.1.4 At suitable intervals the structure should be inspected for fatigue cracks.
In these inspections, full advantage should be taken of advanced inspection
equipment such as articulated video probes; particular attention should be given
to those locations which are known to be fatigue-critical. Early warning of
damage is essential in order to devise repair solutions, enable alternative repair
schemes to be evaluated and, where possible, to confine the damage.
3.1.6 The aim must be to continue the test to the required duration or until the
structure is no longer representative through reasons of repair, production
changes or a major failure.
3.1.7 The fatigue lives and crack-growth periods observed on test must be
reduced to allow for scatter and adjusted to equivalent lives under the service
spectrum using the procedures described in Leaflets 35 and 36.
Page 2 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
3.2.3 The test must proceed at a brisk pace to enable any necessary design
changes to be incorporated into production airframes with a minimum of
retrospective modifications.
3.2.4 Within the constraints of early manufacture and testing, the quality of
materials and manufacture, including influences such as surface treatment,
should be as representative as possible.
3.2.7 When all testing is complete, it is essential that those main load paths
that are still representative of the production structure should be subjected to a
tear-down inspection in order to reveal any significant fatigue damage that has
not been observed during the test.
3.2.9 The Production Major Airframe Fatigue Test (PMAFT) must be done on
a structure, or equivalent representative components, manufactured to the
mature production standard. The PMAFT should be timed to extend the
clearance obtained from the DMAFT and maintain a clearance that, for the bulk
of the life, is at least 10% ahead of the life consumed by the fleet leaders.
3.2.10 The loading used should simulate typical usage. Where the usage
differs appreciably between groups of aircraft and these are not rotated for fleet
management purposes, care should be taken to avoid using extreme loads which
are higher than those which will be experienced by most aircraft in the fleet
(Leaflet 35).
Page 3 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
3.2.11 As the PMAFT proceeds, all structural modifications which have been, or
are to be, embodied on service aircraft should be incorporated on the specimen.
The test life at which each modification is done should exceed the life chosen for
service-embodiment by the appropriate test factor (Leaflet 35). Exceptionally,
earlier modification may be necessary to prevent the cracking in the test
specimen from becoming unrepresentative.
3.2.12 When all testing is complete, the specimen must be subjected to a tear-
down inspection in which the main load paths are carefully dismantled and
inspected in order to reveal any significant fatigue damage that has not been
observed during the test (Paragraph 8). The timing of residual strength testing
and the tear-down inspection will be dependent on the life of the type and should
be balanced against the need to retain the test article for further testing, as would
be necessary if a life extension programme were required.
3.4.2 The Life Extension Fatigue Test (LEFT) must be done on a structure, or
an equivalent assembly of components, which is representative of the fleet,
including any fleet-wide modifications. Supplementary testing may be necessary
to substantiate the life extension of build concessions or other modifications, or
where existing data are inadequate.
3.4.3 The LEFT must proceed at a sufficient rate to ensure that the test retains
an adequate lead over the fleet. This lead should be sufficient to allow for the
design and implementation of repairs to the LEFT article.
Page 4 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
4.1 Components which suffer significant fatigue damage at high cycles may need to be
tested separately using a factor on stress.
4.3 In tests on airframes, it is important to apply the high frequency loading so that the
structure is excited in a realistic way. This means that loading frequency becomes an
important parameter. Successful tests have been done on helicopter airframes using bursts
of severe, but otherwise unfactored, loading interspersed between the low frequency loads.
Aeroplane structures tend to have been tested by applying lesser numbers of higher-
amplitude loads that produce equivalent damage (whilst retaining sufficient lower amplitude
cycles to ensure that any effects of fretting are represented), but here too there are moves to
apply more realistic loading.
5.1 Close control, recording and monitoring of loading is necessary in fatigue tests
because small changes in loads can cause large changes in fatigue life. In tests on complete
structures or systems, or on major components, the accuracy of the applied loads should be
within 3% of the maximum demanded load.
5.2 All applied loads should be measured periodically during the test to check that the
repeatability is within 2% of the particular applied load or 0.5% of the maximum demanded
load, whichever is the greater. This should be done by checking the applied jack loads rather
than the measured loads in the test article.
5.3 Test systems which provide perfect control of applied loads at discrete monitoring
points may produce excessive strains at other locations due to dynamic effects. The aim
should be to achieve the required accuracy and repeatability throughout the structure.
6.1 Fatigue test summaries must be provided in order to meet the requirements of the
Fatigue Type Record (see MAP RA 5309).
Page 5 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
7.1 Where practicable, the test specimen should be used to provide evidence to support
the Critical Crack Lengths which have been used in the interpretation of the test result. This
might be done by removing selected repairs in turn and observing the behaviour of the
cracks under loads of up to the Design Limit Load the residual strength factor. Areas where
repairs have been effected by means of blending or use of oversize holes may also be tested
in this way. At this early stage in the residual strength testing the aim should be to obtain
useful measurements at a number of locations and so the loads should be limited to avoid
premature failure.
7.2 Subsequently the tests should probe the specimen for any debilitating cracks not
detected during the fatigue test. On-line inspection or monitoring techniques should be used
where possible to minimise the chance of premature failure.
8 TEAR-DOWN INSPECTIONS
8.1 The purpose of a tear-down inspection is to identify any significant cracking that has
not been revealed by inspections during the fatigue test and any residual strength tests.
8.2 The inspection also provides a valuable pointer to those further damage sites that
will need to be considered if the life is extended on the basis of inspections.
8.3 The main load-carrying structure must first be completely dismantled. Extreme care
is needed in order to avoid masking or destroying fatigue damage.
8.4 The structure should then be examined using appropriate non-destructive inspection
techniques, paying particular attention to fastener holes, cut-outs and changes of section.
Structure loaded primarily in compression should be included.
8.5 Each damage site should be examined fractographically and, where appropriate,
crack growth characteristics should be determined.
8.6 Where significant damage is discovered in load paths which may not have been
subjected to a residual strength test, calculations must be done to determine the size of the
Critical Crack. Where this is less than the crack size at the end of the fatigue test, the test life
must be reduced. Supplementary testing may be necessary to support this adjustment.
8.7 Data obtained during the tear-down inspection may be used to validate or develop
inspection techniques.
Page 6 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 38
FATIGUE
SERVICE MONITORING
1 INTRODUCTION
1 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
2 TERMINOLOGY
2 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
3 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
4 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
4.6 For helicopters lacking an adequate HUMS fit, automatic flight data
logging (manoeuvre recognition) exercises or Manual Data Recording
Exercises (MDREs) are conducted to provide data to update the SOIUs.
5 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
5.6 Design proving information and fleet statistics are used in conjunction
with ODR/OLM results to substantiate the fleet-wide monitoring process. The
following should be considered:
(a) Flight Loads Measurement data acquired by the Designer/Design
Organisation during development of the type, in conjunction with structural
modelling, are used to confirm that the ODR/OLM monitor points cover the
remaining fatigue critical structure.
(b) Results from the continuous monitor, including usage statistics, are
required to align the ODR/OLM sample to match overall fleet usage. Where
the continuous monitor does not provide adequate usage statistics, periodic
MDRE/data logging should be used.
(c) Fatigue test spectra are compared with the normalised ODR/OLM
results, to confirm that test loading is representative. Test results also help
to confirm fatigue critical feature selection and establish the life associated
with the monitored locations. For aircraft managed under damage tolerance
principles, test results are used in the establishment of inspection thresholds
and intervals.
(d) The performance of the fleet-wide monitor is compared with the
normalised sample of ODR/OLM data. The monitor process can thus be
validated or revised to improve correlation.
5.7 The requirement to carry out a periodic ODR/OLM is reviewed every 2
years but the interval between programmes should not normally exceed 5
years. The review should take into account any change in usage of the aircraft
and whether it is likely to degrade the continuous monitor. Maintaining a
permanent recording capability can be less onerous than trying to reinstate the
system whenever it is deemed further data are required.
5.8 The PT has the overall responsibility for running the ODR/OLM
programme, however, the specific requirements of the system; the equipment
interfaces and analysis routines should be agreed between the
Designer/Design Organisation, the Service specialists and their advisors. This
includes selection and positioning of sensors. This process can be formalised
as a Project Definition Study, which can also define the objectives and scope of
the programme. The aspects to be considered in the realisation of an OLM
programme are detailed in [1].
6 SYSTEM INSTALLATION AND ANALYSIS CONSIDERATIONS
6.1 The guidance in the following paragraphs applies equally to OLM/ODR
and fleet-wide monitoring installations.
6.2 The complexity of the instrumentation fit is normally related to the
number of different loading actions that have an influence on the life of the
structure and the potential variability of flight profiles.
6 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
7 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
6.11 All aspects of the System need to be assessed when meeting the
overall system accuracy requirement. This includes, but is not limited to, sensor
accuracy, bandwidth, sampling rate and digitisation resolution. The level of
accuracy required for a fleet-wide monitoring system is dependent on available
structural margins. Where margins are considered to be large, a less accurate
system than that specified in Clause 3.2.27 might be considered acceptable,
provided that overall damage predictions are not overly optimistic and any
conservatism does not compromise the design life, or cause an unacceptable
inspection burden.
6.12 Data loss can occur for a number of reasons (e.g. processor conflicts,
media jams, lost media, or even media not fitted). Gaps during recording
should be kept to a minimum, should be logged and should not be load
dependent. A significant source of lost data is usually data transfer between the
aircraft and the GSS. Rigorous procedures need to be in place to minimise the
risk of excessive unmonitored flying.
