You are on page 1of 97

FACULTY MECHANICAL, MARITIME AND

MATERIALS ENGINEERING
Delft University of Technology Department Marine and Transport Technology
Mekelweg 2
2628 CD Delft
the Netherlands
Phone +31 (0)15-2782889
Fax +31 (0)15-2781397
www.mtt.tudelft.nl

Specialization: Transport Engineering and Logistics

Report number: 2012.TEL.7733

Title: Wind load on a crane

Author: S.P. Oostlander

Title (in Dutch) Windbelasting op een kraan

Assignment: literature assignment

Confidential: no

Initiator (university): -

Initiator (company): -

Supervisor: Ir. W. van den Bos

Date: February 14, 2013

This report consists of 97 pages and 0 appendices. It may only be reproduced literally and as a whole. For
commercial purposes only with written authorization of Delft University of Technology. Requests for consult are
only taken into consideration under the condition that the applicant denies all legal rights on liabilities concerning
the contents of the advice.
FACULTY OF MECHANICAL, MARITIME AND
MATERIALS ENGINEERING
Delft University of Technology Department of Marine and Transport Technology

Mekelweg 2
2628 CD Delft
the Netherlands
Phone +31 (0)15-2782889
Fax +31 (0)15-2781397
www.mtt.tudelft.nl

Student: S.P. Oostlander Assignment type: Literature


Supervisor (TUD): Ir. W. van den Bos Creditpoints (EC): 12
Specialization: TEL
Report number: 2012.TL.7733
Confidential: No

Subject: Wind load on a crane

During the students research assignment [2011.TEL.7600], in which cyclists aerodynamic position on
the bicycle has been investigated, the interest in wind engineering, wind loads and wind effects grew.

For this literature assignment the world of cycling is changed in the world of cranes and large
structures. Books, papers and the in the Netherlands applicable standards for wind loads on cranes
have been investigated.

An overview about used wind pressures, force coefficients and calculation methods is given.

The professor,

Prof. P. Professor
Preface
This report is the result of my, Sander Oostlander, literature assignment. The assignment has been
performed as part of my Master of Science Transportation Engineering and Logistics at the Technical
University of Delft.

This research has given me the opportunity to increase my knowledge on wind engineering and
working with standards. Working with multiple national, European and American standards Ive come
to know more about how wind loads are calculated, coefficients determined and which dynamical
effects play a part in crane design.

3
Summary
When designing a crane, its obligatory for the design to satisfy with the applicable, national
standards. In the Netherlands the NEN organization is responsible for producing the in the
Netherlands applicable standards. All international standards from CEN and ISO must also be
approved by NEN, before they are obligatory. This report contains: NEN 2018, NEN 13001, ISO 4301,
Eurocode 1 and ASCE 7/10.

These standards describe all (wind) loads and combined load cases which a crane must resist during
its lifetime. Extreme wind conditions (out-of-service conditions) can lead to overload, loss of static
equilibrium and failure of the structure. Moderate, but fluctuating, wind loads (in-service conditions)
can also strongly contribute in the fatigue of a crane.

Wind is an air flow with a certain density, velocity and direction. A wind flow can generate a pressure
on a surface (q) resulting in a force (F). De magnitude of the force depends on the projected area of
the structure (A) and the aerodynamic drag coefficient (C).

The wind pressure (q) is a function of the height (z), the wind velocity (v) and the air density (). The
fundamental wind velocity is recorded at a height of 10meters above flat and open country and
averaged over a chosen time period. Commonly a 3 second mean wind velocity or a 10 minute mean
wind velocity with a return period of 50 years. So an annual risk of exceeding this wind velocity of
p=0.02. The fundamental wind velocity is a static wind velocity. The fundamental wind velocity is
increased or decreased by the absence or presence of nearby structures, hills, nature etc. The
resulting mean wind velocity is still a static wind velocity. In reality a wind flow is not a static flow but
dynamic constantly fluctuating flow. The dynamics of a wind flow are expressed by the gust factor
which is a function of the turbulence intensity. Turbulence is the ratio of the standard deviation of the
wind velocity and the mean wind velocity.

( ) ( ) ( )

The fundamental wind velocity used to determine the wind pressure differs over the standards. Table
A gives an overview of the different static fundamental wind velocities found in the standards used in
this report. Figure A shows where each region or wind area can be found in the Netherlands.

Figure B gives an overview of the resulting wind pressure of the used standards, as a function of the
height of the structure. The difference in wind pressure between the older standards (NEN 2018) and
the newer standards (NEN 13001 and NEN 1991) is big. This mainly because the newer standards use
a factor for turbulence and the older dont take turbulence into account.

4
Table A Fundamental wind velocities

Figure A Locations of the


fundamental wind velocities from
table A

Wind pressure, NEN 1991, NEN 13001, NEN


2018 and ASCE 7-10
4.000
Wind pressure q [N/m2]

3.500
3.000
2.500
2.000
1.500
1.000
500
00
0 50 100 150 200 250
Height z [m]

NEN 13001 out-of-service C NEN 13001 out-of-service D


NEN 2018 out-of-service NEN 1991 WA I and Terrain Cat. 0
NEN 1991 WA II and Terrain Cat. II ASCE 7-10 Inland

Figure B Wind pressure, calculated using the fundamental wind velocities form table A

The aero dynamical coefficients (C) are given in a similar way in all standards. The newer standards
contain extended tables of aero dynamical coefficients which are influenced by the wind angle,
structures dimensions, cross section shapes and details like rounded corners. The aero dynamical
coefficients given are basic values which can be reduced, depending on the structures solidity and
slenderness. The solidity is the ratio between the projected area (A) of all members and the total
enclosed area (Ac). As shown in figure C. The slenderness is at least the ratio between the length (l)
and the characteristic diameter of a member (h). As shown in figure C. In the newer standards the
position of the member within the structure influences the slenderness.

5
Figure C Solidity and slenderness of a structure or member

All standards allow wind tunnel testing or CFD analysis to determine aero dynamical coefficients.
When a structure is sensible to dynamical behaviour or the first natural frequency is below 1Hz wind
tunnel testing is also advised, for a crane this is often the case.

A wind load is one of the many loads acting on a structure. All loads combined lead to a load case, in
which each load has its own contribution and resulting safety factor. Each load is classified:
Permanent, Variable or Accidental. In-service wind loads are variable actions and out of service wind
loads are accidental actions. Wind loads contribute in the occurring stresses and the body stability.

A structure can very well overload under non-maximal wind loads when a dynamic response occurs.
Dynamic behaviour can be caused by the fluctuating nature of the wind flow, by the structures shape
which generates an oscillating wind flow or by a motion of the structure itself which causes an
increase of the wind load. An energy build-up can cause a structure to oscillate until failure. Dynamic
behaviour can be avoid by avoiding critical cross sections, adding spoilers, splitters or adding
mechanical coupling and by increasing the structures stiffness and damping.

Dynamic behaviour can be investigated using wind tunnel research, but only when using an,
expensive, aero-elastic support system instead of a simple force balance. A common force balance
does not allow dynamical behaviour and can only be used to determine aerodynamic coefficients. A
wind tunnel model crane needs to be constructed in such a way that it behaves identical to the real
life crane. Therefor multiple dimensionless coefficients need to be checked. The wind flow inside the
wind tunnel should contain a wind velocity which varies with the height, a boundary layer.

All standards use alike methods, but still differ. The difference is made in the chosen wind velocities
and the amount of detail of the aerodynamic coefficients. Wind loads from standards remain rough
estimates which ensure a safely designed structure. The growth of world trade is will force the
number of standards to reduce and become more consistent. More consistency will develop more
detailed and worldwide applicable standards. Because each part of the world has its own climate, and
resulting wind velocity, temperature, air density etc. a worldwide standard can contain much more
data and detail in determine the wind pressure. The knowledge on wind dynamics and the resulting
wind loads is a constantly developing field, in which wind tunnel and CFD testing provides more and
more insight. Constantly increasing and detailing the data in aerodynamic coefficient tables. A
worldwide standard can increase and accelerate the knowledge on wind loads on structures.
6
Summary (in Dutch)
Bij het ontwerpen van een kraan is het verplicht te voldoen aan de nationaal geldende normen. In
Nederland is de NEN organisatie verantwoordelijk voor het maken en voorschrijven van de normen.
Alle internationale normen van CEN en ISO moeten ook eerst door de NEN getoetst worden, voordat
ze geldend zijn. Dit rapport bevat de volgende normen: NEN 2018, NEN 13001, ISO 4301, Eurocode 1
en ASCE 7/10.

