You are on page 1of 30

Proc. Indian Acad. Sci. (l~nng. Sei.) Vol. 3) Pt. 1, March 1980, pp. 1-30. @ Printed in India.

Modelling of gusts and wind shear for aircraft assessment


and certification

J G JONES
Flight Systems Department, Royal Aircraft Establishment, Bedford, UK

MS received 2 April 1979

Abstract. Recent research in the UK on the subject of turbulence modelling for air-
craft assessment and certification has emphasised the importance of considering
ordered structures existing in apparently disordered air motions. A consequence of
the existence of such order within disorder is that probability distributions of velocity
gradients and associated velocity increments are often strongly non-Gaussian, even
in so-called continuous turbulence. This non-Gaussian characteristic of turbulence
(which may be related to the phenomenon of 'intermittency') is a source of dissatisfac-
tion with the widely-used power-spectral method for modelling turbulence and is the
main incentive for developing an alternative.
One approach to the problem has been the development of a 'statistical discrete-
gust model of turbulence which takes a discrete ramp gust as a basic element from
which representative gust patterns may be built up and related to probability of occur-
fence. On this basis an alternative approach to aircraft certification has been pro-
posed in which the assessment of a particular design (of aircraft, or gust-alleviation
system) involves a systematic search over a specified family of equiprobable gust
patterns for a 'worst case' which produces maximum aircraft response. Such an
equlprobable family is a generalisation of the traditional concept of a ' design gust.'
An important feature of the method is that the ' worst case ' response may he directly
related to the statistical characteristics of response when the input has a random
character representative of real turbulence. The technique is illustrated with parti-
cular reference to the assessment of autoland systems in severe wind shear and to
limit-load criteria (for which a vortex-tube model is proposed to augment the statistical
family of discrete ramp gusts).

Keywords. Gusts; wind shear; turbulence; aircraft certification; random process;


discrete gust model; Reeves' model; autoland systems; limit loads.

I. Introduction

A n i m p o r t a n t prerequisite for the design, assessment a n d certification o f aircraft


a n d their associated control systems is a quantitative specification o f the environ-
ment in which the aircraft are intended to operate. P r o b a b l y the m o s t critical
influence o f the environment, f r o m this point o f view, lies in the disturbances pro-
duced by the atmospheric motions and irregu~rities that are variously described as
gusts, turbulence a n d wind shear. I n this paper, we review some o n g o i n g research
in the U K aimed at developing i m p r o v e d models o f these atmospheric fluctuations
for use in the aircraft design a n d certification process.
Examination o f records o f atmospheric disturbances taken u n d e r a wide variety
o f conditions shows two conflicting trends: o n e towards ' o r d e r ' a n d the other
t o w a r d s ' disorder '. I n some instances, we find cases where isolated discrete gusts
stand out in a clearly identifiable mRnner. I n others, the turb~tlence rc~ords a p p e a r
2 J G Jones

to have a predominantly irregular or random pattern with little obvious structure.


This ambivalent nature of turbulence had led to widely differing ways of attempting
to describe it. Its tendency to fluctuate in a chaotic, random manner has suggested
that the mathematical theory of continuous random processes be taken as a basis;
this approach is typified by the widely-used power-spectral-density (PSD) approach
to problems concerning aircraft dynamic response in turbulence. On the other hand,
an impression that the more intense fluctuations could somehow be singled out as
individual events was the basis of the earliest approach used by aircraft engineers
and there persists a reluctance (in our view, well-founded) on the part of British air-
craft manufacturers (Payne & Cox 1969; James 1973) to dispense with a ' discrete-
gust' model as a means of representing the more severe disturbances. This has
led to research aimed at clarifying the relationship between the two approaches and
at developing a combined gust and turbulence model which takes account of both
the random character of air motions and of the discrete structures which appear to
be particularly relevant to the larger, potentially critical, disturbances. A result
of this research has been the formulation (Jones 1968, 1973) of a ' statistical discrete-
gust (SDG) modal' of turbulence which takes the discrete ramp gust (figure 1) as
a basic element, or building brick, from which representative gust patterns may be
built up and related to probability of occurrence. On this basis, an alternative
approach to aircraft certification is currently under consideration in which the assess-
ment of a particular design (of aircraft, or control system) involves a systematic
(i.e. deterministic) search over a specified family of 'equiprobable' gust patterns
for a 'worst case' or 'design case' which produces maximum aircraft response.
Such an equiprobable gust family is a generalisation of the traditional concept of a
' design gust '. At the same time, an important aspect of the technique is the rela-
tionship that may be shown to exist between such a design case and the overall pro-
bability distribution for the rate of occurrence of large aircraft response values. In
other words, a deterministic procedure may be taken as the basis for satisfying a
statistical criterion. In the following sections we outline and illustrate some aspects
of this work with particular reference to the assessment of autoland systems and to
limit-load criteria.
The point of view underlying the statistical discrete-gust model of turbulence is
that a turbulent flow field, even when apparently 'continuous ', in fact contains
coherent structures which are more appropriately described in terms of spatial velo-
city distributions than by transforming (as in the PSD approach) to a spectral or
frequency distribution. The coherence, or ' order ', is introduced through the use
of ' discrete gusts ', of simple ramp form (figure 1), as basic elements. The ' dis-
order' is tbeu represented by the use of probability distributions to build up a
turbulence record as an aggregate of these basic elements having random gradient

.... J l
gradrent distance
Figure 1. Discrete'amp gust
Modelling of gusts and windshear 3

distance H and amFlitude W and being randomly distributed in space. This em-
phasis upon the existence of coherent structures in turbulence (which mathematically
corresponds to the existence of strong phase correlations between different frequency
components in a spectral representation of the turbulence field--an important feature
that is generally ignored in the PSD approach) is paralleled in recent studies of the
fluid mechanics of turbulence. For example, Roshko (1976), in taking a ' n e w
look' at the structure of turbulent shear flows, described the change in direction
that has occurred in turbulence research, from the use of the various correlations
and spectral functions which are characteristic of the classical statistical theories,
towards a more deterministic description of turbulence which has been initiated by
the discovery of coherent structures in many turbulent shear flows. An objective of
this recent research (Roshko 1976) is to obtain insight into the physical processes
in turbulent flows, such as entrainment, transport, mixing, noise production and
gustiness through an understanding of these coherent structures, including their
growth and their interactions with each other. The growing interest in the study of
coherent structures in turbulence is also reflected in the range of contributions to a
Colloquium on Coherent Structures in Turbulence held at Southampton in 1974
and reviewed by Davies & Yule (1975). At this Colloquium, flow-visualisation
experiments were reviewed which show that shear flows, including mixing layers
and wakes, boundary layers and jets, comprise coherent, but randomly recurring,
structures. The importance of taking explicit account of the existence of ordered
patterns in the study of turbulence is stressed by both Davies (1975), and Kucbemann
(1971) who expressed the view that progress in the understanding of turbulent fluid
motions is most likely to advance if the overall flow field is related to simple basic
structures containing a high degree of inherent order. Particularly relevant in the
present context is the close affinity between coherent velocity distributions described
by Kuchemann (1971) (figure 2) and the velocity distributions (figure 3) measured
using a research aircraft near storm tops and discussed by Anne Burns (1972).

~-) initiot vetocitydistribution

t r o c e m vortex Is ~ distribution. - "~r"

/I core $

----J /
velocity induced
by the Vortex cote
itletf
Figure 2. Possible velocity distribution across double-branched vortex core in an
array representinga shear layer (K~tchemann1971)
4 J G Jones

V'O
,ol
. 1

I I I t' GO00 10000


metres
15OOO
I
2O000

Figure 3. Vertical gust velocity measured in patch of turbulence near storm tops
(reproduced from Anne Burns 1972)

Whilst the studies of coherent structures referred to above have largely been con-
cemed with relatively large-scale features of the flow field, occurring at the ' energy-
containing' end of the spectral range, parallel studies have been in progress of the
energy distributions arising within the inertial range, i.e. the range of wavelengths
between the large-scale ieatures and the small scale at which energy is dissipated by
the action of viscosity. Particular attention has been paid to the ' spottiness' of
the energy distribution, or 'intermittency '. The simplest model of turbulence
structure in this range of scales invokes the concept of ' serf-similarity '. In this
model, the features at differing scales may be related by a ' stretching' o r ' enlarging'
transformation and organiscd structures satisfying simple scaling laws are envisaged
as occurring in the flow field at all scales~down to the viscous range. Whilst such a
picture provides a good first approximation, if the scales to be compared are not too
widely separated, it is now generally accepted that turbulence structure at differing
wavelengths in fact shows gradual but systematic departures from simple self-simi-
larity. In effect, what has been shown is that turbulence energy is concentrated
into smaller and smaller volumes as the scale is decreased. The measure o f ' energy
compression' may be quantified by the self-similarity dimension (or fractal dimen-
sion) D introduced by Mandelbrot 0977). The range of possible values for D is
2 ~ D ~ 3. The case D -- 3 corresponds to simple self-similarity and does not
reflect the energy compression phenomenon. On the other hand, in the case D =2
the energy at the smallest scales is effectively contained in two-dimensional sheet-
like structures. Current estimates of the appropriate value for turbulence lie in the
range 2"5 ~ D ~ 2.7 and imply a small-scale structure comprising highly convoluted
sheets. An explicit geometrical model for turbulence intermittency, in which D is
an arbitrary parameter, has been proposed by Frisch et al 0978). On this basis,
we have shown in recent work (Jones 1979) that, for the purpose of aircraft response
analysis and prediction, a procedure may be followed that takes the simple self-
similar case, D = 3, as a first approximation and, where necessary, incorporates
effects associated with D < 3 by means of easily-applied correction factors. In
fact, for many applications, the results obtained using the self-similar model, D ---- 3,
are sufficiently accurate, corrections only being significant when the range of
significant wavelengths covers a range of several octaves. Analysis of measured
Modelling of gusts and wind shear 5

turbulence records is currently in progress with the objective of comparing turbulence


models with D in the range 2.5 ~< D ~ 3.
In the remainder of this paper, attention will be mainly concentrated on the simple
self-similar case D : 3. However, in w 5.3 we return to the question of the appro-
priate choice of D and recommend, on the basis of recent work, that consideration
be given to the use of a model with D : 2-5 for the prediction of limit loads in
extreme turbulence whilst a self-similar model with D - - - 3 be retained for such
purposes as ride-quality or aireraft-handling assessment and possibly also structural
fatigue.