6.13 Both analysis of operational flight data and limit load exceedance
identification should utilise validated time histories, i.e. these should not contain
interference spikes or voids. Aircraft, that require fast turn-around between
sorties, will have real time limit exceedance monitoring with post flight
indication. This is to prevent continued operation when there is a risk of
permanent structural deformation. Ideally all aircraft should have an on-board
indication of a loads limit exceedance but in some cases, where time history
data are transmitted to the GSS, it may be possible to utilise the GSS to report
limit exceedances. Ground analysis of OLM/ODR data is considered essential
to establish the quality of the data and that the expected relationships between
the recorded parameters are realised.
7 REFERENCES
1. Reed, S.C. and Holford, D.M. Guidance for Aircraft Operational Loads
Measurement Programmes. MASAAG Paper 109, 31 May 2007.
8 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 39
FATIGUE
LIFE EXTENSION
1 INTRODUCTION
1.1 This leaflet provides guidance on the additional considerations which apply to
ensure that the customary standard of compliance is maintained when service lives are
extended.
1.2 The need for life extension should be kept under constant review in order to give
adequate time for preparations to be made.
1.3 The period of extended service life must be clearly defined in terms of loading
spectra and appropriate parameters including, flying hours, landings, Fatigue Index and
numbers of pressurisation cycles.
(b) those features where the safe life may be extended, subject to substantiation by
additional testing and/or analysis, or by in-service monitoring;
(c) those components which must be retired at their original design life;
(d) those components where the life may be extended beyond the safe-life by the
adoption of an inspection-based methodology;
2.3 The assessment should consider the methods and assumptions used in the original
substantiation, together with any design modifications and relevant operational experience
including any instances of cracking, repair, corrosion, wear or accidental damage.
2.5 All airframe and mechanical components considered in the life extension
assessment should be checked against the build standard to identify components which
have been superseded or are obsolescent.
Page 1 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
2.6 Reference must be made to all concessions which have been permitted during the
manufacture of the aircraft. If it cannot be shown that items with concessions have the
necessary additional life, then they must be identified and either modified or withdrawn from
service.
3.1 A supplement to the Fatigue Type Record must be prepared in accordance with the
requirements of MAP RA 5309.
3.2 This supplement must record the evidence provided to demonstrate compliance with
all the requirements of Clause 3.2 and its associated leaflets, for the period of the life
extension.
3.3 If no Fatigue Type Record is available, a life extension document must be prepared
to record the evidence provided to demonstrate compliance with all the requirements of
Clause 3.2 and its associated leaflets, for the period of the life extension.
Page 2 of 2
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 40
FATIGUE
FIBRE-COMPOSITE COMPONENTS
1 INTRODUCTION
1.1 This leaflet provides guidance on methods of substantiating the fatigue life of
polymer-matrix fibre-composite components 1 . In order to minimise the need for in-service
inspections compliance is normally demonstrated using a safe-life approach. However, the
most common failure mode in fatigue-sensitive composites is delamination growth and this
can sometimes be exploited to provide an acceptable inspection-based substantiation.
1.3 The performance is also influenced by the hygrothermal properties of the matrix -
the matrix tending to ‘soften’ after prolonged exposure to moisture and under elevated
temperature. Slight softening of the matrix can have a small beneficial effect on notched-
tensile strength, but can produce a marked reduction in compressive strength. Any such
effects must be taken into account in planning the fatigue test programme.
1.4 The fatigue strength of composite materials may be reduced by impact damage and
this must also be accounted for in the fatigue substantiation.
1.5 The leaflet draws largely upon experience gained in the Helicopter Industry, where
composites are used for the manufacture of dynamic components such as rotor blades and
control systems. In these applications the fatigue environment is more challenging than is
normally found in applications to airframes. Most helicopter dynamic components are of safe-
life design and are substantiated using a factor on stress (see Leaflet 35). This same
procedure has been applied to composite components and is reproduced here for more
general application.
1.6 In aeroplane airframe applications, the relatively high static notch-sensitivity and
modest impact resistance of high-modulus composites, together with the relatively low
gradient of their S-N curves, means that most components are unlikely to experience
significant fatigue problems [1]. The reserves of fatigue strength are normally sufficient to
permit substantiation using a conservative adaptation of the safe-life procedure described in
the following paragraphs, but using static testing to locate the mean S-N curve and drawing
upon library data for S-N curve shape and scatter in fatigue strength. Fatigue substantiation
obtained in this way must be underpinned by the extensive ‘building-block’ programme of
hygrothermal-mechanical tests on coupons, structural elements and components that
accompanies static certification. The next generation of advanced composites for airframe
applications is expected to be less notch-sensitive and to have improved resistance to impact
damage. As these restrictions are eased, then fatigue can be expected to become a more
significant consideration in the design of composite airframe structures [2].
1
Hereafter termed ‘composite’ components
Page 1 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
1.7 Composite materials may have higher coefficients of variation of fatigue strength
than metals. It is important that this variability is not underestimated when determining the
appropriate test factor on stress to be used in constructing a Safe S-N curve.
2.1 A structure or mechanical component can sometimes have more than one critical
feature according to influences such as loading complexity and severity, together with
environmental effects.
2.3 The structural elements used to obtain these S-N curves must be made from the
same materials and by the same processes as the features themselves. They should
generally be of the same size as the feature and be loaded so that the conditions at the test
section represent those in the component.
2.4 These structural elements are normally found to have shallow S-N curves which
may not exhibit a distinct endurance limit. The endurance limit, S inf , should be taken to be the
alternating stress a component can sustain for 109 cycles; for practical purposes S inf may be
associated with a lower endurance providing this does not increase the value by more than
about 1%.
3.1 Clause 3.12 specifies that components which are exposed to impact damage must
be the subject of an inspection-based substantiation. However, the structural performance of
composite components can be degraded by an insidious form of impact damage which is
neither apparent nor readily detectable by inspection. This is termed Barely-Visible Impact
Damage (BVID); it is believed to correspond to the term ‘Undetectable Damage’ used by the
US military authorities.
3.2 The nature of the BVID threat and how it is to be represented for compliance
purposes must be agreed with the Project Authority. The appropriate energy level will
normally be applied using a ‘blunt’ impactor.
3.3 In order to demonstrate tolerance to BVID arising from a defined threat, all full-scale
composite fatigue test specimens, other than the structural elements used to define S-N
curves, must be subjected to the appropriate impact conditions before they are fatigue
tested.
4.1 The coefficient of variation of fatigue strength, which is needed to determine the
appropriate test factor on fatigue stress, can be estimated from tests on representative
structural elements. Lower scatter is likely to be obtained if the tests are done using variable
amplitude loading.
Page 2 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
4.2 At least 20 specimens should be tested. For most purposes tests on four specimens
at each of five stress levels will be suitable. The lowest stress level should be chosen so that
most specimens fail during the test.
4.3 A mean ‘S-N’ curve should be determined from the test data using, for example, a
four-parameter Weibull equation (see Leaflet 35) and taking due account of any run-outs
(non-failures). Run-outs must not be treated as failures, as this can seriously distort the
shape of the curve.
4.4 When the test data do not include run-outs, an estimate of the coefficient of variation
can be determined as below. If run-outs are present, the estimate will normally be an integral
part of the optimisation procedure used in fitting the S-N curve (see, for example, [3]).
However, the procedure given below can also be used, providing the proportion of run-outs is
low and that the results from the unfailed specimens are discarded. Static results must not be
included in the analysis.
2
1 n S i S Mi
n p i1 S Mi
where S i is the stress (amplitude for constant amplitude tests) at test point i, and S Mi
the corresponding stress on the mean curve at the same life, n is the number of test
points corresponding to failures and p is the number of unknown parameters whose
values have been determined in the curve-fitting process (i.e. four in the case of a four-
parameter Weibull equation).
(a) The test specimens should be made from each of at least three batches of each
material.
(b) Alternatively, if the failure modes in static and fatigue tests are similar, the
coefficient of variation in fatigue strength obtained from a single batch can be adjusted
using variability data from static tests on coupon specimens, such as may be used for
batch acceptance purposes [4]:
and b = sb ft / st
where sb = coefficient of variation for static strength batch-to-batch variability
ft = coefficient of variation for fatigue strength variability from coupon tests
st = coefficient of variation for static strength variability from coupon tests
Page 3 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
5.1 The mean S-N curve for a component must be positioned using test results from full-
size specimens.
5.2 In the simplest case, where only a single failure mode need be considered, an
appropriate mean S-N curve (obtained from structural element tests and normally a library
curve) must first be fitted to each test result in turn so that a family of S-N curves is obtained
- each curve corresponding to a Miner damage summation of unity.
5.3 The mean S-N curve for the component must then be positioned to correspond to
the mean fatigue strength given by these curves. It is often convenient to obtain the mean
fatigue strength at the fatigue limit, but by the nature of their construction the relative fatigue
strength given by the curves does not vary with endurance.
5.4 It is usual to substantiate those potentially critical features which do not fail on test
by repeating the above procedure assuming they were about to fail. However, in principle,
these other features could be substantiated using supplementary evidence as described in
Leaflet 35.
5.5 The mean S-N curve for the component is taken to be that of the failure mode which
gives the lowest safe life (see following paragraph) under the service loading.