Deze normen beschrijven alle (wind) belastingen en belastingcombinaties die een kraan moet kunnen
weerstaan tijdens zijn leven. Extreme windcondities (buiten bedrijf) kunnen lijden tot overbelasting en
het bezwijken van de constructie. Mattige, maar fluctuerende, windbelastingen (in bedrijf) kunnen ook
een grote bijdrage leveren aan de vermoeiing van een constructie.

Wind is een lucht stroming met een zekere dichtheid, snelheid en richting. A wind stroming kan een
druk genereren op een oppervlak (q) wat resulteert in een kracht (F), welke constant of fluctuerend
is. De grote van de kracht wordt bepaald door het frontaaloppervlak (A) en de aerodynamische
cofficint (C).

De winddruk (q) is een functie van de hoogte (z), de windsnelheid (v) en de luchtdichtheid (). De
fundamentele windsnelheid wordt geregistreerd op 10meter hoogte boven vlak en open gebied en
gemiddeld over een gekozen tijd. Gebruikelijk is een 3 seconde of 10 minuten gemiddelde wind
snelheid die iedere 50 jaar voorkomt. Dus een jaarlijkse kans op overschrijding van p=0,02. Deze
fundamentele windsnelheid is een statische windsnelheid. De fundamentele windsnelheid wordt
verhoogd of verlaagd door de aanwezigheid of afwezigheid van bebouwing nabij, heuvels, begroeiing,
etc. Dit resulteert in een gemiddelde, statische, windsnelheid. In werkelijkheid is een windstroming
niet statisch maar dynamisch en constant veranderend. Deze dynamica van een windstroming wordt
uitgedrukt door de vlaagfactor, welke een functie is van de turbulentie intensiteit. Turbulentie is de
ratio tussen de standaard afwijking van de windsnelheid en de windsnelheid.

( ) ( ) ( )

De gebruikte fundamentele windsnelheid verschilt tussen de normen. Tabel I geeft een overzicht van
de fundamentele windsnelheden gebruikt in dit rapport. Figuur I geeft een overzicht van waar welke
wind regio of gebied zich bevindt.

Figuur II geeft een overzicht van berekende winddrukken uit de gebruikte normen, als functie van de
hoogte van de constructie. Het verschil tussen de oude normen (NEN 2018) en de nieuwere normen
(NEN 13001 en NEN 1991) is groot. Dit komt voornamelijk doordat the nieuwere normen turbulentie
meenemen in de berekening en de oudere normen doen dit niet.

7
Tabel I Fundamentele windsnelheid

Figuur I Locaties van de


fundamentele windsnelheden uit
tabel I

Wind pressure, NEN 1991, NEN 13001, NEN


2018 and ASCE 7-10
4.000
Wind pressure q [N/m2]

3.500
3.000
2.500
2.000
1.500
1.000
500
00
0 50 100 150 200 250
Height z [m]

NEN 13001 out-of-service C NEN 13001 out-of-service D


NEN 2018 out-of-service NEN 1991 WA I and Terrain Cat. 0
NEN 1991 WA II and Terrain Cat. II ASCE 7-10 Inland

Figuur II Winddruk, berekend met de fundamentele windsnelheden uit tabel I

De aerodynamische weerstandscofficinten kunnen gevonden worden in tabellen en grafieken in de


normen. De nieuwere normen bevatten uitgebreidere tabellen waarin de weerstandscofficinten
worden benvloed door de windhoek, afmeting, vorm van de doorsnede en details zoals afronding van
hoeken. De gegeven aerodynamische weerstandscofficinten zijn basis waarden die gereduceerd
kunnen worden, afhankelijk van de volheid en de slankheid. De volheid van een element is de ratio
tussen het frontaal oppervlak (A) en het omsloten oppervlak (Ac). Zoals te zien in figuur III. De
slankheid is de ratio van tenminste de lengte (l) en de karakteristieke diameter (h) van een element.

8
Zoals te zien in figuur III. In de nieuwere normen wordt de slankheid ook benvloed door de positie
van het element binnenin de constructie.

Figuur III Volheid en slankheid van een constructie of element

Alle normen staan ook windtunnel en CFD analyses toe om de aerodynamische cofficinten te
bepalen. Als een constructie gevoelig is voor dynamisch gedrag of de eerste eigenfrequentie is lager
dan 1Hz worden windtunnel testen ook geadviseerd, voor een kraan is dit vaak het geval.

De windbelasting is n van de vele belastingen op een constructie. Alle belastingen gezamenlijk


vormen een belastingscombinatie, waarbinnen elke belasting zijn eigen bijdrage heeft, afhankelijk van
de veiligheidsfactor. Elke belasting is te classificeren als: Permanent, Variabel of Incidenteel.
Windbelastingen tijdens bedrijf zijn variabel en buiten bedrijf incidenteel. Wind belastingen dragen bij
aan de optredende spanningen in en aan het statisch evenwicht van een constructie.

Een constructie kan overbelast worden bij een niet maximale windbelasting wanneer er een
dynamische response optreedt. Dynamisch gedrag kan worden veroorzaakt door een fluctuerende
wind, door de vorm van de constructie zelf wanneer er een oscillerende loslating ontstaat of door de
bewegingen van de constructie zelf wanneer deze de windbelasting vergroten. Energie kan worden
opgebouwd in de constructie, zodanig dat deze oscilleert totdat de uitwijking tot breuk leidt.
Dynamisch gedrag kan worden vermeden door: kritische doorsneden te vermijden, spoilers of vinnen
toe te voegen, mechanische verbindingen te maken en door de stijfheid en demping van de
constructie te verhogen.

Dynamisch gedrag kan worden onderzocht in een windtunnel. Echter hiervoor is dan een kostbaar
aero-elastische ondersteuning nodig, i.p.v. een eenvoudige krachten balans. Een gebruikelijke vaste
balans kan wel gebruikt worden om de aerodynamische cofficinten te bepalen maar niet om
dynamisch gedrag vast te stellen. Een windtunnelschaalmodel moet zodanig gemaakt worden dat zijn
gedrag identiek is met de werkelijke kraan. Hiervoor moeten de dimensie loze schaal grootheden
worden gecontroleerd. De windstroming moet een snelheidsprofiel bevatten, een grenslaag.

Alle normen gebruiken gelijkaardige methodes, toch verschillen ze. De groei van de wereldwijde
handel zal lijden tot minder en meer consistente normen. Meer consistentie zal lijden tot
gedetailleerdere en universeel toepasbare normen. Omdat iedere wereld deel zijn eigen klimaat heeft,
en daardoor dus eigen windsnelheden, temperatuur, luchtdichtheid, etc. zal een wereldwijde norm
over meer detail en data kunnen beschikken als het gaat om het bepalen van de winddruk. Wind
9
dynamica en de veroorzaakte windbelastingen is een gebied waarover de kennis continu toeneemt.
Windtunneltesten en computer simulaties geven meer en meer inzicht. Waardoor constant de data
over aerodynamische weerstand cofficint toeneemt en gedetailleerder wordt. Een wereldwijde
standaard zou dit proces van toenemende kennis over windbelastingen op constructies kunnen
versnellen.

10
List of symbols
Symbol Unit Description
2
A m Frontal area of the member
2
Ac m Enclosed area of a lattice structure
Afr - Friction area, swept by the wind
Aref - Reference area
b m Reference with of the cross section
2
B - Background factor
C - Aerodynamic drag coefficient of the member
c0 - Aerodynamic coefficient for an infinite length member
c0 - Topography coefficient
ca - Aerodynamic coefficient
Cdir - Constant for wind direction
cf - Force coefficient
cfr - Friction coefficient
cpe - External pressure coefficient
cpi - Internal pressure coefficient
cr - Roughness coefficient
cscd - Structural factor
Cseason - Constant for wind velocity per season
d m Characteristic dimension
frec - Recurrence factor
Fwind N Wind load
G - Gust factor
g - Peak factor
hdis m Displacement height
Iv(z) - Turbulence intensity at a height z
Kd - Directionally factor
kr - Terrain factor
Kz - Exposure coefficient
Kzt - Topographic coefficient
L years Lifetime of a structure
l0 m Length of a member
ni - Natural frequency, mode i
2
p N/m Wind pressure
q N/m2 Wind pressure
2
qb N/m Basic wind pressure
qp(z) N/m2 Wind pressure at a height z
r - Risk of exceeding a curtain wind velocity during the