2. Comments on non-Gauuian turbulence models

2.1. Background

In the previous section we referred to limitations of the power-spectral-density (PSI))


approach as regards its ability to represent the more intense turbulence fluctuations and
their effects on aircraft response. The standard power-spectral model is concerned
with turbulence energy distributions (with respect to wavelength) in patches of tur-
bulence that are sufficiently long for the aircraft response to be evaluated on the
assumption of a state of statistical equilibrium between excitation and response.
Even when tlfis condition is satisfied, whilst the mean-square value of aircraft res-
ponse can be calculated, in order to predict the magnitudes of the large peaks in
response it is necessary to know in addition the probability distribution of peak
amplitudes. A weakness of the power-spectrum method is that no satisfactory
technique for obtaining this amplitude probability distribution exists. The usual
assumption is that each patch of turbulence may be represented by a stationary
Gaussian process; on this basis, the distribution of response peaks can then be cal-
culated.
However, measured peaks in the response of aircraft to turbulence often closely
follow an exponential distribution, containing significantly more large peaks than in
the response to a Gaussian process. An argument widely used by advocates of
the power-spectral approach is that this non-Gaussian behaviour may be due to the
patchiness of turbulence and the resulting averaging of a sequence of possibly Gaus-
sian density functions into a composite one which is approximately exponential~
For example, Dutton (1970) discussed this ' local-Ganssian' hypothesis and concluded
that the hypothesis that turbulence is patchy and locally Gaussian not only explains
observed histograms but is also consistent with the tendency of measured exceedance
curves to be exponential rather than Gaussian. However, this argument has been
effectively refuted by Reeves et al 0976) who show theoretically that if turbulence
could be described as a collection of Ganssian patches then the normalised probabi-
lity densities of gust velocities and gust velocity increments (see below) would neces-
sarily be equal. As discussed by Reeves et al 0976) there is abundant evidence
that this is not the case (including data presented by Dutton and Thomson (1968)
and by Chen (1972)) and that therefore the representation of atmospheric turbulence
by means of sequences of Gaussian patches is questionable. This, of course, does
not rule out the Power-spectral theory as a practical tool--the predictions of a theory
are not necessarily sensitively dependent on the real~m of all its basic assumptions.
6 YGYones
It does, however, imply that there is significant room for the investigation of alter-
natives.
In the previous paragraph we referred to the probability density of gust velocity
increments. For a turbulence record v(s), defined as a function of spatial position
s, we define the velocity increment function, over distance (or lag) S, to be

Ys(s) - v0- s). (1)


The properties of the velocity increment function Ys, regarded as a falter acting on
v(s), are discussed in detail in Jones (1973), where it is related to the concept of an
'averaged gradient' over distance S. The importance, for aeronautical applica-
tions, of probability distributions which quantify the change in turbulence velocity
to be expected over a prescribed distance was emphasised by Jones (1967). This
imp3rtance is ' because such quantities as aircraft loads, or airspeed changes when
the aircraft is controlled by pilot or autopilot, depend not only upon the magnitude
of the gust velocities but also to a large extent upon their rate of change. This is a
consequence of aircraft response decreasing the effects of the low frequency com-
ponents of turbulence: what usually matters is the probable change in turbulence
velocity in an interval of the order of the aircraft response time' (Jones 1967). It
was already suspected at the time this was written that distributions of velocity incre-
ments would tend to be non-Gaussian even in situations where the first-order ampli-
tude distribution of the turbulence may itself be closely Gaussian, on account of the
tendency in turbulent flows for a small number of strong gradients to form rather
than a uniform distribution of small gradients. Subsequent data analysis, both at
the Royal Aircraft Establishanent (Burnham 1968) and elsewhere has amply con-
firmed this result. What has been shown is that in real turbulence the probability
of meeting either very large or very small increments is substantially greater than in
a Gaussian distribution, the probability of meeting average values being smaller, in
compensation. Such a probability distribution is illustrated sehematieally (in ex-
aggerated form) in figure 4. In the language of probability theory, such a distribu-
tion has a larger-than-Gaussian fourth-moment, or flatness-factor, or kurtosis.
Physically, a sequence of velocity increments with a distribution of this form shows
high 'contrast' (to use a photographic analogy), very large increments being
separated by very small increments which form regions of ' low activity '. Such a

v
- G a u s s i o n process
..... intermittent proce~

t
f

_J Or m a t d i s t r i b u t i o n

: ntt~ decoy

m._

Figure 4. Probability i1e~sityof velocityincr9169 (schematic)


Modelling of gusts and wind shear 7

characteristic is a feature of ' intermittency '. For aeronautical applications, what


matters is that the large increments, and associated aircraft response amplitudes, have
significantly higher probability than would be predicted using Gaussian-process
theory. The close relationship that exists between large values of the increment
function and large amplitudes of aircraft response is illustrated in figure 5.
In view of the above limitations of employing Gaussian processes to model tur-
bulence, alternative approaches which take explicit account of non-Ganssian charac-
teristics have been investigated over the past few years. This work is reviewed in
the following section.

2.2. Introductory comments on non.Gaussian models of turbulence

Once it has been accepted that turbulence is significantly non-Gaussian, the problem
arises as to how this non-Gaussian character may be represented and related to the
statistical characteristics of aircraft response. In this section we suggest that there
are two distinct ways of attempting to model turbulence for this purpose. The basis
of this dichotomy may be appreciated by noting that whilst the expressions ' gust'
and ' gust pattern' are commonly employed in the context of atmospheric turbu-
lence there is no commonly accepted precise meaning implied. Indeed, the extensive
use of stationary random process theory to describe turbulence samples has tended
to emphasise the chaotic aspects of turbulence and thus to disregard the concept of
a ' gust' in the sense of an organised discrete structure. As a result there have
evolved two significantly different approaches to the problem, one based on mathe-
matical methods which are in principle capable of treating a wide class of non-
Gaussian random process problems and the other, of a more heuristic nature, based
on the intuitive concept of a ' gust' and introducing hypotheses of a physical nature
concerning the particular non-Gaussian structure of turbulence and the manner in
which this affects aircraft response.
The problem is closely analogous to that of theoretically representing turbulence
for the purpose of clarifying the fluid-dynamic processes involved. In this case the
two approaches to the problem have been well described by Saffmau (1968). The
traditional approach is to employ a general wave number (Fourier mode) representa-
tion of the quantities (pressure, velocity, etc) involved and to transform the equations

a computer simulation of aircraft na'rnot acceleration


(heaving ptus pitching damping ratio 0.6)

b smoothed increment function


u ( s)=f• (s - s') vs (s') ds'
where Ys (ss)'=v ( s l ) - v (s~S) and X ( s - s ' ) is weighting
function
Figure 5. Relationship between smoothed increment function and aircraft response.
Continuous turbulenceinput o(s),lag S equal to tuned gust length of aircraft response.
8 J G Jones

of motion to this space. The non-linear terms then appear as interactions between
components of differing wave number. Whilst from a mathematical point of view
this method looks initially attractive, there are severe practical problems concerning
the truncation of the series expansions involved. For instance, one method, deve-
loped by Meecham & Siegel (1964), makes use of a Wiener-Hermite stochastic
expansion. In principle, this representation is capable of application to general non-
linear interactions and general non-Gaussian processes. However, in practice, there
results a series expansion in which the first few terms are dominant for nearly-
Gaussian processes, such as those resulting from weak non-linear interactions, but
which can lead to severe analytical complexity when applied to the strongly non-
linear turbulence problem, in which strong phase correlations develop between
Fourier modes and the velocity fields correspondingly become strongly spatially
structured.
Saffman (1968) approached this problem by assuming from the outset that turbu-
lence comprises a random array of vortex sheets and tubes. Thus, in this approach,
instead of performing a series expansion in which the first term represents ' chaos'
(Gaussian process) and successive terms slowly introduce' structure ', the alternative
is adopted of postulating the existence of certain well-defined structures at the outset
and representing turbulence as a random collection of these structures. In this
manner empirical features are built into the basic elements of the theory and, at the
expense of mathematical generality, the concepts studied become more directly
related to the physical phenomena observed experimentally.
Returning to the problem of predicting the statistical characteristics of aircraft
response, general mathematical methods are available analogous to the wave number
interaction approach to turbulence dynamics. These techniques make use of series
expansions, the first terms of which provide a complete specification in the case of a
Gaussian process, the non-Ganssian characteristics being introduced in the higher
order terms. Such methods have had practical applications outside the field of
aeronautical engineering, for instance in radio engineering in studying the transmis-
sion of noise through the final amplifier of a receiver following a detector. One
such method, due to the Russians (Kuznetsov et al 1965), makes use of series
expansions by which an arbitrary stoChastic process can be characterised and involves
rules whereby the statistical characteristics of the response of both linear and
non-linear systems to such an input may in principle be obtained. In particular, it
is shown how the problem of evaluating the effects of non-Gaussian flucfuations on
linear control systems can be treated by the approximate 'quasi-moment' method,
in which the one- and two-dimensional probability density functions of the input
fluctuations are expanded as series in 'Chebyshev-Hermite' polynominals. By
means of a recurrence method for calculating the series coefficients, approximate
expressions can then be written for the probability density functions of the output
fluctuations.
It would appear that the above theoretical approach to the problem has a high
degree of generality and deserves a more detailed assessment of its practical
capabilities than has been made to date. On the other hand, it has a high degree
of analytical complexity, and whilst it appears to be appropriate in the case of
nearly-Gaussian processes, when the series involved can be truncated without serious
error after only a few terms, its practical value in the context of aircraft response to
turbulence, which is strongly non-Oaussian, is in our view limited.
Modelling of gusts and wind shear 9

The use of Hermite polynominals, mentioned above, recurs in the so-called'Wiener.