6.1 The factors, given in Table 1 are to be used to reduce the mean S-N curve to obtain
the Safe S-N curve.
6.2 As can be seen, the factors depend upon the population coefficient of variation and
the number of components tested. If high scatter in fatigue strength (coefficient of variation
15%) has been demonstrated, then appropriate factors must be agreed with the Project
Authority.
6.3 The factors have been calculated to correspond to a nominal probability of failure of
1 in 1000, assuming a normal distribution of fatigue strength. However, unlike the factors
used for metallic materials, they have been calculated to provide 95% confidence that this
probability will be achieved (i.e. that it will be achieved on approximately 19 occasions out of
20 rather than on average) and no account has been taken of the notional number of
[identical/identically loaded] components per aircraft [5]. As it happens, in this application, the
factors are closely similar to those that would have been obtained using the procedure
described in Leaflet 35 and so there is no reason for them to be changed.
7.1 As is the case for metallic materials, fatigue life scatter factors depend on the
coefficient of variation of fatigue life which, unlike the coefficient of variation of fatigue
strength, varies significantly with endurance. It follows that life factors cannot be established
directly without extensive testing and so a form of Safe S-N procedure is used.
Page 4 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
7.2 It is inappropriate to use a life factor of 2.8 (Leaflet 35) in constructing Safe S-N
curves for composite materials. This is because the factor is based on the scatter observed
at low endurances when structural elements made from conventional metallic materials are
tested under variable amplitude loading. Composite materials generally exhibit higher life
factors.
7.3 Appropriate life factors for test purposes can be determined by constructing a Safe
S-N curve from the mean [constant amplitude] S-N curve using an appropriate stress factor
(only) from Table 1. Where appropriate, a further stress factor of 1.2 may be required to allow
for uncertainties in loading (see Leaflet 34).
7.4 The life factor corresponding to any [variable amplitude] test severity can then be
obtained by calculating the life at that severity, using Miner’s rule and the mean S-N curve,
and dividing by the corresponding life obtained from the Safe S-N curve. It is emphasised
that factors obtained in this way will not necessarily be achievable on test for all potentially-
critical features. Guidance on their interpretation is provided in Leaflet 35.
7.5 Where a life factor is required for discrete (constant amplitude) stress amplitudes
(as might be interspersed between stress-factored loading in tests on some helicopter
dynamic components), this can be obtained directly from the S-N curves by dividing the
mean life by the safe life at the test amplitude concerned.
8.1 The static and fatigue strengths of polymer-matrix composites are affected by
moisture uptake and temperature. If the operating temperature is close to the resin glass-
transition temperature a significant drop in strength can be expected; moisture uptake lowers
the glass-transition temperature.
8.2 The moisture distribution associated with prolonged operation in the most adverse
environment world-wide can be represented by the equilibrium condition corresponding to a
relative humidity of about 85% [6].
8.3 Often it is impractical to degrade thick components to the equilibrium condition prior
to testing because of the time required to obtain a representative moisture content and
distribution. Instead superfactors (sometimes termed ‘knock-down’ factors) can be applied.
These superfactors can be determined from fatigue tests on coupons and structural elements
with artificial degradation and without. During degradation, travellers should be used to
monitor moisture uptake and specimens should reach at least 95% of the target equilibrium
condition. The factors so obtained can be refined in due course by reference to tests on
naturally-aged full-size specimens, or at least coupons and/or elements cut from them.
8.4 The effect of cyclic and prolonged exposure to both hot/dry and cold/dry conditions
must be considered.
8.5 Consideration must also be given to any deleterious effects of aircraft fluids which
may come into contact with the structure.
Page 5 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
9 INSPECTION-BASED SUBSTANTIATION
9.2 Special techniques may need to be developed to detect the occurrence of BVID or
to register potentially-damaging impacts.
9.3 There is a body of evidence showing that airframe structures containing BVID and
designed to the reduced-strain allowables currently employed do not exhibit significant
damage growth under gust and manoeuvre-dominated spectra. Test substantiation has
focused on demonstrating that damage of an agreed size does not grow under service
conditions. Arguably, such damage should represent defects/damage of ‘detectable size’
(see Leaflet 36) that would be unlikely to be missed in manufacture, or in service in the case
of components exposed to impact damage. Furthermore, since the damage is intended to
represent defects that might be present throughout the life, the damaged structure would
need to be capable of sustaining the Design Ultimate Load; some alleviation of this
requirement would be obtained if sufficient data could be provided to justify a probabilistic
approach, but this would require information on the probability distributions of crack
detection, residual strength and service loading.
REFERENCES
Page 6 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
TABLE 1
Page 7 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 41
FATIGUE LOAD METER INSTALLATIONS
GENERAL CONSIDERATIONS
1 PURPOSE
1.1 The fatigue load meter is intended to provide information that will
enable the consumption of the fatigue life of an aeroplane to be monitored, and
to give immediate warning when the aeroplane may have been overstressed.
2 THE FATIGUE LOAD METER
2.1 The fatigue load meter operates by counting the peaks of normal
acceleration, at the centre of gravity of the aeroplane, that are caused by
external forces in flight.
2.2 The Meter comprises essentially an accelerometer and a group of
counters to record the numbers of times that eight or more predetermined
accelerations are exceeded. For variable geometry aeroplanes, the meter must
separately record the number of times the predetermined accelerations are
exceeded at each of several structurally significant configurations or flight
conditions. It should be installed so that, as nearly as possible, the axis of the
accelerometer lies in the plane of symmetry of the aeroplane and is normal to
the flight path (normally within 5 degrees of vertical, when the aircraft is flying
straight and level). An automatic switch in the 24 volt electrical power supply is
employed to start the meter when the aeroplane has become airborne and to
stop it before the aeroplane lands, so that accelerations caused by contact with
the ground are not counted.
2.3 The accelerometer is a double spring-mass system with eddy-current
damping, designed to respond faithfully to oscillations up to a certain frequency
and to attenuate oscillations of higher frequencies. It has a good transient
response (i.e., a 5% to 10% overshoot after an applied step function). The
meter records absolute accelerations; thus straight and level flight corresponds
to 1.0 g. The meter will not be damaged by accelerations of 12 g.
3 INSTALLATION
3.1 The fatigue meter should always be installed as close as possible to
the centre of gravity of the aeroplane in order to minimise errors in the
measured accelerations. The distance of 1.5 metres, is the maximum that is
tolerable; usually the distance should not exceed one metre.
3.2 The counter display should be installed where it may be read easily
when the aircraft is standing on the ground.
3.3 The fatigue meter should be installed rigidly to the wing centre section,
or to a part of the structure that has an inflexible load path to the wing centre
section. Any flexibility in the connection between the meter and the centre of
gravity of the aeroplane has an adverse effect on the accuracy of
measurement. Therefore the meter should not only be attached firmly to rigid
aeroplane structure but it should be placed where structural modes of vibration
will not cause spurious inputs.
1 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
4.1.1 The switch should start the meter as soon as possible after the
aeroplane leaves the ground and should stop it in as short a time as
possible before the aeroplane lands. Unfortunately the switch cannot be
operated successfully by the extension and compression of the
undercarriage legs. During take-off the wheels remain briefly in contact with
the ground after the legs have extended completely, and the full leg
extension persists for a short time after the initial landing contact.
4.2 COMBAT AND TRAINER AEROPLANES WITH RETRACTABLE
UNDERCARRIAGE
2 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
3 of 3
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 42
STRUCTURAL MONITORING SYSTEMS
USING NON-ADAPTIVE PREDICTION METHODS
1 INTRODUCTION
1.1 Within this leaflet guidance is provided to support of the requirements
identified in Clauses 3.2.33 to 3.2.59 – Structural Monitoring using Non-
Adaptive Prediction Methods. The following topics are addressed:
(a) Definitions;
(b) Related Requirements;
(c) Exclusions;
(d) Task Definition;
(e) System Design;
(f) Training, Testing and Validation Data;
(g) Model Training;
(h) Model Validation;
(i) In-Service Maintenance.
2 DEFINITIONS
2.1 Non-Adaptive Prediction Method
A prediction method is defined as a set of coefficients/weights and a set of
transformation equations that operate on a set of inputs (such as flight
parameters e.g. normal acceleration, roll rate etc) to produce outputs that
approximate a target value (e.g. stress, strain, load or fatigue damage).
Prediction techniques use a range of mathematical or statistical methods that
may include neural networks, model-based analysis, linear or non-linear
regression, clustering algorithms etc. Non-adaptive prediction methods use a
fixed set of weights, which are evaluated through calibration/training using
target data that contains examples of inputs and target outputs. After training,
the coefficients of non-adaptive methods are fixed until the commencement of
any further training.
2.2 Supervised Learning
Supervised learning is a technique where the model is presented with input
data and examples of the required output data, or target. The relationships
within the model are adjusted, for example by iteration, to reduce an error
function. This error function is based upon the difference between the desired
output and that predicted by the model. There is a plethora of training
algorithms used to minimise the error by converging as quickly as possible
while ensuring that the model does not over fit the solution to the training data.
Well-known examples of supervised learning include Multi-Layer Perceptron
(MLP) and Radial Basis Function (RBF) Artificial Neural Networks (ANN).