11
structures lifetime
R years Return period of a curtain wind velocity
2
R - Resonant factor
Re - Reynolds number
Sc - Scruton number
St - Strouhal number
v m/s Wind velocity
vb m/s Basic wind velocity
vb,0 m/s Fundamental basic wind velocity
vCG m/s Critical galloping wind velocity
vcrit m/s Critical wind velocity
vm(z) m/s Mean wind velocity at a height z
vref m/s Reference wind velocity
vs m/s Wind velocity
vw m/s Wind velocity
vwind.max.gust m/s Maximum gust wind velocity
vwind.mean m/s Mean wind velocity
z m Height above the surface
z0 m Roughness length
zeff m Effective height above the surface
zmin m Minimal height above the surface
T - Relative aerodynamic length
- Shielding factor
- Aerodynamic slenderness
kg/m3 Density
v m/s Standard deviation of the wind velocity
- Solidity ratio
- Reduction factor for the aerodynamic coefficient
r - Reduction factor for the roundness of the corners
- Reduction factor for the slenderness

12
Contents
Preface ....................................................................................................................................... 3
Summary .................................................................................................................................... 4
Summary (in Dutch) .................................................................................................................... 7
List of symbols .......................................................................................................................... 11
1. Introduction ....................................................................................................................... 15
1.2 Standards ................................................................................................................... 15

1.3 Wind .......................................................................................................................... 16

1.4 Basic wind load formula ............................................................................................... 17

2. Wind pressure .................................................................................................................... 18


2.1 Wind velocity modelling ............................................................................................... 18

2.2 Standards ................................................................................................................... 21

2.2.1 NEN-EN 13001-2:2011................................................................................................... 21


2.2.2 NEN 2018:1983.............................................................................................................. 24
2.2.3 ISO 4302: 1981 .............................................................................................................. 26
2.2.4 NEN-EN 1991-1-4:2005 ................................................................................................. 27
2.2.5 ASCE/SEI 7-10 ................................................................................................................ 32
2.3 Comparison of the standards, wind pressure.................................................................. 35

3. Wind load .......................................................................................................................... 38


3.1 Wind load modelling .................................................................................................... 38

3.2 Wind load example: Crane hotel ................................................................................... 38

3.3 Standards ................................................................................................................... 42

3.3.1 NEN-EN 13001-2:2011................................................................................................... 42


3.3.2 NEN 2018:1983.............................................................................................................. 48
3.3.3 ISO 4302:1981 ............................................................................................................... 51
3.3.4 NEN-EN 1991-1-4:2005 ................................................................................................. 52
3.4 Comparison of the standards, wind load ........................................................................ 58

4. Design wind load ................................................................................................................ 63


5. Dynamic wind effects .......................................................................................................... 65
5.1 Dynamic response ....................................................................................................... 65

5.2 Structural factor, NEN-EN 1991-1-4............................................................................... 66

5.3 Wind effects and vibrations .......................................................................................... 68

5.3.1 Buffeting ........................................................................................................................ 68


5.3.2 Vortex shedding ............................................................................................................ 68
5.3.3 Galloping ....................................................................................................................... 73

13
5.3.4 Divergence and flutter .................................................................................................. 76
6. Wind tunnel research .......................................................................................................... 79
6.1 Scale model ................................................................................................................ 80

6.2 Wind flow ................................................................................................................... 81

6.3 Support system ........................................................................................................... 82

7. Numeric example................................................................................................................ 83
7.1 Numeric example crane ............................................................................................... 83

7.2 Results numeric example ............................................................................................. 84

7.2.1 NEN-EN 13001-2:2011................................................................................................... 84


7.2.2 NEN 2018:1983.............................................................................................................. 86
7.2.3 ISO 4302:1981 ............................................................................................................... 87
7.2.4 NEN-EN 1991-1-4:2005 ................................................................................................. 88
7.3 Comparison of the results ............................................................................................ 89

7.4 Wind tunnel vs. NEN 13001 vs. NEN 2018 ..................................................................... 90

7.5 NEN 13001 vs. NEN 2018 ............................................................................................. 92

8. Conclusion ......................................................................................................................... 93
References ................................................................................................................................ 96

14
1. Introduction
In my research assignment I performed a wind tunnel test with cyclists. This has generated my
fascination for wind and wind effects. In this research two methods to determine the wind load acting
on real cyclists have been investigated and compared. Being a mechanical engineering student my
interest is not only in cycling but also in mechanical structures. Therefor Ive chosen to perform my
literature study on: determine the wind load on a crane.

The goal of this literature assignment is increasing my knowledge on how the wind load on a large
structure is determined, using standards. The design of cranes is regulated by national and
international standards. Each designed crane must comply with the applicable standards, the use of
these standards is obligatory. Standards specify general conditions, requirements and methods to
prevent hazards of cranes by design and theoretical verification.

The main research questions are:


How is the wind pressure determined and modelled?
How are the shape coefficients of the structures member determined?
How is the wind load on a crane determined?
Which dynamical effects are possible to occur with large structures?

The obligatory standards are of the utmost importance in a crane design, so the standards are
recurring parts of this report. Paragraph 1.2 introduces the national and international standards used
in this report, 1.3 gives the basics about wind and 1.4 the basic physic formula which all standards
use, with added coefficients. Chapter 2 deals with the wind flow, wind velocity and resulting wind
pressure. How is this determined and how do the standards come to a design wind pressure. Chapter
3 deals with how the wind load is calculated. So which aerodynamic coefficients are used to calculate
the wind load, using the design wind pressure? Chapter 4 deals with how wind loads are combined
with all other loads acting on a structure. Chapter 5 further examines the dynamic wind effects which
are possible to occur. Chapter 6 pays attention to how wind tunnel research is used to determine the
wind load on a crane. Chapter 7 gives an example of using the standards. The last chapter, chapter 8,
gives a conclusion and answers to the research questions.

1.2 Standards
In the Netherlands the NEN organisation is responsible for producing the national standards. The NEN
is a member of the CEN (Comit Europen de Normalisation) and of the global organisation ISO
(International Organisation for Standardization).

When the CEN publishes a European standard (EN) the national organisation NEN can copy the
standard, making the standard into a NEN-EN. If necessary the NEN-EN can even be translated into
the Dutch language.

15
When needed a national annex (NB) can be added to the NEN-EN. The national annex can add
information to the NEN-EN and override the NEN-EN.

When working with a standard one should always check if the correct, current and applicable standard
is being used.

In the Netherland these four national standards deal specifically with wind loads on cranes:
NEN-EN 13001-2:2011 en April 2011
NEN 2018:1983 October 1983
NEN 2022 March 1976
ISO 4302:1981 May 1981

More general national standards on wind loads on constructions and referred to in the standard for
wind loads on cranes are:
NEN-EN 1990+A1+A1/C2 December 2011
NEN-EN 1990+A1+A1/C2/NB December 2011
NEN-EN 1991-1-4:2005 nl April 2005
NEN-EN 1991-1-4+A1+C2/NB December 2011

The American national standard for Minimum Design Loads for Buildings and Other Structures is
produced by the ASCE (American Society of Civil Engineers).
ASCE/SEI 7-10 2010

From ASCE 7 especially the wind velocity and wind pressures are interesting, since the east coast of
America is well-known with heavy storms and hurricanes.

1.3 Wind
What is wind? The Dutch van Dale dictionary gives the following meaning to the word wind.

Wind: A sensible, mostly horizontal air flow in the atmosphere.

Wind is an air flow with a certain density and velocity. A wind flow causes a pressure on a surface
which is not parallel to its direction. Wind flows are generated by the pressure differences in the
earths atmosphere. These pressure differences are generated by the heating of the atmosphere by
the sun. Warm air expands resulting is a low pressure zones, with a lower density. Cold air results in
high pressures zones, with a high density. Because nature always seeks for equilibrium a flow is
generated from the high pressure zone to the low pressure zone, wind. Because of the rotation of the
earth and the constant temperature differences between the poles and the equator there will never be
equilibrium and therefore there will always be wind.

16
The density and velocity of the wind flow determines the generated pressure. But this pressure is not
always a constant value. Large pressure zones generate a large, mean wind flow. But locally there are
also pressure differences in a flow causing small fluctuations in the wind velocity and direction. A
perfect smooth wind flow, with no local pressure differences, a constant velocity and direction is called
a laminar flow. A turbulent wind flow is full of local pressure differences, velocity variations and
direction changes.

1.4 Basic wind load formula


For calculating the wind loads all standard use the same basic physics formula
q = wind pressure [N/m2]
C = aerodynamic coefficient [ - ]
A = projected area [m2]

The wind pressure, aerodynamic coefficient and projected area are determined differently in the
standards. The complexity differs between the standards.

17
2. Wind pressure
This chapter deals with the methods used in the standards to determine the wind pressure caused by
the wind velocity, direction, air density, location and nearby structures. The wind pressure is a
function of the wind velocity, the air density and the surrounding landscape en structures. Both wind
pressure and velocity are dealt with in this chapter.