Hermite functional technique' for the representation of turbulence. A description of
this method has been presented by Dutton (1970) who also gives an historical account
of methods that have been employed to treat aircraft response to turbulence. As
Dutton points out, while in principle this technique has sufficient power to treat
general non-linear systems and non-Gaussian inputs, it only takes a simple form if the
problem is 'nearly' linear and Gaussian. In the general case the cost in analytical
complication appears to be excessive. Thus whilst the idea (Dutton 1970) that 'a
Wiener-Hermite expansion carrying the characteristics of turbulence will provide an
input with the true properties of the forcing functions to be found in the atmosphere"
appears superficially to be attractive we believe an alternative approach to the re-
presentation of the strongly non-Ganssian characteristics of atmospheric turbulence
to be necessary for aeronautical applications.
Two alternative practical approaches to this problem have evolved over the period
from 1968 to the present time. One is the non-Gaussian model of continuous atmos-
pheric turbulence initially developed by Reeves et al (1976), and the other is the
statistical-discrete-gust theory (Jones 1968, 1973) which forms the basis of the work
to be outlined subsequently in this paper. These two approaches are introduced
in the two following sections.

2.3 Reeves' model

Reeves' original report (1969) developed a non=Ganssian turbulence simulation


intended to be used in conjunction with piloted flight simulators. The multiplication
of independent Gaussian random signals was the basic approach used to generate
each of three orthogonai gust components. The amplitude probability density of the
simulated gust components was shown to take a non-Gaussian form characterised by
a modified Bessel function. By linear filtering, the desired forms of power spectral
densities could also be obtained. In a later report (Reeves et al 1974) an experiment
was described in which a STOL airplane was programmed into a 6-degree-of-freedom
flight simulator and pilots' reactions to various forms of the model were described.
Subsequently (Reeves et al 1976) the method was developed to calculate aircraft
response statistics analytically, evaluation of response probability d~stributions and
level-crossing frequencies requiring calculation of the eigenfunctions and eigenvalues
of a certain kernel function. The statistical properties of the proposed model have
been compared (Reeves et al 1976) with experimentally measured properties of low
altitude atmospheric turbulence. The results have shown that the power spectral
densities, probability distributions, and level-crossing frequencies observed in the
atmosphere can be modelled quite well. However, it has also been shown (Reeves
et al 1976) that this model, like the Gaussian and Gaussian patch models, does not
reproduce the velocity increment distn~butions of turbulence. Suggestions for
further work were made to correct this deficiency (Reeves et al 1976).
The Reeves model has also been taken as the basis for further work on the modelling
of turbulence for simulation purposes at Delft University of Technology, the Nether-
lands. For example, Gerlach et al (1973) describe a method, based on the Reeves
technique, to generate simulated non-Gaussian turbulence having a prescr~q~edpower
spectral density and amplitude probability density function as well as a chosen value
of a so-called 'patchiness parameter'. Gerlach et al define 'patchiness' in terms of the
10 JGJones

non-Gaussian distribution (characterised by the fourth-order moment) of the turbu-


lence velocity amplitude; this should be contrasted with the concept of 'interndttency',
described earlier, which is based upon the non-Gaussian characteristics of velocity
increments or, equivalently, averaged velocity gradients. In our view, further work
remains to be done to clarify l~ow non-Gaussian characteristics of turbulence velocity
are related to the intuitive concept of 'patchiness'. This subject is important not
only in the context of ground-based simulation but also for investigating effects of
turbulence on aircraft loads, when patch lengths have importance particularly for
predicting response of lightly-damped modes. A more recent description of the work
at Delft is presented in van de Moesdijk 0975) where it is shown that simulated time
histories of turbulence velocities possessing prescribed patchiness characteristics (as
defined above) can be generated in real time with simple programs on any general
purpose digital computer. More recently, van de Moesdijk 0978) has described an
alternative method in which a Gaussian signal is multiplied by its own Hilbert
transform. This development shows promise as a means of generating a random
function with non-Gaussian increments and will no doubt be subject to further
validation.

2.4 Introductory comments on statistical discrete-gust model

It is of some interest that, whereas in the case of the Reeves model (w 2.3) the method
was first developed as a technique for producing simulated time-histories and its
application as a tool for calculating the statistical characteristics of aircraft response
was developed subsequently, the statistical discrete-gust (SDO) model was first
proposed as an analytical tool for predicting the rate of occurrence of large aircraft=
response amplitudes and only later used as the basis of a method for generating
real-time 'synthetic turbulence' histories for simulation purposes. Details of a
technique in which the statistical discrete-gust model is used to produce, in real time,
a continuous turbulence record for aircraft handling qualities studies in a piloted
ground-based simulator have been presented in Tomlinson (1975). For this
application, interest is mainly centred on 'average' conditions, which correspond
approximately to 'moderate' turbulence. In the present paper, we shall on the other
hand be principally concerned with applications in which the relatively rare but
potentially catastrophic gust (i.e. 'severe' or 'extreme' turbulence) is of primary
interest ( ~ 4 and 5).
At the outset, we emphasise that a principal objective of the statistical discrete-
gust theory has been to represent correctly both the power-spectral properties of
turbulence and the amplitude probability distributions of turbulence velocity
increments or, equivalently, averaged gradients. As a basis for introducing
prescribed statistical properties of the velocity increment function into the turbulence
model, the discrete-ramp gust illustrated in figure 1, which essentially introduces an
increment W over a distance H, proves to be an appropriate basic element.
The reason for the explicit emphasis on velocity increments is, as already stated,
that aircraft response amplitudes can be shown to be closely related to turbulence
velocity increments, where the increment is taken over a distance equal to the 'tuned
gust length' of the aircraft response (a concept to be quantified in the following
section). In the case of a well-damped system, the close relationship between these two
functions is well illustrated in figure 5. In the case of more lightly-damped modes
Modelling of gusts and wind shear 11

= H2

Figure & Gust pair pattern

there is not such a simple one-to-one relationship between large peaks in aircraft
response and individual large peaks in the turbulence increment function, as
combined effecst of sequential peaks in the increment function become important.
Nevertheless, by taking into account a relationship between the probability of
meeting a given gust pattern and the number of ramp components comprising that
pattern (see figure 6), information concerning the increment function over an
appropriate tuned distance proves to be sufficient as a basis for predicting the rate of
occurrence of large peaks of specified amplitude in aircraft response.
The discrete-gust concept has been employed in one form or another in aeronautical
engineering for many years. However, in contrast to the viewpoint taken in the
present work, 'discrete-gusts' and 'continuous turbulence' have generally been regard-
ed as quite distinct forms of atmospheric disturbance, corresponding to differing
environmental conditions. Indeed, this distinction is to be found in current formu-
lations of aircraft certification requirements, the implication being that two quite
distinct phenomena are being catered for. Now, whilst it is quite true that gust
disturbances do sometimes take a form which can only be construed as a discrete
event, the converse implication that 'continuous turbulence' is devoid of discrete events
simply reflects an over-emphasis on its irregular or chaotic aspect. In fact, as described
in w l, whilst so-called continuous turbulence may contain little obvious structure
on superficial examination, detailed statistical analysis shows it to differ in significant
respects from the 'structureless' Gaussian process. Spatially-averaged velocity
gradients, and related two-point velocity differences, have probability distributions
which show a high degree of kurtosis which, as we have discussed elsewhere, implies
a preponderance of both very large and very small values, at the expense of average
values, and which may be interpreted as 'intermittency' (figure 4). Thus the mathe-
matical concept of a non-Gaussian distribution of increments may be related to the
intuitive concept of 'structure', and the discrete-gust concept can be applied not only
to manifestly isolated events but also in the context of'continuous turbulence'. More-
over, as we shall indicate, statistical properties of turbulence, modelled in this manner,
can be related theoretically to the associated statistical properties of aircraft response.
Having chosen the discrete-ramp (figure l) as the basic element, the next requirement
is to express the probability of occurrence of this element in terms of H and W.
The emphasis in the method is on large amplitude gusts and the theory is essentially
developed asymptotically for large W. As described in Jones (1973), for values of
H~L, where L is the scale length, a sample of turbulence is modelled by a statistical
family, or 'spectrum' of discrete ramp gusts as follows. The number of discrete gusts
per unit distance in the length (gradient distance) range (H, H~-dH) and having gust
intensity greater than W is taken to be N~,wdH where
12 Y O Jones

W
N& w = ~ i e x p ( - - 1.i 5/~/tf/s). (2)

This equation assumes simple self-similarity (D-----3, see w 1) and incorporates the
following features:

(i) The gust amplitudo distributionfor fixed H is exponential (a non-Gaussian


feature).
(ii) There are more short gusts per unit distance, in length range dH, than long
gusts (the H -t factor).
(iii) a is a dimensionless constant defining the overall rate of occurrence (or
packing density) of discrete gusts.
(iv) ~ H tn, where ~ is a constant, is an overall measure of the amplitude of
discrete gusts of length H. T h e / p / s factor is related to the '--5/3' exponent
in the 'inertial sub-range' form of the power spectrum. The numerical
constant 1.15 is introduced to simplify subsequent formulae.