1 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
2 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
3 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
6.2 Domain knowledge may also be applied to identify the system input
parameters that could improve the accuracy of the prediction method. For
example, if a vertical fin monitor was being developed, domain knowledge may
indicate that sideslip angle, lateral acceleration at the tail, yaw rate and yaw
acceleration are important parameters to include. Many of these parameters
have not routinely been included within an OLM programme parameter data
set. Therefore, early identification and documentation of this type of domain
knowledge for the structural monitor will assist in ensuring that necessary data
are captured within training, testing and validation data sets.
6.3 Each input or output to a model will be either numeric or categorical and
thought needs to be given to the data type and its proposed treatment. This is
particularly significant for categorical data where poor treatment can lead to
non-sense outputs. For example, weight-on-wheels can only be true or false (1
or 0). Therefore, a value of 0.8 is meaningless. In such cases, spread encoding
should be considered, where a single numeric value is transformed into a set of
category variables. This can be envisaged as a large array of pigeonholes,
each of which represents a category. This is particularly useful when the
variable is an output and a number of categories can be chosen. A spread
encoding technique can be used to identify the probability that the answer is in
each of the given possible categories. Examples of this technique could be
helicopter flight condition recognition algorithms.
6.4 Data normalisation methods used for example to match data to non-
linear activation functions and to prevent scale error effects and methods used
to reduce the dimensionality of the input data need to be documented. It is
often desirable to reduce the number of input variables to a minimum to reduce
the size of the training data set needed to produce a suitable solution. This
process is termed dimensionality reduction as each input or output parameter
provides an additional dimension to the input space. Dimensionality reduction
can be achieved in a number of ways. For example, input variables can be
removed that carry the least predictive power. However, in practice
identification of the least predictive variables in a multidimensional space can
be problematic. Input parameter significances can be determined by using
methods such as Automatic Relevance Determination (ARD). This method
uses a Bayesian framework to identify the relative significance of an input to
the network output based upon the overall variance in the weights relating to
each input parameter. Alternatively, methods whereby high-dimensional input
data are projected onto a lower dimensional space are often used. Principal
Component Analysis (PCA) is a well-known tool used for this process. The
input data are resolved into principal components or directions. Those
components that contribute most significantly to the variance of the data are
preserved. Another method is to use an Auto-Associative network as a feature
extractor. This method involves constructing a MLP ANN where the target
output is the same as the input. The hidden layer that contains the least
number of neurons (less than the number of inputs), acts as feature extractors.
The inputs to the prediction method are then generated by entering the full
input vector into the Auto-Associative network and extracting the outputs of the
hidden layer with the least number of neurons.
4 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
5 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
7.5 The training, testing and validation data sets need to be captured from
all practicable areas of the flight envelope. This is to ensure the models can
provide coverage for all flight regimes and to ensure that any significant
changes in the relationships between the model inputs (e.g. flight parameters)
and the model outputs (e.g. strains) are captured. This requirement is identified
as all practicable areas because it may not be possible to obtain flight loads
data from every location in the flight envelope and hence it is essential that
evidence of the coverage of the training, testing and validation data is also
provided.
7.6 Furthermore, where practicable, data from a number of aircraft should
be captured and used within the training, testing and validation process. These
data should be used to demonstrate that the models can provide a generalised
solution across more than one individual aircraft.
7.7 The possible sources of flight data to train, test and validate structural
models are largely restricted to data from Flight Loads Measurement (FLM),
Operational Loads Measurement (OLM) or Operational Data Recording (ODR)
programmes. However, additional data could be obtained from Ground
Resonance Testing (GRT), analytical loads models or Finite Element Models
(FEM). FLM data can exercise areas of the flight envelope not routinely visited
during in-Service operations. Therefore, FLM data could be used to train initial
models to predict the underlying data relationships required for the monitor and
once OLM data becomes available additional training could be undertaken as
required.
7.8 The statistics of the training, testing and validation data need to be
identified and documented. For numeric data this may include the mean,
standard deviation, minimum and maximum values and for categorical data this
may include the range of different categories and the population densities.
Basic statistical data, such as mean, standard deviation, minimum and
maximum values provides the first indication that the training, testing and
validation data are consistent with each other.
7.9 There is no straightforward method of determining how many data are
required to train and test a predictive structural model. In general, the quantity
of data required is governed by the number of training cases that will be
needed to ensure the network performs adequately. A training case, for
supervised learning, is defined as one input vector and one output vector. In
turn, the number of training cases is defined by the nature of the application,
the intrinsic dimensionality of the data, the resolution required and the
frequency distribution of the data. Additionally, the quality of the data and
presence of ambiguous inputs also needs to be considered. Most importantly,
evidence of the coverage of the structural monitor from training data needs to
be provided.
7.10 For a complex system in which there is a significant variation between
nearby points within the input or output space, data that are significantly
different to smooth interpolation may not be modelled correctly. More complex
problems may require more complex models and more data to train with to
provide adequate generalisation capability.
6 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
7.11 Analysis of the frequency distribution of input and output data is a useful
tool in identifying the range and density of the training, test and validation sets.
These distributions can be used to select the most appropriate data for training,
testing and validation. For example, high strain variations are more significant
in fatigue damage terms than steady-state strains. Therefore, it would be ill-
advised for benign flying to dominate the training sets.
7.12 Methods used to check the integrity of the data, including anomaly
detection routines and statistical techniques, need to be applied across
training, testing and validation data sets and documented accordingly.
Assessment of the quality of the data to be presented to the model is essential.
Data anomalies, such as range exceedences, dropouts, sensor failures, datum
shifts and bad blocks should be identified and removed. All raw data should be
stored and a record of all data anomalies should be retained (detailed in next
paragraph). Flight data will not be perfect and where there is noise in the data
then a larger quantity of data will be required to train the model.
8 MODEL TRAINING (Clause 3.2.48 and 3.2.49)
8.1 The function of the training algorithm is to minimise the error between
the predicted output and the required output. Generally, the smaller the error,
the better the performance of the model. This error is calculated either each
time a new data set is presented to the model (on-line mode), or at each pass
(epoch) through the training data (batch mode). These errors are used to
update the internal relationships within the model. During or after training, the
model(s) is/are tested on a data set (test set) to indicate the performance on
unseen data (to confirm generalisation) or to choose an optimal model.
8.2 For some techniques, such as ANNs, a randomisation function may be
used to set the initial weights for training. In this case, a number of models are
required as some may get stuck in local minima on the error surface and the
risk of using a model in this condition is reduced if several models are trained
and the model best performing on the test data is selected. Thereafter, only the
best performing model would be used to compute the performance metric using
the validation data.
8.3 There are numerous training algorithms used for non-adaptive
prediction methods. Broadly, these training algorithms work by presenting each
pattern of data to the network and either at each data presentation of after the
entire set (epoch) has been presented the model is updated based upon the
error criteria used. The most appropriate method will depend on the
application. Whichever method is chosen by the designer, it needs to be
documented clearly and unambiguously together with supporting material
including the performance of the selected non-adaptive prediction method on
training and testing data. As well as the algorithm itself, the methods used to
prevent effects such as over fitting to the training data need to be documented
clearly.
9 MODEL VALIDATION (Clause 3.2.50 to 3.2.52)
7 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
8 of 8
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 43
GUST
PROBABILISTIC ANALYSIS
1. INTRODUCTION
1.1 The gust amplitudes quoted in the discrete gust cases are set at levels based on
collected gust data and which have given satisfactory gust strength on previous
aeroplanes. The loads produced by gusts of these amplitudes are at the Design Limit
Load level that is to be expected once in a specified time (e.g. aircraft lifetime).
1.2 The probability of meeting a severe gust depends on the aircraft altitude,
geographical terrain and weather patterns. Whether or not that gust generates loads in
excess of limit load on the airframe structure depends on the aircraft speed when the
gust is encountered, the gust amplitude and sharpness and the configuration of the
aircraft – mass, fuel quantity, payload, store load, etc.
1.3 The rigid application of the gust design cases may result in an aircraft being
judged unacceptable and consequently having a limit imposed on the flight envelope or
requiring an excessive weight penalty to strengthen the structure. These penalties have
a overall impact on the aircraft, whereas it is possible that there is only a low probability
of encountering a severe gust in the particular combination of flight conditions which
infringe the gust design case.
1.4 The Gust Probabilistic Analysis quantifies the overall risk of encountering a gust
which reaches a limit load capability level during the planned full operational usage of
the aircraft, which yields the frequency with which limit load level will be reached. This
frequency is considered acceptable if it implies a period between encounters which is at
least as long as the specified life in the definition of the Design Limit Load.
1.5 It may be noted that such probabilistic analysis can also be used to maximise the
level of clearance offered to an aircraft of a given gust capability while maintaining the
risk of a limit load gust encounter at acceptable levels. The analysis can also provide an
insight into the implications of sortie planning to minimise gust risk
1.6 The overall process is illustrated in figure 1 and is described in the following
paragraphs. Figure 2 shows more detail for the process applied to one flight condition. A
discussion of the effect of parameter variations is included in reference 1 and the generic
aircraft model used to generate examples in this leaflet is described in reference 2.
2. SORTIE DEFINITION
2.1 The starting point for the process is the definition of operational usage of the
aircraft. The military operator often defines utilisation in a Statement of Operating Intent
and Usage (SOIU), expressed in terms of a number of different sorties reflecting specific
operational duties and the fraction of total aircraft flight time expected to be devoted to
each duty.