Paragraph 2.1 gives a general introduction on how a static design wind velocity is made out of
dynamic wind velocity data. Paragraph 2.2 contains the standards used. Paragraph 2.3 compares the
standards.

2.1 Wind velocity modelling


Establishing an appropriate design wind velocity is a critical step in the calculation of the wind load on
a structure. It requires the statistical analysis of historical data on recorded wind velocities. Recording
of wind velocities commonly takes place at 10meters above flat and open country. The maximum wind
velocity each year, month, week etc. is recorded. All recorded storms must be statistically
independent. Its also possible to set a threshold value and to only record and count the storms which
exceed this threshold value.

The statistical analysis of the wind velocity is often done using the type III generalized extreme value
distribution, (Jenkinson, 1955). This distribution has an upper limit, which we would expect the
atmosphere indeed has. When looking at fatigue not only the maximum wind velocities is interesting,
but also the distribution of the wind velocity as a whole.

Figure 1 shows the wind speed recorded around Melbourne, Australia, as a function of the return
period.

Figure 1 Extreme wind velocities, Holmes (2001)

18
The risk of the wind velocity exceeding during the lifetime of a structure (r) can be calculated using
the return period (R) and the lifetime (L). A commonly used return period for the wind velocity is 50
years.

[ ] R = return period of the wind velocity

L = life time of the structure

The wind velocity changes with the height above the ground. Figure 2 shows a graph of recorded
wind velocities at three heights. The gusty or turbulent nature of the wind is constant at all heights.
The lower frequencies are also quit the same at all heights, the higher frequencies differ.

Figure 2 Wind velocities at 3 heights, Holmes (2001)

A layer of wind in which the wind velocity is increasing is known as the boundary layer and is caused
by the influence of the earths surface to the wind flow. The boundary layer can extend up to 1km.
The mean wind velocity vector may locally change in direction and magnitude as a function of height.
This change in direction is small over the height range of normal structures and normally therefor
neglected in wind engineering.

The wind velocity distribution in the boundary layer is a logarithmic function, derived by Prandtl and
influenced by the air density and surface shear stress. However a power law is often used to describe
the wind velocity distribution, because integrating becomes much easier. Figure 3 shows the wind
velocity distribution in a boundary layer, as a function of height, for both methods.

The turbulence or gustiness of a wind flow is expressed as the standard deviation or root-mean-
square of the wind velocity as a function of time. The turbulence intensity is the ratio between the
standard deviation and the mean wind velocity. The turbulence and fluctuating nature of wind flows
can have a significant effect on wind pressures, drag coefficients and the resulting wind load on
structures.

19
Figure 3 Wind velocity (U) in the boundary layer, Holmes (2001)

The effect of the turbulence to the wind velocity is expressed by the maximum gust wind velocity.
Which stands for the mean wind velocity times the gust factor. In reality the gustiness will not act
simultaneously at all heights. Topography can considerably increase the gust factor.

G = gust factor, , g = peak factor and v =

standard deviation of the wind velocity, the turbulence.

The effect a gust of wind has on a structure depends on the scale of the gust and the dimensions of
the structure. Large structure will never be entirely enveloped by a single gust. Small structures or
small parts of a structure can be enveloped by a single gust, resulting in a higher and more fluctuating
wind loads.

The wind pressure (q) is calculated by multiplying the maximum gust wind velocity with the air
density. This leads to the assumption of quasi-static wind pressure. This assumption ignores pressure
fluctuations over the structure.

= air density, common = 1,25 kg/m3

v = mean wind speed

20
2.2 Standards
The following paragraph deals with how the common and crane specific wind loading standards
determine the wind pressure.

2.2.1 NEN-EN 13001-2:2011


The NEN-EN 13001-2 standard divides wind loads in two categories: occasional in-service wind
pressure q(3) and exceptional out-of-service wind pressure q(z).

( ) ( ) q(3) = in-service wind pressure at v(3)


= density of the air
v(3) = gust wind velocity averaged over a period of 3s
( ) = mean wind velocity, averaged over 10 minutes at in 10m
height above flat ground or sea level
1,5 = factor between 3s and 10min mean wind velocity

NEN-EN 13001-2 distinguishes the three in-service wind states light, normal and heavy which depend
on the type of crane. Table 1 and figure 4 show the three wind states (1, 2 and 3), mean wind
velocities ( and ( )), resulting wind pressures ( ( )) and the Beaufort scale (X). The Beaufort wind
scale wind velocity is a 10minutes mean wind velocity. The in-service wind load is assumed to act
perpendicularly to the longitudinal axis of a crane member and constant at all heights.

Table 1 In-service wind states [NEN-EN 13001-2:2011 Table 4]

Figure 4 Correlation of the mean wind velocity, wind pressure and


Beaufort scale [NEN-EN 13001-2:2011 Figure 7]

21
The design of the in-service wind load is based on the following requirement for the operation of the
crane: If the wind velocity, measured at the highest point of the crane tends to reach the wind state:
heavy, the crane shall be secured or its configuration shall be transformed into a safe configuration.

The out-of-service wind load is calculated using the assumption that the wind blows horizontally at a
velocity increasing with the height (z) above the surrounding ground level.

( ) ( ) q(z) = wind pressure


v(z) = the 3s mean equivalent static out-of-service
wind velocity

( )
( ) [ ]

( ) [( ) ] frec = a factor depending on the recurrence (R) of the

extreme out-of-service wind condition. R = 5 years


frec = 0,8155 for R = 50 years frec = 1,0
vm(z) = 10 min mean storm wind velocity in the
height z
vref = 10min. mean reference storm wind velocity.
Differs over regions in Europe (see figure 5).
8 = gust response factor (8 = 1,1)
vg = is a 3s gust amplitude beyond the
10min mean storm wind, with K = 0,0055.

( ) = simplified roughness coefficient

Figure 5 and table 2 show the map of Europe with the four wind regions and the reference wind
velocities in these regions. These wind velocities are 10minute mean wind velocities with a recurrence
interval of 50 years. The reference wind velocity is increased with a factor depending on the height of
the crane and the recurrence.
z = 1m 1,0(0,72+0,4)=1,23 v(1) = 1,12vref
z = 25m 1,0(1,13+0,4)=1,68 v(25) = 1,53vref
z = 100m 1,0(1,38+0,4)=1,96 v(100) = 1,78vref
z = 200m 1,0(1,52+0,4)=2,11 v(200) = 1,92vref

Table 2 Reference storm wind velocities, 10min mean. [NEN-EN 13001-2:2011 Table 6]

Figure 6 shows a graph of the in-service wind pressure for wind state 2 and 3 and the out-of-service
wind pressure for region C (inland of the Netherlands) and D (north-west coast of the Netherlands).

22
Figure 5 Reference storm wind map of Europe [NEN-EN 13001-2:2011 Figure 12]

Wind pressure, NEN-EN 13001-2:2011


2.500

2.000
Wind pressure q [N/m2]

1.500

1.000

500

00
0 50 100 150 200 250
Height z [m]

In-service state 2 normal In-service state 3 heavy


Out-of-service region C Out-of-service region D

Figure 6 Wind pressure in- and out-of-service NEN-EN 13001-2:2011

23
2.2.2 NEN 2018:1983
NEN 2018 divides the wind load in wind and storm conditions, Sw and Ss. The wind load is assumed to
act horizontally, in the least favourable direction.

q = wind pressure

vw = wind velocity

The value of the wind velocity is found in NEN 2018 table 20 and a function of the type of crane (A, B
and C), height and wind condition (wind or storm). All wind velocities and wind pressure are based on
a 3s mean wind velocity. See table 3 for the table (in Dutch). Because all wind velocities vw are based
on 3s mean wind velocities they dont correspond to the wind velocities in the Beaufort scale! Table 4
shows the Beaufort scale, 10min. mean wind velocities and wind pressures.

The maximum 10min. mean wind velocity at which a crane type A may operate in is: 14/1,5=9,3m/s
which corresponds to wind force 5. For crane type B the maximum wind force is 6 and for crane type
C wind force 8. The storm wind corresponds to a wind force of 9 to 11, depending on the height.

Figure 7 shows a graph of the in-service wind pressure for crane type B and C and the out-of-service
wind pressure for crane type B and C.