As regards assumption (i), analysis of measured data has shown that turbulence-
velocity increments tend to follow a non-Gaussian distribution. The exponential
model not only gives a good fit to measured data (e.g. figures 13 and 14) but has the
advantage of traditional use in aircraft applications and moreover leads (as outlined
in w 3) to significant analytical simplifications.
Equation (2) applies to the aggregate of discrete gusts of either sign (e.g. up-gusts
and down-gnsts). If necessary, these could be represented by two separate families,
each having its own (not necessarily equal) values of a and 8.
Experience has shown that, asymptotically for large values of W, equation (2) is
equally applicable to an entire aircraft life history, a segment of a mission, or to a
relatively short patch of 'continuous' turbulence, although the appropriate values of
a and//differ from case to case. In the case of a measured sample of continuous
turbulence, Jones (1971, 1973) describes how the values of a and ~ that characterise
that sample may be obtained by passing the data through a set of bandpass filters, of
relatively wide bandwidth, each of which is derived by adding a smoothing or
weighting function (see figure 5) to the velocity increment function Ys, equation (1).
The overall s~t of filters is obtained by taking a sequence of values of the increment
$, where S is generally of the form 2u x sampling interval (of the dlgitised data).
To complete the statistical formulation of a turbulence record as an array of ramp
gusts subject to equation (2) we need to make some assumption concerning the
probability of meeting 'clusters' of ramp gusts of large amplitude. In the technique
described in Tomlinson (1975), in which equation (2) is taken as the basis of a digital
software program for generating 'synthetic' turbulence for studying aircraft hand-
ling and ride qualities in a ground-based simulator, it is assumed that the individual
ramp gusts occur at random times and independently of one another, as in the
generation of shot noise (Middleton 1960). Whilst this simplifying hypothesis
appears to be adequate at the 'moderate' levels of turbulence relevant to the above
application it is not valid at the high levels of gust intensity which define critical
conditions from the point of view of autoland certification or limit loads. For such
conditions ( ~ 4.1 and 5.1) the occurrence of one very large gust implies the existence
of a local meteorological situation which increases the probability of occurrence of
Modelling of gusts and wind shear !3

a second or further large gusts. For these applications, the theory thus has to be
formulated in a manner that takes account of the probability of occurrence of isolated
gusts, joint probability of occurrence of gust pairs, and so on, where these factors
may be determined empirically on the basis of real turbulence data. This feature is
discussed further subsequently. In the following section, the process by which this
turbulence model may be used to predict analytically the statistical characteristics of
aircraft response is outlined.

3. Use of statistical discrete-gust theory to predict aircraft response in turbulence

We now indicate how the discrete-gust model defined by equation (2) may be applied
to calculate the statistical characteristics of aircraft response. We will review the
theory in two steps. First (w 3.1) we consider the case where it is assumed that the
effects of individual ramp gusts upon peak values of aircraft response may be
considered independently. This hypothesis will give good results in practice if the
'overswing' in the transient response to a single ramp gust (figure 7) is small relative
to the primary peak. In addition, this analysis may equally be applied to predict
the response of a broad-band digital filter, and thus provides the basis of a quanti-
tative method for fitting the statistical discrete-gust model to measured samples of
turbulence (w3.3). In w3.2 we describe a generalisation of the theory to cater for the
important case in which there is significant overswing (figure 7), and the effects of
simple gust patterns must be considered. For many applications it is sufficient to
consider only gust pairs.

3.1 The case of discrete gusts with independent effects

We now consider the evaluation of aircraft (or numerical falter) response on the basis
of the discrete gust model described by equation (2). We consider the response y(t)
associated with a linear transfer function (or frequency-response function) 7', to the
turbulence velocity component v(t) (making the usual hypothesis of 'frozen' turbu-
lence and relating time t to distance s through the equation s = Vt, where, in the
case of aircraft response, V is the mean airspeed).
Although the family of discrete gusts, equation (2), comprises gusts covering a wide
range of lengths, it is a feature of statistical discrete-gust theory that, for any parti-
cular aircraft response quantity y(t), the large peaks in response can be related (Jones
1968) to a 'critical case'. In situations where the dominant response modes are
moderately well-damped this critical case consists in general of a single ramp gust
with definite length which tunes to the transient aircraft response. This follows
from the fact that the significant contribution to large response peaks, expressed as

macroscaLe Peak over owing - overmw0ng

Figure 7. Typicaltransient ~ to diNempnN~_JK~tUultt~ti~ovmwings.


14 Y G Jones

an integral over all gust lengths, is dominated by the comribution from only a limited
range of gusts. Explicitly, the rate-of-occurrence of large response peaks may be
expressed as tho product of the contribution from the 'critical case' and a 'sensiti-
vity' factor, which allows for variations in gust length about the 'tuned length'.
Mathematically, the above result stems from an application (Jones 1968) of the
Laplace asymptotic approximation to the integral, and depends essentially upon the
exponential function in equation (2).
The result may be expressed explicitly as follows. A response function 7(H) is
defined (figure 8) to be the magnitude of the peak response to an isolated ramp gust
(figure l) of length (gradient distance) H and intensity W ~ H l/a, assurfling zero
(quiescent) initial conditions at gust onset. Denoting by Ny the rate of occurrence,
per unit distance, of response peaks greater than magnitude y it can be shown (Jones
1968) that, for given values of a and fl, ~HNy is a universal function ofy/7(H), where
(figure 8) H is the 'tuned gust length' at which 7(H) attains its maximum and thus
7(H) is the 'worst case' response with respect to the specified gust family. ;~ is the
'gust length sensitivity', a measure of the breadth of the peak in 7(H). Application
of the Laplace approximation leads to an expression for Adependent upon 7"(ff)s
a measure of the sharpness of the peak at H ~ H . For practical purposes, it is
convenient to replace the term in 7"(H) by an approximation based on fitting a
parabola to the function 7(H) over the range from/~/2 to H (the result is relatively
insensitive to the particular choice of range). Then ~ may be evaluated using the
equation

1 - (3)
-log,2. /
On the basis of the assumed exponential form, equation (2), for the rate of occur-
rence Na, w of discrete gusts, application of the Laplace approximation leads to an
approximate proportional expression

N, ~ y-;/' exp {-yl[l.15

distonce s distance s

H : t u n e d gust tength

I
I
t
H 'tOgH
.Fk~re- 8. --Disc~e.~st respons~ fun~o~
Modelling of gusts and wind shear 15

The final simplification is to note that over an extensive range of y, adequate for
practical applications, N, can be well approximated by a simple exponential such that

N, : - = exp ..... . (4)


AH fl 7

For equation (4) to apply, it is necessary that the dominant response modes should
be sufficiently well-damped for the transient response to a tuned ramp gust to
have relatively small overswing (figure 7). If this condition is not satisfied it is
necessary to go on to examine the response to simple gust patterns, particularly those
comprising a sequential pair of ramp gusts of opposite sign, spaced so that the over-
swing in response to the first gust coincides with the primary peak associated with the
second gust. Such a situation is considered in w 3.2.

3.2 Response to simple gust patterns

Provided that the overswings (figure 7) in the response to an isolated tuned ramp gust
are relatively small, the assumption that there exists a one-to-one relationship between
individual ramp gusts and large peaks in response is adequate in practice. The
effects of the 'background' level of turbulence fluctuation or of previous discrete gusts
cannot produce significant peak amplification. If, however, there exist substantial
overswings, the possibility of a resonant build-up of response over a sequence of gusts
must be considered. We concentrate in the following on the important case where
only the first overswing is significantly large and when, in consequence, the critical
pattern consists simply of a pair of gusts. In related work, aimed at determining
the practical limits of the statistical discrete-gust approach, more complex patterns
comprising up to eight component ramp gusts have been studied (Card 1978) in the
context of very lightly-damped modes. This extension is employed i n w 5.2 (see
equation (7)).
In the case where the response to an isolated ramp gust consists of a single significant
response peak we have shown in the previous section how a tuned gust length H may
be determined as the value of H at which the response function y(H) attains its
maximum. In an exactly analogous manner, we can determine a 'tuned overswing'
gust of length H 0 in terms of the maximum value e r a response function y0(H) defined,
to be the magnitude of the overswing in response to an isolated ramp gust of length
H and intensity W--H x/8. By combining these two cases it can be seen that, if the
spacing between the gusts is appropriately chosen, a resonant condition in response
to a pair of gusts generally exists in which the maximum response is equal to the sum
~9.-----y(H)+ y0(Ho) of the amplitudes associated with H and H0 respectively. A
simple gust pattern of this form is illustrated in figure 6, consisting of two differing
ramp gusts separated by spacing distance H~. The wall-known "one-minus-cosine'

F~ 9. 'Onc-minus-cos~' gust l~ttern


i6 J G Jones

gust pattern (figure 9) is a particular instance of the more general gust pattern shown
in figure 6.
For gust patterns of this form it is possible to derive an approximate equation,
similar to equation (4), describing the rate of occurrence of large response peaks
associated with gust pairs and produced by the 'resonance' mechanism described.
Although, in contrast to the single isolated ramp gust, the gust pattern (figure 6)
is multi-dimensional, the argument leading to equation (4) may be extended to apply
to this more general situation. We recall that equation (4) resulted from an integral
over isolated ramp gusts of all lengths but was reduced, by means of an asymptotic
approximation, to a relatively simple form comprising an exponential factor deter-
mined by the response to a particular critical gust and a sensitivity factor which
allowed for variations about this critical case. Following the same approach,
we may simplify the corresponding general expression associated with gust pairs,
consisting of a multi-dimensional integral over all patterns of the form illustrated in
figure 6, to a corresponding simple form comprising an exponential factor determined
by the response to a 'critical pattern' and a sensitivity factor which allows for variations
about this critical pattern.
For the response to a single ramp gust, the critical case is found by searching over
the family of discrete gusts iUustrated in figure l0 (talcing unit intensity parameter U0)
for the critical, or tuned, gust which produces the maximum peak response value.
This peak response value is the term 7(H) appearing in equation (4). Following
this approach, an analogous 'critical pattern' may be found by searching over a
set of gust pairs (figure 6). Then, denoting by {N,)2 the rate of occurrence per unit
distance of resonant response peaks greater than magnitude y, produced by the
mechanism described, we obtain:

{Nv},= as exp --~-2-~I" (5)


~a:3f~

In equation (5), ~s is the peak response to the critical gust pair, i.e. the 'worst case'
response, ~-'~ is a measure of the spatial extent of the critical gust pair, and a s, ~ are
respectively parameters dependent on the rate of occurrence and amplitude probability
distribution of gust pairs. ~ is a sensitivity factor, which allows for variations

gust l UoH1/3(H~<L)
intensitygradient
disto~ce . . . ~

w=Uo
Figure 10. Family of equiprobable ramp gusts defined for H < L by intensity para.
n~ter U.
Modelling of gusts and wind shear 17

about the critical pattern, analogous to the gustqength sensitivity (equation (3)) of
the isolated gust ease.