Page 1 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
2.2 For the purpose of the gust analysis greater detail is necessary, so that a range
of variations of the same sortie type may be defined to cover, for example, different flight
duration, take-off fuel, and payload. Each sortie variation will have a probability
associated with it, represented by a predicted number of sorties with those conditions
during a period of operation. The variations of sorties are necessary to ensure that the
effects of the most severe operating conditions are included in the analysis. For
example, flight with low fuel and high payload may entail a much higher gust risk than
the same flight with average fuel and payload weights. It is important to include sorties
which take the aircraft into the extremes of the flight envelope, with a wide range of
values of take-off fuel and payloads studied. The improbability of operating sorties at
these conditions will be accounted for via a small number of flights associated with that
combination of parameters.
2.3 The effect of different flight duration cannot be achieved by a linear scaling of
similar flights, since duration is a significant variable in the analysis for two reasons. One
is the effect of fuel consumption, so that the aircraft will be carrying less fuel at the end
of a longer flight with the same take-off conditions. The other effect is the modified
balance between flight at different altitudes: on a shorter flight a higher proportion of total
time will be spent in, for example, the descent and landing phases.
2.4 In general, take-off and climb will be defined by handling and performance
requirements and have been taken to remain fixed at the baseline values for the initial
phase of all flights of a particular sortie type, subject to a possible reduction of rate of
climb when flying at high gross weights. Similarly, descent, approach, and landing are
likely to be fixed in time relative to the end of the flight. The adjustments to duration after
climb and before final descent should be appropriate to the operational task in the sortie.
For example, a transit or transport flight is easily modified by suitable reduction or
extension of cruise, possibly with altitude changes as appropriate. A maritime patrol
mission, with low altitude patrol interspersed between higher altitude transits, can be
extended by introducing additional cycles of patrol and transit.
2.5 There is also the possibility of variable discrete changes of weight during the
flight, from causes such as air-dropping of some payload, release of stores, and
receiving or dispensing fuel in air-to-air refuelling. These must be included as further
variations of the basic sorties.
2.6 On an existing aircraft, operational data should be available to guide the likely
range of variations in all these sortie parameters. For the analysis of a new design,
assumptions will have to be made on the basis of operational experience with similar
aircraft. In both circumstances these data will have to be checked by operational
recording on the aircraft in service and a revised analysis undertaken if significant
utilisation changes become apparent.
Page 2 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
3. SORTIE SUBDIVISION
3.1 The variations of aircraft operations have now been defined in terms of a
multiplicity of sorties, each having a specific flight duration, take-off fuel and payload,
and possibly discrete mass changes during the flight. The SOIU often defines the
baseline sorties in terms of altitude and airspeed variations between a number of key
events during the flight, as shown in figures 3 and 4.
3.2 The probabilistic gust analysis requires the sortie to be broken down into small
segments within which there are only small changes in the aircraft gust response
characteristics and in the frequency of gusts in the atmosphere. The breakpoints in the
sortie definition are very likely to be too coarse for the subdivisions needed to meet
these requirements.
3.3 The altitude variations of frequency of occurrence of two gust velocities are
shown in figure 5 and table 1, taken from reference 3. At low altitude the gust
frequencies (on a logarithmic scale) shows a rapid decrease with increasing altitude and
this drives a requirement for fine subdivisions of altitude at low altitude. It has been
found that the use of 1000 ft altitude bands up to 4000 ft has been the minimum required
to give an adequate prediction of the low altitude gust risk, with the gust frequency for
each band evaluated for the mean altitude of the band. The preferred choice of altitude
bands above 4000 ft is to use bands of 2000 ft up to 10000 ft and 2500 ft bands at all
higher altitudes.
3.4 By comparison the need for subdivision of the flight due to the variation of the
gust response is less significant and only has an effect during cruise at constant altitude,
when the reduction of weight due to fuel consumption may become the determining
factor on segment duration. For a large aircraft a segment time of 10 minutes during
cruise gives a negligible error in the gust response compared to the precise value. For
combat aircraft, with a reduced impact of fuel on the loads, a longer interval may be
acceptable, but in this case it is more likely that speed or altitude variations may
intervene in the sortie definition. Indeed, for a combat aircraft the organisation of the
analysis may be more oriented towards the time spent in specific point conditions within
the envelope rather than the steady progression through a flight profile which is more
appropriate for a large or transport type of aircraft.
3.5 To illustrate the subdivisions required for the bands described here, figure 6
shows the segmentation of a simple short sortie typical of transport operations (i.e. a
one-hour flight with climb to 27000 ft, cruise, descend and land). This sortie is
subdivided into 32 segments. More segments will be required to model a longer sortie or
a complex military sortie in which a considerable proportion of flight time is spent at low
altitude and with frequent changes of flight condition.
Page 3 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
Page 4 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
Page 5 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
ratio of overswing/
1 0.7 0.6 0.5
direct response
maximum error in
50% 6% 1.4% 0.2%
gust frequency
5.8 These results emphasise the need to include the overswing response for highly
oscillatory quantities and the error at 60% or less overswing is well within the overall
accuracy of the complete process. This may be used as a guideline for ignoring
overswing properties on a specific quantity if the simplification of analysis is considered
significant.
5.9 For each load quantity the selection process gives a pair of gust velocities which
produce a limit load, one for an up-gust and one for a down-gust. These limit load gust
velocities are calculated separately for all the loads considered on the aircraft for each
gust length.
6. GUST PROBABILITY
6.1 Having calculated the gust velocities which produce limit load it is now necessary
to predict the probability of encountering a gust of this magnitude. This stage of analysis
is repeated for each of the gust lengths considered applicable to the aircraft under
review and for each load quantity. The minimum gust length should be 15.20m (50 ft) as
specified in Clause 3.5.9 The maximum gust length should be either 100m (305 ft) or
double the gust length which produces the maximum tuned gust response if this is
longer than 50m. Across this defined range, it is recommended that at least 8 gust
lengths be analysed.
6.2 It is necessary to account for the effect of gust length on the gust probabilities.
This is carried out via an assumption that gusts of different lengths H are equally
probable when the gust velocities are related by the H1/6 power law. The H1/6 law was
derived by Jones 4, 5 as the most realistic model of gust characteristics in severe
turbulence. In the analysis the gust velocities are adjusted to the velocity of an equally
probable gust with a standard reference length, i.e.
U R =U G * (H/H ref )1/6
where U G is a limit load gust velocity from the calculation of response to a gust of length
H, H ref is the reference gust length and U R is the velocity of an equally probable gust
with length H ref . A family of gusts with amplitude related to gust length by the H1/6 law is
illustrated in figure 9, all the gusts shown having the same probability of occurrence as a
20 m/sec gust with length 40m.
Page 6 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
6.3 A review of gust data is given in reference 1, combining records from early RAE
data and more recent records from 747s in the service of two airlines. The
recommended gust frequency models which resulted from these studies are biased
towards the older RAE data below 15000 ft, where data are reckoned to be more
representative of the environment for military operations, and blend into the data from
747s by 25000 ft. The frequency distributions for 50 ft/sec (15.24 m/sec) and 66 ft/sec
(20.12 m/sec) gusts from reference 5 are shown in figure 5 and the values are given in
table 1. Note that the data has been standardised for a gust length of 40m and this value
should be used for H ref in the analysis.
6.4 In order to find the probability of any gust velocity, interpolation is used on both
altitude and gust velocity. Interpolation is linear with altitude and gust velocity used with
the logarithm of the frequencies of occurrence from table 1. First, for each of the two
basic gust velocities 15.24 and 20.12 m/sec the altitude variation of the frequencies is
removed by interpolating to the altitude of the mission segment under analysis:
Y= Y 1 + (Y 2 -Y 1 ) * (A - A 1 ) / (A 2 - A 1 )
where A is the required altitude, A 1 and A 2 are two altitudes from table 1 (one above and
one below A) and Y 1 =log(F 1 ), Y 2 =log(F 2 ) with F 1 and F 2 is the frequency of an up or
down gust at the altitudes A 1 and A 2 (for either 15.24 m/sec or 20.12 m/sec gusts). Note
that the frequencies defined in table 1 are for up and down gusts. Since this analysis
accounts for the different response of the aircraft to up and down gusts, it is necessary
to divide the values in table 1 by 2 to obtain the probability of an explicit up gust or down
gust.
6.5 The result Y = log(F) defines the frequency of the basic gust at this altitude. A
second interpolation is then used to find the frequency of the required gust velocity U R
(normalised to the gust reference length).
Y= Y A + (Y B -Y A ) * (U R - U A ) / (U B - U A )
where U A and U B denote the basic gust velocities 15.24 and 20.12 m/sec and Y A and Y B
are the corresponding values of log frequency at the required altitude. The exponential
of the result Y gives the frequency of the desired gust velocity at the specific altitude of
the mission segment.
6.6 For a single gust length, the two different probabilities of a up gust and a down
gust are added together. Since the gusts have been normalised to an equi-probable
family of gusts the different gust lengths may be handled by finding the mean probability
over all the gust lengths in the analysis (i.e. the sum of all the probabilities divided by the
number of gust lengths in the analysis). The calculation of gust probability in this flight
segment is repeated for each load quantity.