Table 3 Wind load on a crane [NEN 2018 table 20]

24
Table 4 Beaufort wind scale [NEN 2018 table 21]

Wind pressure, NEN 2018


1.600

1.400
Wind pressure q [N/m2]

1.200

1.000

800

600

400

200

00
0 50 100 150 200 250
Height z [m]

In-service crane type B In-service crane type C


Out-of-service crane type B and C

Figure 7 Wind pressure in- and out-of-service NEN 2018

25
2.2.3 ISO 4302: 1981
The ISO 4302 standard uses a single methodology for calculating the wind pressure.

p = wind pressure
vs = wind speed

The design wind speed for in-service conditions depends on the type of crane and is shown in table 5.
ISO 4302 presumes that the wind load is applied in the least favourable direction. For the out-of-
service wind velocities ISO 4302 refers to appropriate national standards. The wind pressure may be
used as an constant, as a step function or a function of the height. The in-service wind pressure from
ISO 4302 is identical to NEN 2018. So the in-service wind pressures in figure 7 also apply for ISO
4302.

Table 5 In-service design wind velocity and pressure [ISO 4302]

26
2.2.4 NEN-EN 1991-1-4:2005
NEN-EN 1991-1-4 also known as and referred to as: the Eurocode 1, is the general standard in Europe
for wind actions. In this standard wind pressure calculations are much more complex and contain
many factors. The wind velocity and pressure is made up by an average and fluctuating component.
The average component is represented by mean wind velocity vm(z) and the fluctuating component by
the turbulence intensity Iv(z). The wind actions are characteristic values. There is an annual risk of
0.02 that the wind actions are being exceeded of r=0,02. This agrees with a recurrence period of
R=50 years.

The fundamental wind velocity NEN-EN 1991-1-4 uses is vb,0. This is the 10 minute mean wind
velocity, independent of direction and season, at 10 meters height above the surface in a terrain
category II.

Cdir = direction factor (Cdir = 1 in the Netherlands)


Cseason = seasonal factor (Cseason = 1 in the Netherlands)
vb,0 = fundamental wind velocity

Figure 8 shows the three wind areas in the Netherlands. At adjacent wind areas a linear relation
between both wind velocities over 5km length must be used to determine the wind velocity. Table 6
shows the fundamental 10min. mean wind velocities for each wind area.

Table 6 Fundamental wind velocity [NEN-


EN 1991-1-4+A1+C2:2011/NB:2011
Table NB.1]

Figure 8 Wind area's in the


Netherlands [NEN-EN 1991-1-
4+A1+C2:2011/NB:2011 Figure NB.1]

27
The mean wind velocity vm(z) is the mean wind velocity as a function of the height, under influence of
the environment.

( ) ( ) ( ) cr(z) = roughness coefficient


co(z) = topography coefficient (co(z) = 1 in the Netherlands)

The roughness coefficient cr(z) at height z is defined by two logarithmic profiles

( ) ( ) for

( ) ( ) for

( )

kr (terrain factor), z0 (roughness length) and zmin are parameters depending on the terrain category.
Table 7 shows these factors for three terrain categories in the Netherlands. Figure 9 shows possible
locations for terrain category 0 in the Netherlands.

Table 7 Factors for the roughness coefficient cr [NEN-


EN 1991-1-4+A1+C2:2011/NB:2011 Table NB.3]

Figure 9 Possible locations for terrain


category 0 (bold line) [NEN-EN 1991-1-
4+A1+C2:2011/NB:2011 Figure NB.4]

28
Terrain factor kr for all three terrain categories

TC 0: kr = 0,162 TC II: kr = 0,209 TC III: kr = 0,223

Roughness coefficients for four heights and all three terrain categories

z=1 cr,TC 0(1) = 0,858 vm(1) = 0,858vb


cr,TC II(1) = 0,626 vm(1) = 0,626vb
cr,TC III(1) = 0,589 vm(1) = 0,589vb

z = 25 cr,TC 0(25) = 1,38 vm(25) = 1,38vb


cr,TC II(25) = 1,01 vm(25) = 1,01vb
cr,TC III(25) = 0,872 vm(25) = 0,872vb

z = 100 cr,TC 0(100) = 1,60 vm(100) = 1,60vb


cr,TC II(100) = 1,30 vm(100) = 1,30vb
cr,TC III(100) = 1,18 vm(100) = 1,18vb

z = 200 cr,TC 0(200) = 1,72 vm(200) = 1,72vb


cr,TC II(200) = 1,44 vm(200) = 1,44vb
cr,TC III(200) = 1,34 vm(200) = 1,34vb

The resulting mean wind velocity vm(z) is a 3s mean wind velocity, without unsteady turbulence
effects. Including the turbulence effects means calculating the turbulence intensity Iv(z). Which at any
height is the standard deviation of the wind velocity divided by the mean value of the wind velocity.

( )
( )
kl = turbulence factor (kl = 1 in the Netherlands)
( ) ( )

The peak velocity pressure qp(z) contains the mean steady flow and the influence of the unsteady
turbulent flow.

( ) ( ( )) ( )

( ) ( ) [ ( ) ( ) ]
( )

29
Example locations for each situation in the Netherlands, shown in figure 10.

A. Wind Area I and terrain category 0 Den Helder


B. Wind Area I and terrain category II Centre of the province of North Holland
C. Wind Area I and terrain category III Alkmaar
D. Wind Area II and terrain category 0 Rotterdam, Maasvlakte 2
E. Wind Area II and terrain category II Rotterdam, Botlek
F. Wind Area II and terrain category III Rotterdam, city centre
G. Wind Area III and terrain category II Uncultivated areas in all Dutch provinces
without a coast line
H. Wind Area III and terrain category III Cities and towns in all Dutch provinces,
without the coast line: Breda, Utrecht,
Arnhem, Zwolle, etc.

Figure 10 Example locations of WA en Terrain Cat.

Figure 11 shows for eight possible wind situations the wind pressure. Based on the wind areas and
the terrain categories.

30
Wind pressure, NEN-EN 1991
3.000

2.500

2.000
Wind pressure q [N/m2]

1.500

1.000

500

00
0 50 100 150 200 250
Height z [m]

A: NEN 1991 WA I and Terrain Cat. 0 B: NEN 1991 WA I and Terrain Cat. II
C: NEN 1991 WA I and Terrain Cat. III D: NEN 1991 WA II and Terrain Cat. 0
E: NEN 1991 WA II and Terrain Cat. II F: NEN 1991 WA II and Terrain Cat. III
G: NEN 1991 WA III and Terrain Cat. II H: NEN 1991 WA III and Terrain Cat. III

Figure 11 Wind pressure Eurocode 1, the Netherlands

31
2.2.5 ASCE/SEI 7-10
The American standard describes wind pressure as a function which is mostly influenced by the basic
wind speed (V). The basic wind speed is specified by ASCE 7 as a 3s gust wind speed at 10m height in
open terrain with a 3% probability of exceedence in 50 years.
0,613 = a constant for air density [kg/m3], may be
changed when sufficient climatic data is available to
justify the change
Kz = velocity pressure exposure coefficient
Kzt = topographic factor
Kd = wind directionality factor, a value of 1 or less
V = basic wind speed

The basic wind speed is determined by the risk category of the structure, for a crane category risk
category: III + IV. Figure 12 and 13 show a map of the USA with the corresponding basic wind
velocities. Special regions, like mountainous terrain, gorges and special wind regions need to be
examined for unusual wind conditions. The wind speed is in: miles per hour (meter per second).

Figure 12 Basic wind speed for cat. III + IV [ASCE 7-10, Figure 26.5-1B]

32
A
B

C
Figure 13 Basic wind speed for cat. III + IV [ASCE 7-10, Figure 26.5-1B] +
Locations A,B and C for wind pressure calculations

In ASCE 7 all factors increase the wind pressure. In the Eurocode all factors increase the wind
velocity. The velocity pressure exposure coefficient is a function of the surface roughness and the
structures height. For port side cranes an exposure category D is used.

( ) for z < 5m (z = structures height)


( ) for 5m < z < 213,36m

z = 1m Kz = 1,05
z = 25m Kz = 1,38
z = 100m Kz = 1,76
z = 200m Kz = 1,99

The topographic factor takes wind velocity change due to slopes and hills into account. For flat and
open country a value of 1 is used.

The wind directionality factor depends on the structure type and can be found in ASCE/SEI 7-10,
Table 26.6-1. For a crane a conservative value would be 0.95.

33
The gust-effect factor is not included in the calculation of the wind pressure but in the calculation of
the resulting wind force. A crane is a flexible and/or dynamically sensitive structure so the gust-effect
factor needs to be hand calculated.