3.3 Analysis of measured turbulence samplesfrom discrete-gustpoint of view

In addition to studies of spectra and of probability distributions of turbulence velocity


and velocity increments (discussed earlier), the results of which have been to confirm
the non-Gaussian character of measured turbulence velocity increments (figures 11 and
12) extensive analysis of turbulence samples has been performed at the Royal Aircraft
Establishment using a wide-band filter technique (Jones 1971, 1973), based on the
smoothed increment function (figure 5). This technique combines aspects of time-
plane analysis, associated with explicit consideration of turbulence-velocity incre-
ments, with aspects of frequency-plane analysis, associated with the interpretation of
the smoothed-increment function as a bandpass filter. Data has been analysed from
a wide range of sources, including low-altitude data from tower-based anemometers
(figure 14) and thunderstorm turbulence obtained by means of specially instrumented
research aircraft (figure 13). To assist in the physical understanding of turbulence
structure over the widest possible range of wavelengths, measurements of fine-scale
turbulence from a hot-wire anemometer have also been analysed, at waveIengths
approaching the viscous range (wavelengths of the order of 1 cm). Quantitative
results have been obtained by fitting the theoretical expression, equation (4), to

0.8 ~0"8 I' "

~0.6
x
~0.4
O
~ O.2
A
-5 -3 -1 1 3 5 -5 -3 -I I 3 5
turbulence velocity turbulence velocity increments
(norrnotisedto unit varionce) normaLisedtoun/t variance)
99.99
99.90
99.90 / ~
norrnot .~ /
distributionS/ 99.0
99.0

90
/
/ .Co_ 90

bU 50 ! ~
C 5o

///
"0 / ~'~ increment
10 10 ~/~ o 185ft(56.39m)
.-: ~ x 3?lft (,3.O8m)
It 1.0 1.0 ~=e~o~) r ~ 742ft(226.16m)
o 0.1 ~x7
:~ 9 1483ft (452.01m)
G 0.1
0.01 0.01 // o 296eft (904.34m)
i I L l I I l I I
-5 -3 -1 1 3 -5 -3 -1 1 3 5
turbulence velocity turbulence' velocity increments
Figure 11. Probability densities and distributions for vertical turbulence component.
Thunderstorm data measured at 27,000 ft (8,229.6 m)
Pro. C.-2
18 J G Jones

0.6 .', ' 0.8

'~ 0.6 ~ O.G


C
Q

~. 0.4

o 0.
.13
0.2
O o

-3 -1 1 3 -5 -3 -1 1 3 5
t u r b u l e n c e velocity turbulence velocity increments
(norrr~t~sed to unit ~3r~once) ( normat~sed to u~it vorJonce )
99.99 99.99
99.90 99.90
90.00 90.0C
:o
~ 60.0~
,O
/ nor mat
,, d~stribclt~ 50.00 nt
i' ~( o 64ft(t950m)
/ ~ 12eft (39.00m)
--?_ 10.00 10.0C
~ 256ft (78.00m)
o ~.00 1.OO ~wj~' 9 511ft (155 75m)
r
0
0.10 '~ 0,10 ~ / ~ r o 1022ft(31150m)
0,01 GO1
i ~ * 9 l J 2 I --t
-S -3 -1 3 S -5 -3 -I ~ 3 5
t u r b u l e n c e veLooty t u r b u l e n c e velocity ,ncreme~tll.
Figure 12. Probability densities and distributions for longitudinal turbulence com-
ponent. Tower-based anemometer data measured at 600 ft (182-88 m)

0~ H- increment

[
g l L 9
increment
371ft (113.08m)
_ l t V 7&2ft(226.16m)

-4 -yivi~f '

F i g ~ e 13. Peak distribution for s m ~ t h e d i n ~ e m e n t f ~ c t i o n of vertical-turbu-


lence COml~nent. Thunderstorm data measured at 27.000 ft (8,229.6 m)
Modelling of gusts and wind shear 19

R =incf'e rr1~nt
increment
9 128ft (39.00m)
A 256ft (78.00m)
-1 f t (155.75 m)

-3

-4
Y/y (R)
Fignlre 14. Peak distribution for smoothed increment function of longitudinal tur-
bulenco component. Tower-based anemometer data m o a s u r ~ at 600 ft (182.88 m)

measured rates of occurrence of peaks in the response to the set of wide-band filters
(figures 13 and 14). The filter response quantities H, )~and ~(H) have been established
by using isolated discrete gusts as calibrating inputs. In this way the basic
hypotheses underlying equation (2) (self-similarity, the law of exponential decay, and
the H 1Is law) have been confirmed and the turbulence parameters (a,/3) measured.
A detailed survey of these results is in preparation.
At long wavelengths, of the order of the scale length and beyond, it has been found
in the case of low altitude turbulence that the non-Gaussian characteristics of turbu-
lence become less pronounced. This is t o b e expected as the 'exponential' peaks in
velocity increments are interpreted as being due to 'organised structures' in the turbu-
lence velocity field, and the scale length is physically a measure of the spatial extent oi"
the largest of these structures. In the case of high-altitude turbulence, there is an
unfortunate lack of reliable long wavelength data (it is difficult to ensure that the
effects of aircraft motion are adequately removed from the turbulence measurements
at these wavelengths). However, what evidence there is suggests that organised shears
may exist in the upper atmosphere up to extremely long wavelengths, with associated
non-Gaussian characteristics of velocity increments. Measurements of these long
wavelength characteristics of fluctuating air motions remain an important area for
further research.
20 J G Jones

4. Gust and wind shear model for autoland 'systems

4.1 Nature of large disturbances

An important area where gusts and wind shear can cause aircraft operational problems
concerns take-off and landing. There is a need to be able to assess the effects of these
disturbances both for manual landing and for the assessment of the safety of automatic
landing systems.
From the operational point of view, it is required to relate the severity of distur-
bances to be met on the landing approach to the prevailing meteorological conditions,
to allow both the prediction of the severity of disturbances for piloted aircraft and the
placing of sensible limits on the conditions in which automatic landing systems may
be used.
There is considerable current interest, stimulated by recent fatal accidents, in the
problem of 'wind shear'. Wind shears come in differing forms depending on the
meteorological conditions. Shears characterised by strong vertical gradients of
horizontal wind speed can occur under conditions of high atmospheric stability, asso-
ciated with 'inversions' involving large temperature stratification. Under these
conditions time-varying fluctuations tend to be highly damped and the 'turbulence'
level, or 'gustiness', is generally low (although significant wave effects can sometimes
occur). Art assessment of the influence on aircraft of these conditions may generally
be made simply by choosing a 'mean wind' profile with appropriate variation with
height. The operational significance of this type of wind shear is perhaps greatest
on take-off, where conditions can occur in which the surface wind is calm but at
heights as low as 50 ft (15.24 m) the wind may be 10 kn or more. Since, under
inversion conditions the temperature increases with height (10~ increase in 100 ft
(30.48 m) was quoted in a recem accident report), the influence of the shear, possibly
causing a sharp reduction in airspeed, can occur simultaneously with a loss of power.
A quite different class of wind shears is r by strong horizontal gradients
of horizontal wind speed (although possibly in conjunction with vertical gradients).
In this ease, the distinction between 'shear' and "gustiness' is largely a matter of
definition, and in the following section we shall indicatehow the 'discrete gust' model
of turbulence is being developed to eater for this class of phenomena. The most
significant disturbances of this type tend to be relatively isolated, and are often
referred to as 'squalls' or 'gust fronts'. They are associated with -thunderstorms or
other strong sources of convection. From the operational point of view, it is significant
that large wind variations associated with convection tend to occur when the wind had
previously been relatively light (Br~raham & Colmer 1969) and their presence on a
landing approach flight path can go unsuspected. In particular, their effects are not
likely to be reduced significantly by imposing wind limits simply based on mean
windspeeds as conventionally used by controllers at airports. As a result of its
investigation of an accident to an Eastern Airlines Boeing 727 at New York's John
F Kennedy International Airport in 1975, the US National Transportation Safety
Board made recommendations to expedite a programme to develop and instal equip-
ment which would facilitate the detection and classification, by severity, of thunder-
storms within 5 n mile of the departure or threshold ends of active runways at airports
having precision instrument approaches, and to instal equipment capable of detect-
ing variations in the speed of wind components along takeoff and approach flight
Modelling of gusts and wind shear 21

paths within 1 n mile of ends of active runways. The findings of this accident investi-
gation described, in particular, a headwind increase of about 15 kn followed later by
a headwind decrease of about 15 kn in 4 s. This type of wind variation has the basic
pattern illustrated in figure 6, and in the following section we describe a technique
for the assessment ofautoland systems which is directly relevant to the above type of
disturbance. In addition to the changes in headwind, the Eastern 727 also en-
countered significant down-drafts of a type which must also be modelled adequately
for aircraft assessment purposes.
Figure 15 shows a schematic diagram, reproduced from Goff(1975), of the type of
cold outflow ahead of a thunderstorm which is believed to be associated with the
squalls or gust fronts referred to above. Data (Goff 1975) obtained from tower
measurements made by the National Severe Storms Laboratory have resolved many
small details of thunderstorm outflow not previously observed and are currently being
analysed at the Royal Aircraft Establishment to provide quantitative parameters for
use in the discrete gust assessment technique for autoland systems described below.