Page 7 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
6.7 There is the possibility that the aircraft would stall before generating the lift
required to produce a limit load, most likely for a slow flying aircraft on which the gust
velocities represent a greater proportion of the flight speed and hence higher
incremental aerodynamic incidence. In these cases it is logical to keep track of stall
boundaries in the extrapolation to find the gust velocity which produces a limit load. If a
limit load could not be generated in the conditions of this flight segment, the probability
of encountering a gust which generates a limit load on the structure is zero. This value
should be included in the structural risk analysis. However, a risk of encountering this
severe gust can be quantified and this should be recorded for the risk assessment of
stability or handling issues which might arise from such an upset.
Page 8 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
9. GUST DIRECTION
9.1 The process description has concentrated on a single gust direction. The major
interest for gust load design on a large aircraft is the vertical gust, which relate closely to
the normal accelerations which were measured on aircraft collecting the gust data used
in the analysis. The same considerations apply to lateral gusts and the method can be
applied directly to lateral gust load prediction. However, there would be difficulties in
extending the probability analysis to encompass gusts from other directions. The relative
importance of gusts from other directions should be explored against the basic design
cases.
LEAFLET REFERENCES
1. Kaynes I W. Probabilistic gust analysis example. QINETIQ/FST/CR031374, March
2003
2. Patel H. Generic Gust Response Investigation. QINETIQ/FST/CR030991, February
2003
3. Kaynes I W. Flight loads data for gust load prediction.
QINETIQ/FST/SMC/CR021764, March 2002
4. Jones J G. A theory for extreme gust loads on an aircraft based on the
representation of the atmosphere as a self-similar intermittent random
process. RAE TR68030, February 1968
5. Jones J G, Foster G W, Haynes A. Fractal properties if inertial range turbulence with
implications for aircraft response. Aeronautical Journal, October 1988.
Page 9 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
0 2.0E-5 3.0E-6
Table 1 - Gust probabilities for 15.24 and 20.12 m/sec (50 and 66 ft/sec) gusts
Page 10 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
Page 11 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
Page 12 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
30000
25000
20000
altitude (ft)
15000
10000
5000
0
0 10 20 30 40 50 60
time (min)
350 350
300 300
250 250
airspeed (kt EAS)
airspeed (kt EAS)
200 200
150
150
100
100
50
50
0
0 0 10 20 30 40 50 60
0 10 20 30
time (min)
40 50 60
time (min)
Page 13 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
1.00E-04
1.00E-05
Up and down gusts per n.mile
1.00E-06
1.00E-07
1.00E-08
1.00E-09
1.00E-10
1.00E-11
1.00E-12
0 10000 20000 30000 40000
Altitude (ft)
30000
25000
20000
altitude (ft)
15000
10000
5000
0
0 10 20 30 40 50 60
time (min)
Page 14 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
160
120
bending moment
80
40
0
0.0 0.5 1.0 1.5
-40
-80
time (sec)
response to Ug=15
response for Ug=20.2, upper ALE limit
response for Ug=98.2, lower ALE limit
upper ALE
lower ALE
a) Response quantity 1
50
bending moment
25
0
0.0 0.5 1.0 1.5
-25
time (sec)
response to Ug=15
response for Ug=27.8, upper ALE limit
response for Ug=19.5, lower ALE limit
upper ALE
lower ALE
b) Response quantity 2
Figure 7 – Up gust velocities required to produce limit loads for two sample response
quantities
Page 15 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
a) Response quantity 1
b) Response quantity 2
Figure 8 – Down gust velocities required to produce limit loads for two sample response
quantities
Page 16 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
Page 17 of 17
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 44
1 INTRODUCTION
1.1 Structures are at a threat from impact damage from a number of sources, during
manufacture and in service. Impact damage can have consequences both for the immediate
structural integrity and the supportability of structures. Composite structures require
particular consideration of this type of damage, because while there may be little or no
outward sign, the structural integrity of the structure can be significantly reduced.
1.2 This Leaflet gives guidance on the certification and qualification route for composite
structures, with particular reference to impact damage resistance. Composite aircraft
structures in this case means structures fabricated from continuous fibre-reinforced polymer
matrix materials, for example, carbon, boron, aramid and glass reinforced plastics. The
guidance is equally applicable to monolithic and sandwich structures.
1.3 Compliance with the requirements of Clause 3.1 (Static Strength) and Clause 3.2
(Fatigue) must be demonstrated. Guidance on acceptable means of compliance is given in
the Leaflets accompanying these Clauses. The impact damage resistance issues and
associated residual strength requirements are covered in this Leaflet, but fatigue and
damage tolerance issues are not, although some guidance on this is provided in Leaflet 40,
which accompanies the fatigue design requirements of Clause 3.2 Damage resistance and
damage tolerance differ in that the former quantifies the damage caused by a specific
damage event, while the latter addresses the ability of the structure to retain the required
residual strength, in the presence of specific damage.
1.4 Comprehensive guidance and discussion of impact damage resistance is also given
in Chapter 7, Volume 3 of The Composite Materials Handbook Mil 17 (formerly know as Mil-
Hdbk-17) [1], which describes means of compliance likely to be acceptable to US military
certification authorities. AC20-107A [2] presents an acceptable means of compliance with
civil certification requirements; essentially the same information is contained in CS-25 [3] at
AMC No. 1 to Paragraph 25.603. For a discussion of certification issues for composite
structures, see Chapter 7 of [1].
Page 1 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
2.2 Different areas of the air vehicle will be subject to different types and levels of
impact threat. Therefore, an impact threat assessment should be done, to determine which
areas of the air system are susceptible to particular types of impact damage. For example,
lower surfaces of the air vehicle (the underside of the airframe, the landing gear, and external
stores) may be more vulnerable to runway stone damage, while upper surfaces may be more
susceptible to damage by dropped tools or equipment. Edges may also be vulnerable to
impacts, for example those of control surfaces, or of panels; where this is identified as a risk,
it should be taken into consideration. It should also be considered that removable structural
parts could be damaged during maintenance, or storage off-aircraft.
2.3 Having identified the likely sources and locations of impact damage, probabilistic
analysis of the impact threat should be made, in terms of the likelihood of occurrence. For
these threats, impact energy, velocity and mass/geometry of the impact source should also
be taken into account, coupled with the probability of the aircraft encountering an extreme
load. The impact resistance criteria for the composite structures under consideration should
be selected on the basis of this assessment.
2.4 The criticality of the structure should also be taken into account when assessing the
consequences of an impact. The guidance given here is principally aimed at Grade A
structure (see Part 1, Section 4, Clause 4.1).
*
In this Leaflet, “blunt” and “sharp” refer to the geometry of the impactor. A blunt impactor will generally have a high
radius of curvature, whereas a sharp impactor will have a low radius of curvature, or sharp edges.
Page 2 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
3.3.1 It may be desirable, depending on the type of aircraft and the desire to trade
off inspections against the increased structural weight brought about by more robust
structure, to allow visually-detectable damage from ‘exceptional’ impacts. Where a
visible impact criterion is used, a level of damage is defined, which is on the verge of
being visually detectable. In such circumstances, ‘exceptional’ damage should be
detectable as Barely Visible Impact Damage (BVID), detectable during routine or
periodic inspections or Clearly Visible Impact Damage (CVID), which can be found
before the next flight (see also Leaflet 40). Damage visibility criteria should be set and
justified on a project-specific basis, as the visibility of damage may vary with, for
example, surface finish, and lighting conditions during inspection.
3.3.2 Damage visibility criteria are more often used in civil aircraft, but can negate
the requirement to characterise the impact threat fully. The setting of the damage
visibility criterion is also subjective, and its use may be less appropriate for military
aircraft, where visual inspections may have to be carried out in less than ideal
conditions. These conditions may lead to a greater dent dept being chosen, leading to
a heavier structure, which in turn may not be practical for performance reasons.
3.3.3 Furthermore, dent depths depend on a number of variables, including impact
source shape, the type of structure, and the location on the structure. For many blunt
impacts, a very shallow dent, which may relax with time, or no dent at all, can result; in
either case, considerable delamination may occur, with consequent loss of strength,
without any dent being visible. For this reason, a visible damage criterion may not be
appropriate for ‘routine’ levels of impact damage. In some structures, planar damage
size (such as may be detected by non-destructive examination) may give an indication
of the loss of structural integrity due to impact (see, for example, [5]), but this should be
demonstrated as part of compliance with the requirements.
3.3.4 In the energy cut-off approach, it is assumed that the damage threshold
energy (the impact energy level at which damage first forms) of the structure is chosen
to be greater than that imparted by the pre-characterised threat. This relies on the
accurate characterisation of the impact threat. The probability of a large impact event
may be quite low, however, and designing against this threat may cause the structure
to be overly heavy. The approach may also lead to a “window of uncertainty”, where
impact damage may occur just above the chosen threat level, but just below that which
would be readily visible. Non-destructive inspections would need to be implemented in
service to ensure that damage, which would compromise structural integrity, has not
occurred within this “window of uncertainty”. Appropriate non-destructive inspection
methods should be chosen and substantiated against realistic damage scenarios.
3.4 The investigation of the advantages and disadvantages of the methods above,
coupled with a consideration of the damage tolerance criteria, will lead to a selection of the
most suitable design approach. The selection of the most suitable approach will be a trade-
off between structural weight and supportability considerations.
Page 3 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
Page 4 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
4.3 Consideration should be given to how any damage introduced into specimens
before testing is to be measured and characterised, pre- and post-test. Measurement of
damage generated by testing must be correlated with additional parameters, which can be
used to judge residual strength.