Figure 14 shows the wind pressure for three possible locations (see figure 13):

A. Inland Basic wind velocity 54m/s


B. East coast Basic wind velocity 72m/s
C. Florida Basic wind velocity 89m/s

Wind pressure, ASCE 7-10, risk cat. III and IV


10.000

9.000

8.000
Wind pressure q [N/m2]

7.000

6.000

5.000

4.000

3.000

2.000

1.000

00
0 50 100 150 200 250
Height z [m]

A: ASCE 7-10 Inland B: ASCE 7-10 East Coast C: ASCE 7-10 Florida

Figure 14 Wind pressure, ASCE 7-10

34
2.3 Comparison of the standards, wind pressure
The standards display great similarity in wind pressures, but still differ. NEN-EN 13001-2, NEN 2018
and ISO 4302 make a distinction between in-service wind pressure and out-of-service wind pressure.
The NEN-EN 1991-1-4 and ASCE 7 only provides the out-of-service, maximum, wind pressure. NEN-EN
1991-1-4 and NEN-EN 13001-2 also add the fluctuating effect of turbulence to the wind velocity or
resulting pressure.

The wind pressure in all standards is calculated using the multiplication of wind velocity and air
density. The used wind velocity defers. NEN-EN 13001-2, NEN 2018 and ASCE 7 uses a gust wind
velocity averaged over a period of 3s. NEN-EN 1991-1-4 uses a fundamental wind velocity which is
averaged over 10 minutes, which is increased with a factor to a 3s mean wind velocity.

Because of the great differences between the defined basic wind velocities in the standards around
the world, a general classification system for design wind speeds and levels should be generated.
Such a system could create an unambiguous and clearly defined basic wind velocity for different levels
or categories. Table 8 shows such a possible system, generated by Holmes (2001).

Table 8 Possible international wind velocity distribution, Holmes (2001)

From table 8 you can conclude that the factor between the 3s mean wind velocity and 10min mean
wind velocity should be 1,5 - 1,6. NEN-EN 13001-2 uses a factor 1,5.

Table 9 shows a comparison of the (out-of-service) basic wind velocities in both 3s and 10min mean
value used in the paragraph 2.2. All Dutch standards use similar wind velocities, when using the same
average time. The American wind velocities, for risk cat. III and IV are much higher than the Dutch.
Mainly because of the completely different climate and the possibility off hurricanes.

Figure 15 shows a comparison of the out-of-service wind pressures for the Dutch standards and for
the American. Its clear that the much higher basic wind velocity in America results in a much higher
wind pressure. Because wind pressure is the square of the wind velocity the differences are even
clearer between the Dutch and American standards.

35
The difference between the older NEN2018 and the newer NEN 13001 and the Eurocode 1 is quite
big. This mainly because the newer standards increase the wind pressure with a fluctuating
component. In figure 16 the out-of-service wind pressure of NEN 2018 and the Eurocode 1 wind
pressure with and without the fluctuating component are shown. Without the fluctuating component
the Eurocode 1 wind pressure is similar to the NEN 2018 wind pressure. Clearly showing that the
newer standards use newer wind models which should result in more accurate wind modelling.

Table 9 Basic wind velocities, 3s and 10min mean values

Wind velocity 3s [m/s] 10min [m/s]


13001 48.0 32.0 Region D
42.0 28.0 Region C
2018 46.0 30.7 z > 100m
42.0 28.0 20 < z < 100
36.0 24.0 z < 20
1991 44.3 29.5 Wind Area I
Eurocode 1 40.5 27.0 Wind Area II
36.8 24.5 Wind Area III
ASCE 7-10 89.0 59.3 Florida
72.0 48.0 East coast
54.0 36.0 Inland

In all Dutch standards an air density of =1,25kg/m3 is used. In the American =1.23kg/m3. When a
crane is placed on a location with great diversity in weather multiple air densities should be
investigated. A cold and dry frizzing winter day results in a 30% higher wind pressure than a warm
and moist summer day.

-25C sea level =1,418


40C sea level 100% humidity =1,093

36
Wind pressure, NEN 1991, NEN 13001, NEN
2018 and ASCE 7-10
4.000
3.500
Wind pressure q [N/m2]

3.000
2.500
2.000
1.500
1.000
500
00
0 50 100 150 200 250
Height z [m]

NEN 13001 out-of-service C NEN 13001 out-of-service D


NEN 2018 out-of-service NEN 1991 WA I and Terrain Cat. 0
NEN 1991 WA II and Terrain Cat. II ASCE 7-10 Inland

Figure 15 Wind pressure, the Dutch and American standards

Wind pressure, NEN 2018 and Eurocode with


and without fluctuating component
3.000

2.500
Wind pressure q [N/m2]

2.000

1.500

1.000

500

00
0 50 100 150 200 250
Height z [m]

NEN 1991 WA I and Terrain Cat. 0


NEN 1991 WA I and Terrain Cat. 0 without turbulence
NEN 1991 WA I and Terrain cat. II
NEN 1991 WA 1 and Terrain cat. II without turbulence
NEN 2018 out-of-service

Figure 16 Wind pressure, with and without the fluctuating component

37
3. Wind load
This chapter deals with determine the wind load. When the wind pressure in known as a constant or
as a function of the height, multiple coefficient need to be determined to calculated the force
generated by the wind pressure. These coefficients contain aerodynamic shape constants, projected
areas, reduction factors, shielding factors etc.

Paragraph 3.1 gives a general introduction on how the wind load is calculated. Paragraph 3.2
describes an example crane which is used to visualize the standards which are described in paragraph
3.3. Paragraph 3.4 compares the standards.

3.1 Wind load modelling


The established wind pressure is used to determine the resulting wind force on a structure. The
magnitude of the wind force depends on the frontal area of the structure and its shape, the
aerodynamic coefficient. When the quasi-static wind pressure is used a quasi-static wind load is
calculated. Therefor the wind load must be multiplied with a dynamic constant which accounts for the
dynamic effects.

Due to the turbulent nature of wind flows, wind loads will act highly fluctuating. When this fluctuating
frequency resembles the natural frequency of the structure, or parts of it, there is a potential to a
resonant response. Especially for all natural frequencies below 1Hz. The amount of resonant response
of a structure depends on the aerodynamic or structural damping. The magnitude of the structural
forces generated by the response depend on the mass, damping and stiffness of the structure. The
turbulent nature of wind flows can also, at moderate wind loads, have a significant contribution to the
fatigue load of a structure. Chapter 5 will go deeper into dynamical behaviour.

3.2 Wind load example: Crane hotel


The wind load standards in this chapter will be shown together with an example based on a crane
located in Amsterdam. The crane shown in figure 17 will be transformed into a Hotel. This will change
the purpose of the crane into a hotel and therefor also the design wind load. Figure 18 and 19 show
the dimensions of the crane hotel. All without the
added minute structure for the hotel. The crane is
vertically divided in multiple sections for which one
general aero dynamical coefficient needs to be
determined. The wind load on the crane is
determined in the feathering position (position a)
and with the boom perpendicular to the wind flow
(position b). The calculated wind load is including
the safety factor for stability.

Figure 17 Crane Hotel Amsterdam

38
Figure 18 Dimensions crane hotel, including the divided area's

39
Figure 19 Crane hotel with area A7b within the green box

40
All sections are described below. The dimensions are rough estimates based on computer drawings.
Exact dimensions where not obtained. The purpose of this example is to provide a practical application
of the abstract standards, so the lack of exact and detailed dimension is no problem.

Section A1 Area A1 is the rail support system of the crane. All four support systems have
an area of 13.3m2 (total 53.2m2) and an average height of 0.792m above ground. Dimensions: 12m
long, 2m high, 1m depth and a spacing between the rails of 10m.

Section A2 Area A2 is a horizontal beam connecting the lower parts of the legs. One on
each side with an area of 5.1m2 (total 10.2m2) and an average height of 2.634m. Dimensions: 15m
long, 1,5m high, 1m depth and a spacing between the beams of 10m.

Section A3 Area A3 are the legs of the crane. Each leg has an area of 9.2m2 (total
36.8m2) with an average height of 5.463m. Dimensions: 10m long and 2.5m wide. Spacing of 10m.

Section A4 Area A4 is the rotational part of the crane. Its simplified to a closed box with
an area of 30.7m2 with an average height of 9.947m Dimensions 10m wide and 6m high.

Section A5 Area A5 is the tower part which is not converted into the hotel rooms. It
contains a plate structure at the core and a lattice structure. The projected area is: 45.3m2 and the
enclosed area is: 77.3m2 all at an average height of 17.8m. Dimensions: Height 15m, width 8m and
beam width 1m.

Section A6 Area A6 is the tower part which contains the hotel rooms. Its simplified to
one closed box. Area of 75.6m2 at an average height of 34.478m. Dimensions Height 15m and width
8m.