4.2 Proposed alternative approach to turbulence simulation for autoland certification

The British CAA have expressed the view (CAA 1973) that the power-spectral type of
analysis and associated simulation programme which has been generally used to
assess autoland systems may not be the most satisfactory way to prove the safety
of the system, as operational experience and experience in flight testing suggests that
there may be a class of gusts, or combination of gusts, which arises more frequently
than the spectrum would suggest and which needs to be treated separately in some
way. As a result there is a need (CAA 1973) for an improved procedure for assessing
the effect of wind and turbulence on the safety of automatic landing systems. The
work required includes data collection, studies to improve and extend the wind model,
and any additions or improvements which may be desirable to simulation techniques.
As a result of the CAA views quoted above we have suggested an alternative
approach to turbulence simulation for autoland certification. A note describing
the new approach has been presented in Appendix 1 of Jones (1974). The principal
objective is to develop a single procedure for both the sudden gust (squall) ease and
the case of turbulence associated with relatively steady winds. The guiding principle
is the belief that in both these cases a relatively simple gust pattern consisting either
of a single ramp (figure 1) or of a pair of ramp gusts in opposite directions (such as
in figure 6), may be treated as a 'critical case' which dominates the situation. A major
change in the process of assessment, on the basis of the above approach based on
discrete-gust patterns, could be a move from the current process by which the aircraft

high turbutence

cotaair --.- - - \
(from thundemtorm-')-'~. . . . . . . . . ~ I~ose~warm air
.. :~...... ~ r ~'-..'~'~_:/V - .

under~ur~ent, ( high
turbutence
Figure 15. Schematicdiagram of thunderstorm outflow (reproducodfrom Goff 1975)
22 J G Jones

is validated on an explicit probability basis towards a procedure in which the proba-


bility concepts are implicit in the ' critical cases' prescribed in the formulation of the
requirements.
The basis of the proposed method lies in equations (4) and (5), which show that at
large amplitudes of response y, the most significant term in the expression Ny, denot-
ing the rate of occurrence of response amplitudes greater than y, is the ' tuned res-
ponse ', 7 (H---)or ~2. The gust, or gust pattern, associated with the tuned response
is the ' critical case' referred to above. Whilst equations (4) and (5) apply strictly
to the rate of occurrence per unit distance of large response amplitudes for flight in
turbulence, they may be generalised in a straightforward way to give the percentage
of occurrences on which, for example, the rate of aircraft descent at touchdown ex-
ceeds a specified value (Jones 1977). Equations of the form of (4) and (5) may thus
be used to give explicit consideration to probability in the assessment of a particular
system, where ' gust parameters' such as a, %, fl, fl~ would be specified in the turbu-
lence model on which the assessment was ba.~ed. On the other hand, we have out-
lined in Appendix 2 of Jones (1974) how equations such as (4) and (5) may be used to
derive design or assessment procedures o f ' design envelope' form which would be
based simply on a ' critical' response amplitude of the form U0 ~, (H), or (Uo)2 Y2,
where Uo and (U0)~ define families of design gusts (figure 10). Results of recent
research indicate that we should take (U0)2 : constant • U0 where the constant
lies in the range 0.70 to 0.85. Thus the explicit consideration of probability is re-
moved from the sphere of the aircraft designer, who is then concerned with applying
a completely deterministic procedure, to the responsibility of the air-worthiness
authority who would be concerned with the relationship between the amplitude of
U0 and the level of probability implicit in the requirement (see Appendix 2 of Jones
1974 for details of this relationship). It should be noted that this would be a move
towards the more traditional philosophy of gust assessment based on the concept of
' design gusts '.
Whereas, in equations (4) and (5), the variables defining the gust input refer to
the definition of the gust or gust pattern, in the case of autoland assessment it is
neccessary to include the positions of the component gusts with respect to the flight
path as additional variable~. Thus t h e ' search' for the critical case involves a larger
number of degrees of freedom. Corbin (1973) describes a deterministic computer
search procedure developed a t t h e Royal Aircraft Establishment based on variable
patterns of the form illustrated in figure 6 occurring at random p~sitions along the
flight path. A simplex search procedure is used to find the gust patterns and posi-
tions giving greatest rate of descent at touchdown or greatest range from desired
touchdown point. Figures 16 and 17 (reproduced from Corbin 1973) show the
gust time-histories for a particular control system in the respective cases:

205 SL

9 . wind /t S-=H I'"IJ',,l J, 'i


,.=r.o.ing doob,.rom0,.v,
/

time of
tou ch down
Figure 16. Horizontal ramp gusts causing greatest rate of descent at touchdown
(reproduced from Corbin t973)
Modelling of gusts and wind shear 23

20S

increasing
h.odwind ]
mls
Ht! ii!q~L
I I I I I
' " ' I" I ' I I.
,1 I14 ,
, t..lan
I I dotlbte romp giving
,I ,I ,I' , m o x i m u m r a n g e o t
I II I I I I J [ I I l l~ touchdown of 9 8 4 m
II~l [ I1 I I tl] It
tbne of
touchdown
Figure 17. Horizontal ramp gusts causing greatest error in range at touchdown
(Corbin 1973)

(i) horizontal gust pattern giving greatest rate of descent;


(ii) horizontal gust pattern rnaximising error in range.

In case (i) (figure 16) it cart be seen that the critical case for rate Of descent com-
prises an increase in headwind occurring 7 s before touchdown followed by a decrease
in headwind just prior (2-3 s) to touchdown. In case (ii) (figure 17), it can be seen
that the critical case for range is completely different, comprising a tailwind com-
ponent occurring 21 s before touchdown followed later by a headwind increase.
Subsequent work on a different control system suggests that the type of gust pattern
illustrated in figure 17 is typical of patterns causing maximum error in range, bur
that in the critical pattern causing maximum rate oi descent (figure 16) the initial
gust (increase in headwind) is only present if the system is in a sense ' overactive '.
It is intended to follow up this re.~ultas regards its implications fol piloting techniques
in eases of large gust distrurbances influencing manual landings.
A recent assessment by industry of the proposed approach to the assessment of
autoland systems is described in Rosenberg (1978) where the comment is made that
the technique could also be used as a design tool to aid in the development of the
automatic landing control laws of an aircraft, for instance, to identify over-sensitive
areas in which the control laws designed by traditional means may require modifica-
tion. The conclusion was reached that, for this purpose, the single discrete ramp
gust, using a raster scan and search technique (Rosenberg 1978), should be sufficient,
thus avoiding the need to perform more extensive searches involving pattenas com-
prising ramp-gust pairs.

5. Gust model for limit loads

5.1. Nature of large disturbances ~

A basic aim of research at the Royal Aircraft Establishment over the past 12 years
on the subject of atmospheric disturbances has been the study of severe gusts and
the situations in which they occur. As described in Burnham (1968a), catastrophic
accidents to civil transport aircraft in which encounters with gusts play a significant
part are in the main clearly linked with thunderstorms--the remainder being associat-
ed with airflow over mountains.
At a rather lower level of severity, a significant number of injuries (and occasionally
deaths) occur to passengers and cabin crew during flight through turbulence, in
almost all cases the victim not having been securely strapped into his seat. It appears
t h t t many such incidents are connected with storm encounters and some with gusts
24 J G Jones

associated with mountain wave systems. Very few seem to be associated with clear
air turbulence (CAT) in the absence of storm or mountain influence.
The introduction of the supersonic transport aircraft (SST) on an operational basis
has increased the range of altitudes over which turbulence models are required for
aviation purposes up to at least 60,000 ft (18,288 m), thus extending beyond the
tropopause level into the lower stratosphere. Results of work in this area carried
out in connection with the development of the SST are described in Burnham
(1968b).
Gusts large enough for their avoidance by civil aircraft to be desirable have been
found in RAE work in clear air near the tops of thunderstorms (up to 20 miles (32"19
kin) laterally and 10,000 ft (3,048 m) vertically from the top of the cloud) and in
association with mountain waves. As described in Burnham (1968a), one patch of
severe turbulence of the latter type, found at 46,000 ft (14,020 m) contained a
horizontal gust which reduced the indicated airspeed of the aircraft by 50 kn in 1 s, a
similar gust of opposite sign occurring some 20 s later.
Patches of severe turbulence, particularly those associated with storms, are often
very short. Data on large gust disturbances recorded during routine operational
flying (King 1967, 1969 and Morton 1966) indicate the importance of the short patch
and the relatively isolated gust. Experience shows that, although large loads may
occur within patches of rough air, nevertheless the really large loads can often be
singled out from the general level of roughness. It is not uncommon for a relatively
isolated large disturbance to occur at or near the boundary of a patch of turbulence.
This suggests that in this case the large disturbance was associated with a shear
layer on the boundary separating two distinguishable bodies of air. For the case of
flight at constant altitude, the overall pattern of disturbance is often compatible with
a penetration by the aircraft of the tilted or inclined boundary between two bodies
of air separated by a quasi-horizontal stratification having strong gradient in both
density and velocity. The form of the large disturbance influencing an aircraft
penetrating the bounding shear layer is sometimes suggestive of an encounter with
a vortex-like structure.
The statistical discrete-gust model of turbulence as described in w 2.4 comprises
gust patterns which employ the discrete-ramp gust illustrated in figure 1 as a basic
element or building brick. In this way, it proves possible to reproduce the essentially
non-Gaussian characteristics of turbulence velocity increments, large increments
tending to stand out from the background level of fluctuation. The aggregate of
discrete ramp gusts defined by equation (2) applies for values of gradient distance
H ~< L, where L is the scale length, and implies that the largest increments in velocity
are to be found over distances of the order of the scale length (see figure It)). Whilst
analysis of real turbulence samples has indicated that this model is compatible with
much of the atmospheric turbulence encountered, extremely large wind changes of
order 100 ft/s (30.48 m/s) are sometimes found over relatively short distances, of
order 150 ft (45.72 m) or less (substantially less than the scale length of the surround-
ing turbulence). These velocity increments suggest the possibility of a class of
phenomena of independent significance. Whilst the majority of outstanding wind
velocity increments encountered in the atmosphere may be associated with sections
of shears that are adequately modelled by the discrete gust of figure 1, the intr
wind gradient illustrated in figure 3 may be convincingly interpreted, as shown by
Anne Burns (1972), as an encounter with a vortex core. Whilst intense vortex cores,
Modelling of gusts and wind shear 25