5 RESIDUAL STRENGTH
5.1 The results of the probabilistic impact threat assessment, and the selection of a
design and certification/qualification approach will be used in establishing impact resistance
test conditions. These should be expressed in terms of the probability of occurrence of the
impact event, associated with the probability of the aircraft encountering an extreme load.
Following a routine, ‘probable’ † , impact event, the structure must be able to sustain Ultimate
Load. After an exceptional, ‘incredible’, event, the structure must be capable of sustaining 1.2
Limit Load (this is partly for consistency with the residual strength criterion of the fatigue
design requirements ‡ ).
5.2 Where the concepts of BVID and CVID have been applied to exceptional impact
events, the following criteria apply. Structure which has been subjected to an impact or
impacts resulting in BVID should be shown, by analysis and test evidence, to be capable of
sustaining the design Ultimate Load. Where flaw tolerance (no growth) is to be
demonstrated, residual static strength testing should be carried out after repeated load
cycling. Structure which has been subject to an impact or impacts resulting in CVID should
be shown, by analysis and test evidence, to be capable of withstanding 1.2 Limit Load § .
This method is to be used in conjunction with a flaw tolerant or fail safe approach that
produces either an enhanced safe life, or an on-condition inspection interval (see also Leaflet
40).
5.3 For all impacts where the probability of occurrence is shown to be improbable,
impact testing should be carried out to demonstrate that in the damaged state, the static
strength requirements of Clause 3.1 and the fatigue requirements (particularly the residual
strength criterion) of Clause 3.2 are met. The functional requirements of Clause 3.1 shall also
be met.
†
See Definitions.
‡
See Clause 3.2.16 for guidance on circumstances where the residual strength factor on Limit Load may be reduced.
§
See Clause 3.2.16 for guidance on circumstances where the residual strength factor on Limit Load may be reduced.
Page 5 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
6.3 Where impact resistance assessment criteria have been established on the basis of
probabilistic considerations, aircraft usage and exposure to impact threat must be monitored
and analysed in service to ensure that the assumptions underlying the analysis remain valid,
and that the impact threat has not changed. Feedback from in-service inspections should
also be used to validate the assumptions concerning impact threat levels.
DEFINITIONS
Probable – Will occur to a significant proportion of aircraft during the life of a particular fleet.
Probability of occurrence between 10–3 to 10–4 per flying hour for combat aircraft and
between 10–4 to 10–5 per flying hour for combat support aircraft and helicopters.
Improbable – Remote likelihood of occurrence during the operational life of a particular fleet.
Probability of occurrence between 10–6 to 10–7 per flying hour for combat aircraft and
between 10–7 to 10–8 per flying hour for combat support aircraft and helicopters.
Incredible – Less likely than improbable. Probability of occurrence less than 10–7 for combat
aircraft and 10–8 for combat support aircraft and helicopters.
REFERENCES
1. American Society for Testing and Materials International. The Composite Materials
Handbook Mil 17 Volume 3. Polymer Matrix Composites: Materials, Usage, Design and
Analysis. Mil-Hdbk-17-3F, August 2002.
2. Federal Aviation Administration. Composite aircraft structure (Advisory Circular). AC20-
107A, 25 April 1984.
3. European Aviation Safety Agency. Certification Specifications for Large Aeroplanes.
CS-25, Amendment 2, 2 October 2006.
4. Advanced certification methodology for Composite Structures. Report No.
DOT/FAA/AR-96/111, April 1997.
5. Impact Damage Characterization and Damage Tolerance of Composite Sandwich
Airframe Structures. Report No. DOT/FAA/AR-00/44, January 2001.
Note: it may be possible to simulate high velocity impacts by changing the mass/shape of a
standard test set-up, providing it can be shown that the damage mode is equivalent or worse.
Page 6 of 6
DEF STAN 00-970 PART 1/15
SECTION 3
LEAFLET 45
1 INTRODUCTION
1.1 This Leaflet gives guidance on the requirements to be followed in evaluating the
structural performance (that is, the capability of the aircraft to meet its specified structural
requirements) of aircraft (aeroplanes, helicopters and remotely piloted aerial vehicles) with
systems, the failure or malfunction of which during a mission may influence safe flight and
landing. System failure conditions for the purposes of this Leaflet are those that affect
structural performance and induce loads, change the response of the aircraft to inputs such as
gusts or pilot actions, or lower flutter margins. These requirements are based on Appendix K
to CS 25 [1], with changes appropriate to operation in the military environment.
1.2 The criteria given in this Leaflet must be used to demonstrate compliance with the
requirements of Clause 3.1.3 and 3.1.4 (Static Strength). They are intended for application to
aircraft equipped with flight control systems, autopilots, stability augmentation systems, load
alleviation systems, flutter control systems (see also Part 1, Section 4, Clause 4.8), and fuel
management systems. If the requirements of this Leaflet are applied to other systems, it may
be necessary to adapt the criteria to the specific system.
1.3 The criteria defined in this Leaflet only address the direct structural consequences of the
system responses and performances and cannot be considered in isolation, but should be
included in the overall safety evaluation of the aircraft. These criteria may in some instances
duplicate standards already established for this evaluation. These criteria are only applicable
to structure whose failure could prevent continued safe flight and landing. Specific criteria that
define acceptable limits on handling characteristics or stability requirements when operating
the system-degraded or -inoperative mode are not provided in this Leaflet.
1.4 Depending on the specific characteristics of the aircraft, additional studies may be
required that go beyond the criteria provided in this Leaflet, in order to demonstrate the
capability of the aircraft to meet other realistic conditions, such as alternative gust or
manoeuvre descriptions, for an aircraft equipped with a load alleviation system.
1.5 The criteria prescribed in this Leaflet should be used in determining the influence of a
system and its failure conditions on the aircraft structure.
1.6 For specific guidance on the requirements to be followed in evaluating the aeroelastic
stability of aircraft with systems, the failure or malfunction of which during a mission may
influence safe flight, see Part 1, Section 4, Clause 4.8, and Part 1, Section 4, Leaflets 23 and
97.
1.8 When the system under consideration is fully operative, the following criteria apply:
Page 1 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
(a) Limit loads must be derived in all normal operating configurations of the
system from all the limit conditions defined for the aircraft (consistent with the
requirements of Clause 3.1.2), taking into account any special behaviour of such a
system or associated functions or any effect on the structural performance of the
aircraft that may occur up to the limit loads. In particular, any significant nonlinearity
(rate of displacement of control surface, thresholds or any other system
nonlinearities) must be accounted for in a realistic or conservative way when
deriving limit loads from limit conditions.
(b) The aircraft must meet the strength requirements of Clause 3.1 (with respect
to static strength and residual strength), using the specified factors to derive
ultimate loads from the limit loads defined above. The effect of nonlinearities must
be investigated beyond limit conditions to ensure the behaviour of the system
presents no anomaly compared to the behaviour below limit conditions. However,
conditions beyond limit conditions need not be considered when it can be shown
that the aircraft has design features that will not allow it to exceed those limit
conditions.
(c) The aircraft must meet the aeroelastic stability requirements of Part 1, Section
4, Clause 4.8; specific guidance on flutter considerations is given in Part 1, Section
4, Clause 4.8 and Part 1, Section 4, Leaflets 23 and 97.
2.1 For any system failure condition not shown to be extremely improbable (i.e., so unlikely
as not anticipated to occur during the entire operational life of all aircraft of the type), the effect
of the failure on the structure in terms of static and residual strength must be determined, for
specified flight conditions. Two states should be addressed: the conditions prevailing at the
time of failure and immediately thereafter, and continued flight. For continued flight, one of two
situations should be considered: one where the pilot does not know that a failure has
occurred, and the other where the pilot is aware of the failure and a limited flight envelope can
be attained.
2.2 Load cases to be considered for each of these states are set out in the sections below.
3.1 The initial conditions to be considered at the time of failure should be established on the
basis of a realistic scenario, including pilot corrective actions, to determine the loads occurring
at the time of failure and immediately thereafter.
3.2 For the civil certification requirements, it is deemed sufficient to start from a 1g level
flight condition. However, this is unlikely to be realistic for military aircraft, and analysis should
therefore be done to establish and justify the criteria to be applied.
3.3 The criteria to be applied may be determined by a consideration of the proportion of time
spent at different manoeuvre conditions, to give a probabilistic assessment of the condition at
time of failure, and the resultant load levels.
Page 2 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
(b) for tactical and training configurations: 1±0.25g (covering nominally 1g conditions),
pull-up at 2/3*n1 g, and 40˚ turn;
3.5 When the loads have been established, the static strength, residual strength and
aeroelastic response (see Part 1, Section 4, Clause 4.8 and Part 1, Section 4, Leaflet 97)
must be substantiated.
3.5.1 For static strength substantiation, these loads, multiplied by an appropriate factor
of safety that is related to the probability of occurrence of the failure, are ultimate loads
to be considered for design. The required factor of safety (FS) is defined in conjunction
with the probability of occurrence per hour of the failure mode j, Pj.
3.5.2 Failures with Pj ≥ ‘once per lifetime’ shall be protected by a factor of safety of 1.5.
Remote failure conditions down to 10–9/hr shall be considered, for which a smaller factor
of safety, down to 1.25, may be used. The ‘lifetime’ to be considered shall not be less
than the specified life of the aircraft. The impact of a life extension should also be
considered.