Section A7a Area A7a is the top beam in feathering position and simplified as on big box
(figure 18). Area 24.6m2 at an average height of 44.374m. Dimensions Height 10m and width 8m.
Because the top beam has a depth of over 60m there are many member shielding behind each other
in the feathering position. These shielded members contribute to the wind load, to take this into
account the projected area is doubled.

Section A7b Area A7b is the top beam perpendicular to the wind flow, as shown in figure
19. The top beam is a lattice structure, just like section A5, with an enclosed area of 267m2 and a
projected area of 85m2 at an average height of 44.374m. A second lattice frame is shielding 2m
downstream of the first lattice frame. All dimensions for area A7b have been obtained from the
original microfilm drawings of the crane. Sadly these scanned drawings lack resolution so some
dimensions are determined roughly.

41
3.3 Standards
The following paragraph deal with the common and crane specific wind loading standards which are
currently active in the Netherlands.

3.3.1 NEN-EN 13001-2:2011


Annex A in NEN-EN 13001-2 provides a method, tables and graphs to determine the aerodynamic
coefficient (ca) and characteristic area (A). Aerodynamic coefficients may also be derived from
theoretical or experimental methods. For instance results from CFD or wind tunnel testing.

= reduction factor, depends on the aerodynamic


slenderness () and solidity ratio () see figure 22
c0 = aerodynamic coefficient of a member with infinite length.
NEN-EN 13001-2, table A.2 A.5

The aerodynamic slenderness

= la / d la = aerodynamic length of a member


d = characteristic dimension
l0 = length of the member
= relative aerodynamic length, can be a function of the
Reynolds number. See figure 20.

Figure 20 Relative aerodynamic length [NEN-EN 13001 Table A.1]

42
The solidity ratio (see figure 21)


AJ = area of the individual members of a lattice structure

Ac = enclosed area of a lattice structure

Figure 21 Solidity ratio of a lattice structure member [NEN-EN 13001 Figure A.2]

Figure 22 Reduction factor [NEN-EN13001-2, Figure A.1]

The aerodynamic coefficients of round shapes (for instance lattice structures) are a function of the
Reynolds number (Re). This results in a non-linearly relation between the wind load and the wind
pressure.

v = v(3) or v(z)

When a crane consists of multiple members behind each other shielding can become in issue. The
shielding factor () depends on the solidarity and the spacing ratio a/d, where a = distance between
the members, d = characteristic dimension (see figure 23). Figure 24 shows the shielding factor.

43
Figure 23 Multiple arrangement of structural members [segment of: NEN-EN 13001 Table A.6]

Figure 24 NEN-EN 13001-2, Figure A.9, Shielding factor

The characteristic area A of the whole structure

= shielding factor

nm = number of parallel and identical members


A1 = characteristic area of the first member

The wind load acting perpendicularly to the longitudinal axis of a crane member in-service

( )

The wind load acting perpendicularly to the longitudinal axis of a crane member out-of-service

( )

44
For the crane hotel the resulting wind loads are shown in table 10 including all coefficients used for
calculating. Figure 25 28 show segments of tables form NEN 13001 used for the reduction factor,
the shielding factor and aero dynamical coefficients.
Table 10 Wind load crane hotel, NEN 13001

Solidity For all section except A5 and A7b the solidity is 1. Section A5 has an enclosed area of
77.3m divided by the projected area of 45.3m2 resulting in a solidity of 0.59. For A7b the solidity is
2

0.32.

Slenderness The slenderness of the members and elements of each section depends on the
dimensions, cross sections and within the structure. The relative aerodynamic length of sections A1,
A4 A7a and A7b is 1, for A2 and A3 it is infinite.

Reduction factor Using figure 22 and the calculated solidity and slenderness the reduction
factor has been determined. Shown in figure 25.

Shielding factor The shielding factor is determined using the distance between the members
and figure 24. For section A1 and A3 the projected area is reduced because of shielding of the front
rail support and legs. The projected area of A5 is increased with 12% because the framework at the
back side is largely shielded but not entirely. For section A7a the projected area is doubled, because
the top beam has an depth of over 60meters. For section A7b the projected area is increased because
a second identical lattice frame is 2m downstream, with 30%. See figure 26 for how the shielding
factors are determined using NEN 13001-2 figure A.9.

45
Basic aerodynamic coefficients For sections A1, A2, A4, A5, A5, A7a and A7b NEN 13001 Table A.3
have been used, see figure 27. For section A3 NEN 13001 Table A.4 has been used, see figure 28.
Because the crane structure is simplified to seven sections not much detail is left. NEN 13001 is a very
detailed standard with extensive table off aerodynamic coefficients. So more detailed crane sections
could lead to a more accurate and perhaps reduced wind load.
A1: b/d=0.5 A2: b/d=0.7 A3: b/d=1 A4 - A7a and A7b: b/d=1

Wind pressure For the calculations a wind region D and C has been used to determine the wind
pressure q (figure 6).

Safety factor For the safety factor load combination C2 from NEN 13001 Table 11 has been used:
1,16. See chapter 4.

Figure 25 Reduction factor for the crane hotel, NEN 13001

Figure 26 Shielding factor for the crane hotel, NEN 13001

46
Figure 27 Segment of: NEN 13001 Table A.3

Figure 28 Segment of: NEN 13001 Table A.4

47
3.3.2 NEN 2018:1983
The aerodynamic coefficients (C) in NEN 2018 are a function of the aerodynamic slenderness and
presented in a table. NEN 2018, table 22. See table 11 (in Dutch). Favourable aerodynamic
coefficients may be applied, if obtained by wind tunnel experiments.

The aerodynamic slenderness

a = l / h or l/D l = length of a member


h = height of the member perpendicular to the wind direction
D = Outside diameter of the member

When the crane consists of multiple members behind each other a reduction factor is applied to the
aerodynamic coefficient. The reduction factor can be found in NEN 2018 table 23 and NEN 2018 figure
13. See figure 29 for the graph.

( ) C = aerodynamic coefficient, function of the slenderness


= reduction factor, function of the solidity ratio (same as
NEN 13001), distance between the members b and the
height of the member h.

A = area of the individual members of a lattice structure [m2]

Ae = enclosed area of a lattice structure [m2]

The wind load acting horizontally in the least favourable direction

Table 11 NEN 2018, table 22, aerodynamic coefficients

48
Figure 29 NEN 2018: figure 13, reduction factor

For the crane hotel the resulting wind loads are shown in table 12 including all coefficients used to
calculate. Table 12 shows the used aerodynamic coefficients.

Table 12 Wind load crane hotel, NEN 2018

Solidity For all sections except A5 and A7b the solidity is 1. Section A5 has an enclosed area
of 77.3m2 divided by the projected area of 45.3m2 resulting in a solidity of 0.59. For A7b the solidity is
0.32.

Slenderness The slenderness of the members and elements of each section depends on the
dimensions, cross sections and within the structure.

Shielding factor The shielding factor is determined using the distance between the members.
For section A1 and A3 the projected area is reduced because of shielding of the front rail support and
legs. The projected area of A5 is increased with 12% because the framework at the back side is
largely shielded but not entirely.

49
For section A7a the projected area is doubled, because the top beam has a depth of over 60meters.
For section A7b the projected area is increased because a second identical lattice frame is 2m
downstream, with 30%. See figure 26 for how the shielding factors are determined. NEN 2018 and
NEN 13001 use an identical graph for the shielding factors.

Basic aerodynamic coefficients For all sections table 13 shows the aerodynamic coefficients
used. The slenderness and b/d ratio determines the value. For the aerodynamic coefficients of section
A1, A3, A4 and A5 a simplified a boxed member / flat plate has been used. Section A2 is simplified as
a boxed section with a h/b of 1,5. Sections A5, A7a and A7b are simplified to a single lattice structure.

Wind pressure For calculations the out-of-service wind pressure from NEN 2018 (figure 7) has been
used.

Safety factor For the safety factor the general factor 1,4 has been used for cranes with stowing
devices. See chapter 4.

Table 13 Aerodynamic coefficients used for the crane hotel [NEN 2018, table 22]

50
3.3.3 ISO 4302:1981
The aerodynamic coefficients (Cf) in ISO 4302 are a function of the aerodynamic slenderness and
presented in ISO 4302, table 2. See table 4. Aerodynamic coefficients obtained by wind tunnel or full
scale tests may also be used. When the crane consists of multiple members behind each other a
reduction factor (shielding factor ) is applied to the aerodynamic coefficient. The reduction factor can
be found in ISO 4302, table 3. See table 5.