with vertical axes, are well known in the form of tornado at low altitude (Morton
1966), the case documented in Anne Burns (1972) is of an aircraft encounter, at an
altitude of 45,000 ft, with an intense vortex core with a nearly horizontal axis in
the neighbourhood of storm tops.
It thus appears that, to complete the picture, we should include in our discrete-
gust model not only representations of discrete velocity fluctuations associated with
turbulence structures such as shear layers but also, at a level of relatively high inten-
sity but relatively low encounter probability, and thus influencing only the extreme
tails of the distributions, velocity fluctuations associated with sections of vortex
cores.
To summarise, we take the view that the ideal form for an overall discrete-gust
model of the atmosphere, representing the total family of discrete-gust encounters
experienced in the life-time of an aircraft or a fleet of aircraft, would take the form of
a sum of terms of the form of equation (2), each corresponding to a particular class
of atmospheric disturbances, and each having its appropriate scale length (i.e. maxi-
mum value of H). The separate representation of vortex cores is one step in this
direction. A particular advantage of this approach is that effects of aircraft operat-
ing procedures, and in particular storm-avoidance procedures, would be more easily
taken into account in the design criteria. However, this type of formulation awaits
such time as the necessary detailed statistical measurements are separately available
for the classes of atmospheric phenomena represented.

5.2. Proposedform of discrete-gust modelfor limit loads

In Jones (1973) we suggested ways in which the statistical discrete-gust theory may
be used as the basis of either ' design-envelope' or ' mission-analysis' criteria. In
Appendix 2 of Jones (1974) we have shown how an equation such as equation (4)
may be used to derive gust intensities for use in design-envelope studies. British
industry has expressed doubts (James 1973) about the practical use of the mission-
analysis concept (irrespective of the form of the turbulence model) and thus atten-
tion has recently been concentrated on the use of the statistical discrete-gust model
in design-envelope studies. Determination of appropriate design-gust intensities
(denoted by U0 in 9 4.2) has followed two courses. On the one hand, values have
been related to measured statistical distributions of extensive inflight counting-
accelerometer data. On the other, Uo values have been derived which, together
with appropriate calculations of aircraft response, produce estimated limit loads on
existing aircraft equal to the loads for which those aircraft were designed.
The current form of the proposed discrete-gust model is outlined below (but see
also 9 5.3). It should be emphasised that the particular model described is at present
entirely provisional and subject to further investigation.
The proposed gust model consists of three basic families which we denote by A, B
and C. Family A comprises gust patterns made up from discrete-ramp gusts as
described in 99 2 and 3, but includes patterns with up to eight component ramps to
cater for the case of response dominated by very lightly-damped modes. Making
the usual assumption of linear system response, the critical patterns comprising more
than two basic ramps cart be found by linear superposition in terms of successive
overswings in the response to a single ramp, just as the critical gust pair was found
in 9 3.2 by considering both 7 (H) and ~0 (H).
26 J G Jones

Family 2~ is made up o f ' vortex tubes' and is intended to represent atmospheric


vortices such as have been detected within thunderstorms and also sometimes in
clear air as described in w 5.1. These are relatively intense short gusts. Family C
is similar to B but with a different' scale length' (i.e. maximum vortex core diameter)
and is intended to represent aircraft trailing vortices (recent analysis of ' special
events' recorded in operational flying suggests the advisability of including this
family).
In terms of design gust amplitudes of the form W : UoH~/s, the three families of
gusts are illustrated in figure 18.
The discrete-gust profile for use with Family A is that illustrated in figure 1. Family
A gust amplitudes are given by

H1/s, H ~< 2500 ft (762 m),


W = ((U~ (6)
H > 2500 ft (762 m),

where L a ----2,500 ft (762 m) is the associated scale length. Patterns of ramp gusts
should be .considered and the critical patterns producing maximum response should
be found. The ' design amplitude' of the response is then given by

(u0)A 71
0.85 (u0)~ ~
(U0)a "~ = max
0.60 (uo)a ~
0.40 (Uo)A ~8, (7)

where, for the gust family W = H 1/s,~i istbe m a x i m u m response to an isolated gust,
~a the m a x i m u m response to a gust pair, and so on. The factors 0"85, 0"60 and 0"40
are provisional estimates and are based on turbulence analysis,and dynamic simula-
tion of the associated response of simple systems, using relativelylong samples of
measured 'moderate' turbulence. Appropriate values of (Uo)A have recently been
proposed by Card 0978) and include a variation of design-envelope gust amplitudes
with altitude as in existing airworthiness requirements. In more recent work, the
possible dependency of the appropriate factors (0"85, 0"60 and 0"40 above) on the
turbulence intensity has been investigated and it has bccn suggested (Jones 1978)
that for the representation of ' severe' or 'extreme ' turbulence the higher-order
factors should be reduced.
The gust profile proposed for use with Family B represents a section across a
vortex tube and is illustratedin figure 19. This profilehas been chosen by matching
to measured vortex data, particularlythose in Anne Burns (1972). Family B gust
amplitudes are given by (figure 18)

~ (U0)s H l/s,

w ---- ((tJo)a--~-,
L~s
H ~< 100 ft (30"48 m)

H > 100 ft (30"48 m),


(8)
Modelling of gusts and wind shear 27

W-,(Uo)A Hll3

A I

Lc'5Oft (15.24 m) I LA-2500ft (762m1 grodient


t distGnce
LB~lOOft (30.68m)
Figure 18. Proposed composite discrete-gust model for limit loads
I I
I
I I centre of profge as figure 1

taiL of profile based on


W

I
'L_ X
H ~ 3H ~

Figure 19. Profile to represent section across vortex tube

where L n = 100 ft (30"48 m) is a ' scale length' for the entire aggregate of vortex
cores. The H 1/3 part of the family (H ~ 100 ft (30"48 m)) is intended to represent
' equiprobable ' vortices of different sizes but of similar relative ages. There is no
direct experimental evidence for this H 1/3 law, but it is known that vortices with cores
of varying diameters exist, with H : : 100 ft (30"48 m) a typical ' large' diameter.
Clearly W--> 0 as H-> 0 and so some power law W ~-, H ~, 0 < k < 1, appears to
be appropriate. Since k ~ 1/3 holds approximately for other turbulence structures,
it seems reasonable to also take k ~ 1/3 for families of eqniprobable vortex cores
until such time as experimental evidence suggests an alternative (but see w 5.3). The
/../-1 family (H > 100 ft (30"48 m)) is included to represent different stages in the
decay of the design vortex that has H = 100 ff (30"48 m) initiaUy. There are good
fluid-mechanical reasons (conservation of circulation) for choosing this W ~ H -1
law.
Only single isolated gusts of Family B need be considered. The appropriate
choice of (U0)B is a subject for further investigation.
Family C is similar to Family B but with smaller core diameter Lc -----50 ft (15-24 m)
(figure 18). Current work on aircraft trailing vortices is helping to establish an
appropriate value for (Uo) c. The value (U0) c -- 17 ft/s (5"18 m/s) is being assumed
provisionally. The W , ~ H -1 relationship for H > 50 ft (15"24 m) is based on vor-
tex decay as for Family B. The W - ~ H x/s relationship for H ~ 50 ft (15"24 m) has
been shown in recent theoretical work (to be published) to correspond to trailing-
vortex sections measured at the minimum separation distance between aircraft, on
the assumption that the separation distance is taken to decrease with aircraft weight
in proportion to (weight/span) x/=.
28 J G Jones

One aspect of the use of Families B and C for limitqoad studies that is under
investigation is the extent to which aircraft flight-path encounters with the vortex
tube other than the transverse (constant gust velocity across the span) intersection,
usually assumed in discrete-gust applications, need to be considered. For example,
a critical case might be a longitudinal vortex, paraUel to the flight-path, producing
upwash on one side, and dowawash on the other side, of a mid-span wing section.
Questions of this type are particularly relevant in studies of gust-load alleviation
systems, for which the distributions of gust-induced loading over the surface of the
aircraft can be an important factor.