3.5.3 Figure 1 shows how the factor of safety is to be defined, using the specified life of
a typical transport aircraft as an example. In this case, the aircraft has a ‘lifetime’ of
30000 hrs, so that the breakpoint becomes (3.33 × 10–5, 1.5) and the end points are (1,
1.5) and (1 × 10–9, 1.25): this is illustrated by the blue line in Figure 1. For an aircraft with
a ‘lifetime’ of 6000 hours (typical for a combat aircraft), the breakpoint would be (1.67 ×
10–3, 1.5) and the end points remain at (1, 1.5) and (1 × 10–9, 1.25). Values for rotorcraft
and remotely piloted air vehicles should be derived in a similar manner.
3.5.4 For life extension of large aircraft, the value of Pj to be considered should be that
of the civil requirements of CS-25, i.e., those shown by the black line in Figure 1. For
combat aircraft, a value of Pj commensurate with a life extension of at least 100%
should be considered. Thus, for the typical combat aircraft from the example above, the
breakpoint would be (8.33 × 10–4, 1.5) and the endpoints are (1, 1.5) and (1 × 10–9,
1.25). Values for other aircraft types should be derived in a similar manner.
3.5.5 For residual strength substantiation, the aircraft must be able to withstand 80% of
the ultimate loads defined above. For pressurised cabins, these loads must be
combined with the normal operating differential pressure. The value of 80% of ultimate
loads is consistent with the residual strength criteria of Clause 3.2, and is selected
because military aircraft are used to the extremes of their manoeuvre envelopes. Those
parts of the flight envelope which do not conform to this criterion shall be advised.
3.5.6 Freedom from aeroelastic instability must be shown up to the speeds specified in
Part 1, Section 4, Clause 4.8
3.5.7 Failures in the system that result in forced structural vibrations (oscillatory failures)
must not produce loads that could result in detrimental deformation of primary structure.
Page 3 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
4.1.1 These conditions are to be considered, if the failure mode is such that the pilot is
unaware of the failure, and therefore would not apply manoeuvre limitations appropriate
to the failure condition.
4.1.2 For the continuation of the flight, with the aircraft having the system in a failed
state, and taking into account any appropriate reconfiguration, the following apply.
4.1.3 The loads derived from the following conditions at speeds up to V H must be
determined:
(a) the specified limit symmetrical manoeuvring conditions (see also Clause 3.3);
(b) the specified limit gust and turbulence conditions (see also Clause 3.5);
(c) the limit rolling conditions and the limit unsymmetrical conditions specified for
the aircraft (see also Clause 3.4);
(d) the specified limit yaw manoeuvring conditions (see also Clause 3.4);
4.1.4 For static strength substantiation, each part of the structure must be able to
withstand the loads specified in Paragraph 5.1.3 above, multiplied by a factor of safety
depending on the probability of being in this failure state. The factor of safety is defined
in conjunction with the probability of being in the failed condition j, Qj (where Qj = (Tj)(Pj)
and Tj = average time in hours spent in the failed condition).
4.1.5 The factor of safety in the failed condition is 1.5 for Qj = 1, reducing to 1.25 for
Qj = 1 hour / ‘lifetime’, thence reducing to 1.0 for Qj = 1 × 10–9. Figure 2a shows how the
factor of safety may be derived, using the transport aircraft with a ‘lifetime’ of 30000 hrs
as an example: the breakpoints become (1, 1.5), (3.33 × 10–5, 1.25) and (1 × 10–9, 1).
The end points at Qj = 1 and Qj = 1 × 10–9 remain unchanged.
4.1.6 For aircraft having different ‘lifetimes’, the breakpoints should be derived in an
analogous manner.
Note: If Pj is greater than ‘10 / lifetime’ per flight hour, then a 1.5 factor of safety must be
applied to all specified limit load conditions. For an aircraft with a ‘lifetime’ of 30000 hrs,
this condition shall be applied for Pj > 3.33 ×10–4 / flight hour. For aircraft having different
‘lifetimes’, the applicable value should be substituted.
4.2.1 These conditions are to be considered for failure modes where the pilot is aware of
the failure. A typical failure of this type is one where a flight control system fails to a
basic state, and pilot-observed limitations are to be used. The limited manoeuvre
envelope must be sufficient to permit recovery to base in all necessary configurations,
including those at landing.
Page 4 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
4.2.2 For the aircraft, in the system-failed state and considering, over the whole flight
envelope, any appropriate reconfiguration, the loads from the following manoeuvre
limitations and gust conditions apply:
(a) A manoeuvre capability not less than half that used for design in the un-failed
state. The limited manoeuvre envelope shall be sufficient to permit recovery to base
in all necessary configurations including those at landing.
(b) Gust capability not less than half that used for design in the un-failed state.
(c) Ground loading conditions specified for the aircraft in the un-failed state.
4.2.3 For static strength substantiation, each part of the structure must also be able to
withstand the loads in the failed state, multiplied by a factor of safety depending on the
probability of being in this failure state. The Factor of Safety is defined in conjunction
with the probability of being in the failed condition j, Qj (where Qj = (Tj)(Pj) and Tj =
average time in hours spent in the failed condition).
4.2.4 For overall military usage, the factor of safety is maintained at 1.5 for
Qj ≥ 1 hour / lifetime. For Qj < 1 hour / lifetime, the factor of safety may be reduced to
1.25 at Qj = 1 × 10–9. Figure 2b shows the factor of safety versus Qj for an aircraft with a
‘lifetime’ of 30000 hours. For aircraft having different ‘lifetimes’, the breakpoints should
be derived in an analogous manner.
Note: If Pj is greater than ‘10 / lifetime’ per flight hour, then a 1.5 factor of safety must be
applied to all specified limit load conditions. For an aircraft with a ‘lifetime’ of 30000
hours this condition shall be applied for Pj > 3.33 × 10–4 / flight hour.
4.3 For residual strength substantiation, the aircraft must be able to withstand 80% of the
ultimate loads defined by the appropriate situation above. For pressurised cabins, these loads
must be combined with the normal operating differential pressure.
4.4 If the loads induced by the failure condition have a significant effect on fatigue or
damage tolerance, then their effects must be taken into account.
4.5 Freedom from aeroelastic instability must be shown up to the speeds specified in Part 1,
Section 4, Clause 4.8
4.6 Regardless of the calculated system reliability, certain failure conditions may require
consideration. Where analysis shows the probability of failure for these conditions to be less
that 10–9, criteria other than those described in this Leaflet may be used for structural
substantiation to show continued safe flight and landing.
5 FAILURE INDICATIONS
5.1 For system failure detection and indication, the following apply:
5.1.1 The system must be checked for failure conditions, not extremely improbable that
degrade the structural capability below the level required by Def Stan 00-970 or
significantly reduce the reliability of the remaining system. As far as reasonably
practicable, the flight crew must be made aware of these failures before flight. Certain
elements of the control system, such as mechanical and hydraulic components, may
use special periodic inspections, and electronic systems may use daily checks, in lieu of
detection and indication systems to achieve the objective of this requirement. These
certification maintenance requirements must be limited to components that are not
Page 5 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
readily detectable by normal detection and indication systems and where service history
shows that inspections will provide an adequate level of safety.
5.1.2 The existence of any failure condition during flight, not extremely improbable, that
could significantly affect the structural capability of the aircraft and for which the
associated reduction in airworthiness can be minimised by suitable flight limitations,
must be signalled to the flightcrew. For example, failure conditions that result in a factor
of safety between the aircraft strength and the loads below 1.25, or flutter margins below
those specified in Part 1, Section 4, Clause 4.8, must be signalled to the crew during
flight.
6.1 If the aircraft is to be dispatched in a known system failure condition that affects
structural performance, or affects the reliability of the remaining system to maintain structural
performance, then the provisions of Clause 3.1.4 must be met for the dispatched condition
and for subsequent failures. Flight limitations and expected operational limitations may be
taken into account in establishing Qj as the combined probability of being in the dispatched
failure condition and the subsequent failure condition for the safety margins in Figures 2a and
2b, and including the flutter margins of Part 1, Section 4, Clause 4.8 These limitations must be
such that the probability of being in this combined failure state and then subsequently
encountering limit load conditions is extremely improbable. No reduction in these safety
margins is allowed if the subsequent system failure rate is greater than ‘10 / lifetime’. For an
aircraft with a lifetime of 30000 hours, this equates to 3.33 × 10–4 per hour.
REFERENCE
1. European Aviation Safety Agency. Certification Specifications for Large Aeroplanes. CS-25,
Amendment 2, 2 October 2006.
Page 6 of 7
DEF STAN 00-970 PART 1/15
SECTION 3
1.5
1.25
1
FS
0.75
0.5
0.25
0
1.E-10 1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00
1.5
1.25
FS
0.75
0.5
0.25
0
1.E-10 1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00
Qj - probability of being in failure condition j
1.5
1.25
1
FS
0.75
0.5
0.25
0
1.E-10 1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00
Qj - probability of being in failure condition j
Page 7 of 7
©Crown Copyright 2015
DStan Helpdesk:
Tel: 44 (0) 141 224 2531/2
Fax: 44 (0) 141 224 2503
Internet e-mail: enquiries@dstan.mod.uk
File Reference
The DStan file reference relating to work on this standard is D/DStan/21/970/1
Contract Requirements
When Defence Standards are incorporated into contracts users are responsible for their correct
application and for complying with contractual and statutory requirements. Compliance with a Defence
Standard does not in itself confer immunity from legal obligations.