5.3. Incorporation of recent work on scale-dependent turbulence intermittency

The turbulence models that have been reviewed in this paper are based upon a
fluid mechanics background that relates back to scaling laws given in the classical
work by Kolmogorov in 1941. This led in particular to the well-known k -5/a law
for the energy spectrum in the inertial range and the associated H t/3 behaviour of
velocity increments. However, more recent work (for example Kolmogorov 1962),
indicates that the above scaling laws are not exact as they neglect the pheno-
menon of scale-dependent intermittency, in which the turbulence energy becomes
concentrated into smaller and smaller volumes as shorter and shorter wavelengths
are considered, leading to an effective concentration into sheet-like structures
at the fine scales at which the influence of viscosity becomes predominant and
dissipation occurs. Recent theoretical work on this phenomenon includes the
concept of self-similarity dimension D, where 2 ~< D ~< 3, due to Manddbrot (1977)
and an explicit geometrical model due to Frisch et al (1978). These theories reduce
to the classical (i.e. Kolmogorov 1941) case when D := 3. However, it is currently
proposed (Mandelbrot 1977; Frisch et al 1978) that for the more structured eddies
in turbulence a value of D < 3, such as D = 2"5, is more appropriate. We have
recently (Jones 1979) reviewed a selection of experimental data that supports the
use of a traditional self-similar model (D = 3) for turbulence fluctuations of low
to average intensity but are equally consistent with a scale-dependent intermittency
model with D ----2"5 for the higher intensities. Although direct evidence for the
choice of D is inconclusive at these higher levels of intensity, the current fluid-mecha-
nics background supports the use of a value such as D = 2"5 for the scaling of intense
gusts. In Jones (1979) we have thus made the suggestion that, for the prediction of
aircraft limit loads, where ' extreme' levels of turbulence are relevant, consideration
be given to the use of a model with D ~ 2"5 (leading to a revision of the proposals
in w 5.2 and Jones 1978), whilst a self-similar model with D ~ 3 (as reviewed in the
earlier parts of this report) be retained for such purposes as ride-quality or handling
qualities assessment, and possibly structural fatigue. In terms of velocity incre-
ments, this would imply (Jones 1979) using a W,~- H 1/6 scaling law at the highest
(e.g. limit-load) intensities but retaining a W ~ H 1/3 law at lower levels. The engi-
neering implementation of aircraft design or assessment procedures using the statis-
tical discrete-gust (SDG) theory, modified as outlined in Jones 0979) to take account
of scale-dependent intermitteney (say D = 2"5, W ~ H ~/6) would be no more difficult
in practice than is the case using the standard SDG theory. The implementation
in this latter ease has already been assessed as comparable with the use of PSD
methods in Card (1978).
Modelling of gusts and wind shear 29

The application to aircraft response of a turbulence model with self-similarity


dimension D < 3 has been discussed by Mandelbrot (1977) as follows: ' Every so
often the aircraft is shaken ~ibout, which shows that certain regions of the atmos-
phere are strongly dissipative, the remainder seeming to be l a m i n a r . . . w h e n a
strongly dissipative piece o f the cr0ss-section is examined in detail laminar inserts
become apparent and further, small inserts are seen when the analysis is defined
further'. As discussed in Jones (1979) the magnitude o f this effect will only be
significant in practice when comparing the response of aircraft having markedly
dissimilar dynamic characteristics. It should be noted that the expected future
trend towards increasing use of Active Control Technology, e.g. for gust-alleviation,
will lead to a wider range of dissimilarity between aircraft dynamic characteristics
than has been the case in the past and it is likely to be in this context that the
improved modelling of gusts and wind shear for aircraft assessment and certification
will find its principal application.
This paper is a revised version o f a paper presented at a C A A R C Symposium
on Operational Problems held at Bangalore in October 1976, and published in the
Proceedings of that Symposium (ARDB-Proc.-701). The paper is copyright 9
Controller, HMSO, London, 1979.

References

Burnham J 1968a Atmospheric gusts--a review of the results of some recent RAE research. RAE
Technical Report 68244
Burnham J 1968b Atmospheric turbulence and the SST: a review in the light of recent research.
RAE Technical Report 68096
Burnham J & Colmer M J 1969 On large and rapid wind fluctuations which occur when the wind
had previously been relatively light. Aeronautical Research Council CP No. 1158
Burns, Anne 1972 On the nature of largo clear air gusts near storm tops. RAB Technical Report
72036
CAA 1973 The use of a wind model in the certification of automatic landing systems. Unpublished
CAA note
Card V 1978 An analytical investigation of the statistical discrete gust method of aircraft gust loads
prediction. BAe (Weybridgo-Bristol)BAe/WBD/D/M/283
Chen W-Y 1972 On the application of Rice's exceedance statistics to atmospheric turbulence. AIAA
Paper No 72-136
Corbin M J 1973 Turbulence time-histories causing greatest touchdown errors following automatic
flare. RAE Technical Memorandum Avionics 152
Davies P O A L 1975 Order in disorder: studies in turbulence. University of Southampton,
Inaugural Lecture
Davies P O A L & Yule A J 1975J. FluidMech. 69 513
Dutton J A 1970 Progress in aerospace sciences ed. D K~chomarm (Oxford: Pergamon) Vol. 11
Dutton J A & Thomson G J 1968 Probabilistie determination of aircraft response to turbulence at
low altitudes. Air Force Aeronautical Systems Division Technical Report ASD-TR-68-39
Frisch V, Sulem P-L & Nelkin M 1978J. Fluid. Mech. 87 719
Gerlach O H, van de Moesdijk G A J & van der Vaart J C 1973 Progress in the mathematical
modelling of flight in turbulence. AGARD Conference Proceedings, CP-140
Goff R C 1975 Thunderstorm-outflow kinematics and dynamics. National Severe Storms Labo-
ratory, NOAA Technical Memorandum ERL NSSL-75
James D O N 1973 Structural loads and gust criteria. AGARD Conference Proceedings CP-140
Jones J G 1967 Gradient properties of a model of stationary random turbulence. Aeronautical
Research Council C P No. 998
Jones J G 1968 A theory for extreme gust loads on aircraft based on the representation of the atmos-
phere as a self-similar intermittent random process. RAE Technical Report 68030 (ARC30592)
30 J G Jones

Jones J G 1971 A unified discrete-gust and power-spectral treatment of atmospheric turbulence.


RAeS/CASI/AIAA International Conference on atmospheri~ turbulence, London
Jones J G 1973 Statistical discrete gust theory for aircraft loads: A progress report. RAE Technical
Report 73167
Jones J G 1974 UK research on aeronautical effects of surface winds and gusts: AGARD Report
No. 626 on Effects of surface winds and gusts on aircraft design and operation
Jones J G 1977 Notes on the application of the statistical discrete gust method to autoland assess-
ment. RAE Technical Memorandum FS 129
Jones J G 1978 On the formulation of gust-load requirements in terms of the statistical discrete
gust method. RAE Technical Memorandum FS208
Jones J (3 1979 On self-similarity, fractal dimension and aircraft response to gusts. RAE Technical
Memorandum FS 244
Jones J G 1979 O n the representation of fluctuatingsignalenergy in time and frequency. R A E Tech-
nical Memorandum FS 230
King G E 1967 Civil aircraftairworthiness data recording programme. Study of sevele turbulence
encountered by civilaircraft.R A E Technical Report 67106
King (3 E 1969 Civil aircraft airworthiness data recording programme. Some characteristics of
severe turbulence. R A E Technical Report 69150
Kolmogorov A N 1941 Dokl. Acad. Sci. U S S R 30 4
Kolmogutov A N 1962 J. Fluid Mech. 13 82
Kttchemama D 1971 Turbulence and vortex motions. ZFW 19 part 9 p 305
Kuznetsov P I, Stratonovich R L & Tikhonov V I 1965 Nonlinear transformations of stochastic
processes (Oxford: Pergamon Press)
Mandelbrot 1977 Fractals: form, chance and dimension, (San Francisco: W H Freeman)
Meeeham W C & Siegel A 1964 Phys. Fluids 7 1178
Middleton D 1960 Introduction to statistical communication theory (New York: McGraw Hill)
Morton B R 1966 Progress in aeronautical sciences ed. D Kltchemann (Oxford: Pergamon) Vol. 7
Payne B W & Cox R A 1969 Aircr. Engg,
Reeves P M 1969 A non-gaussian turbulence simulation. University of Washington, Dept. of Aero-
nautics and Astronautics, College of Engineering Report 69-6. Also released as Air Force
Flight Dynamics Laboratory Technical Report AFFDL-TR-69-67
Reeves P M, Campbell G S, Ganzer V M & Joppa R G 1974 Development and application of a
non-gaussian atmospheric turbulence model for use in flight simulators. NASA Contractor
Report CR-2451
Reeves P M, Joppa R G & Ganzer V M 1976 A non-gaussian model of continuous atmospheric
turbulence for use in aircraftdesign, N A S A Contractor Report CR-2639
Rosenberg K W 1978 Hybrid computer investigationof discretegust and wind shear effectson auto-
matic landing system performance; A G A R D Flight Mechanics Panel Specialist Meeting,
Ottawa
Roshko A 1976 Structure of turbulent shear flows: a new look. AIAA Paper No. 76-78
Saffman P G 1968 Lectures on homogenous turbulence, Topics in non-linear physics, ed. N J Zabusky
(Berlin: Springer-Verlag)
Tomlinson B N 1975 Developments in the simulation of atmospheric turbulence. AGARD Flight
Mechanics/Guidance and Control Panels, Joint Symposium on Flight Simulation/Guidance
Systems Simulation, The Hague, Netherlands
van de Moesdijk G A J 1975 Simulation of patchy atmospheric turbulence based on measurements
of actual turbulence. AGARD Flight Mechanics]Guidance and Control Panels, Joint Sympo-
sium on Flight Simulation/GuidanceSystem~ Simulation, The Hague, Netherlands
van de Moesdijk G A J 1978 Non-gaussian structure of the simulated turbulent environment in
piloted flight simulation. Paper presented at AGARD/FMP Specialists Meeting, Brussels

You might also like