( ) C = aerodynamic coefficient, function of the slenderness


= shielding factor, function of the solidity ratio , spacing
ratio: distance between the members a and the height of the
member b.

A = area of the individual members of a lattice structure [m2]

Ae = enclosed area of a lattice structure [m2]

The wind load acting in the least favourable direction

The crane hotel example is skipped because ISO 4302 tables are identical to NEN 2018.

Table 14 ISO 4302-1981, Table 2, Force coefficients

Table 15 ISO 4302-1981, Table 3, Shielding factors

51
3.3.4 NEN-EN 1991-1-4:2005
The Eurocode 1 provides two methods to determine the wind load. Indirectly by summing the
components of surface pressures and friction stresses over the structure. Or directly by force
coefficients appropriate to the whole structure. In both methods the peak velocity pressure qp(z) is
calculated for a reference height ze. The value of the reference height depends on the ratio between
the width and height of the structure. Figure 30 shows how the reference height is determined and
the resulting peak velocity pressure distribution over the structure.

Figure 30 NEN-EN 1991-1-4, Figure 7.4, reference height ze

Indirect method

[ ( ) ] [ ( ) ] ( )

Direct method

[ ( ) ]

cscd = structural factor


cpe = external pressure coefficient
cpi = internal pressure coefficient
cfr = friction coefficient
cf = force coefficient
Aref =reference area, function of reference height ze
Afr = friction area, area parallel to wind
52
Friction forces may be ignored when the swept area is less than four times the total area of windward
and leeward surfaces, which is the case for cranes. The NEN-EN 1991-1-4 does not supply pressure
coefficients for structural elements with rectangular sections. The direct method is further explained.

The structural factor cscd accounts for the combined effect of the non-simultaneous action of peak
wind pressures over faces of the structure (size effect = cs) and the vibration of the structure in its
fundamental mode due to the action of turbulence (dynamic response = cd). See paragraph 5.2 for
more information about the structural factor.

When slender buildings (h/d > 4) are arranged in groups it is possible upwind structures generate
turbulent wakes onto downwind structures and increasing their dynamic response. This is called wake
buffeting. When the natural frequency of the structure is less than 1 Hz or unknown, wind tunnel tests
or specialists advice is recommended. The natural frequency of a reasonably stiff container crane is in
the range of f = 0,70Hz (J.Verschoof, 1999).

The aero dynamic coefficients are called force coefficients given in multiple tables in the Eurocode 1.

cf,0 = aerodynamic coefficient of a member with sharp corners


and infinite length
r = reduction factor which accounts for rounding of corners
= reduction factor which accounts for the length of the
element, end-effect factor

The reference area Aref is the projected or frontal area

b = characteristic diameter
l = length of the element

Graphs for the aerodynamic coefficient of a rectangle cross section and the reduction factors for
corner are given in figure 31 and 32. The reduction factor which accounts for the length of the
element , depends on the solidity ratio, same as NEN 13001, () and the aerodynamic slenderness
(). Figure 33 shows the graph of the , the end-effect factor.

A = area of the individual members of a lattice structure [m2]

Ac = enclosed area of a lattice structure [m2]

C = Constant from NEN-EN 1991-1-4 table 7.16 (see table

16)

53
Figure 31 NEN-EN 1991-1-4, Figure 7.23, aerodynamic coefficient

Figure 32 NEN-EN 1991-1-4, Figure 7.24, reduction factor corners

Figure 33 NEN-EN 1991-1-4, Figure 7.36, end-effect factor

54
Table 16 Effective slenderness [NEN-EN 1991-1-4, table 7.16]

For the crane hotel the resulting wind loads are calculated for wind area I and II and two terrain
categories. All are shown in table 17 on the next page, including all coefficients used to calculate.

55
Table 17 Wind load crane hotel, NEN-EN 1991-1-4, terrain category II and III

Solidity For all sections except A5 and A7b the solidity is 1. Section A5 has an enclosed area
of 77.3m divided by the projected area of 45.3m2 resulting in a solidity of 0.59. For A7b the solidity is
2

0.32.

Slenderness The slenderness of the members and elements of each section depends on the
dimensions, cross sections and within the structure. Table 16 shows how the slenderness is
determined.

Reduction factor Using figure 33 and the calculated solidity and slenderness the reduction
factor has been determined. No reduction has been applied for rounded corners.

56
Shielding factor The shielding factor is not described in the Eurocode so no reductions on the
areas have been made. For section A7a the projected area is doubled, because the top beam has a
depth of over 60meters. For section A5 and A7b an increase in the projected area has been made
using the shielding factors form NEN 13001-2.

Basic aerodynamic coefficients For the aerodynamic coefficients of sections A1, A2, A4 and
A6 a simplified boxed section has been used without rounded corners (see figure 31). Section A3 is
simplified to a sharp edged element (see NEN-EN 1991-1-4 chapter 7.7). Section A5, A7a and A7b are
simplified to a lattice structure (see figure 34).
A1: d/b=0.5 A2: d/b=0.7 A4 and A6: d/b=1

Wind pressure For the calculations of the out-of-service wind pressure a wind area I and terrain
catergory II and III have been used, see figure 10 and 11, B and C.

Safety factor For the safety factor the general safety factor from Eurocode 0 is 1,5. has been
used. See chapter 4.

Figure 34 Aerodynamic coefficient of a lattice structure [NEN-EN 1991-1-4 Figure 7.34]

57
3.4 Comparison of the standards, wind load
The standards display great diversity in wind loads; the differences are made by the amount of
parameters used to fine-tune the aerodynamic coefficient, the projected area and the applied wind
pressure.

The older standards (ISO 4302:1981 and NEN 2018:1983) provide one single table of aerodynamic
coefficients, which in both standards contains exactly the same data. The aerodynamic coefficient in
ISO 4302 and NEN 2018 is a function of the aerodynamic slenderness and the cross section of the
member. Both ISO 4302 and NEN 2018 allow the usage of aerodynamic coefficient obtained from
wind tunnel experiments.

The newer standards (NEN-EN 1991-1-4:2005 and NEN-EN 13001-2:2011) provide multiple tables and
a more extensive method to determine the aerodynamic coefficients. The aerodynamic coefficient in
NEN-EN 1991-1-4 and NEN-EN 13001-2 is a function of at least the aerodynamic slenderness, cross
section, aerodynamic length and the solidity ratio. For specific cross sections the aerodynamic
coefficient is also a function of the Reynolds number, surface roughness, and roundness of the
corners and orientation of the cross section to the wind direction.

The aerodynamic slenderness () is a constant which allows for the reducing resistance of the wind
flow on a structure caused by the end effects and the ground of the structure. In the newer standards
the aerodynamic slenderness is not the ratio between the length divided by the characteristic
diameter, but the aerodynamic length dived by the characteristic diameter. Both NEN-EN 1991-1-4
and NEN-EN 13001 use the same method and values to determine the aerodynamic length and the
resulting aerodynamic slenderness (see figure 20 and table 16).

The basic aerodynamic coefficient in the newer standards is spread over multiple tables and graphs
depending on the cross section, Reynolds number, surface roughness and orientation to the wind
direction. Especially the aerodynamic coefficient of round cross sections is highly dependable on the
Reynolds number, surface roughness and orientation. Both NEN-EN 1991-1-4 and NEN-EN 13001-2
provide similar basic aerodynamic coefficients, but NEN-EN 13001-2 contains the most extended
tables and graphs of aerodynamic coefficients. In the newer standards the aerodynamic coefficient is
also influenced by the solidity ratio (). The bigger the solidity ratio of a structure, the smaller the
resulting aerodynamic coefficient will be. Both NEN-EN 1991-1-4 and NEN-EN 13001-2 use the same
graph and values for determine the resulting reduction factor of the aerodynamic coefficient (see
figure 22 and 33).

( ) ( )

58
In all four standards the characteristic area, reference area, projected area or frontal area is defined
by sum of the member length and the characteristic diameter.

When a structure contains of multiple members parallel behind each other, the shielding of the
member influences the resulting wind load on the total structure. ISO 4302, NEN 2018 and NEN-EN
13001-2 all consider this effect, using a shielding factor (). The shielding factor is, in all three
standards, a function of the distance between the parallel members and the solidity ratio. The
shielding factor of a member can reduce the aerodynamic coefficient or characteristic area of the
considered member. ISO 4302, NEN 2018 and NEN-EN 13001-2 all use the same graph and values to
determine the shielding factor, see figure 24 and 29 and table 15.

The NEN-EN 1991-1-4 standard multiplies the wind load with a structural factor which takes the
structures size. NEN-EN 1991-1-4 calls it the size effect (cs). The effect a gust of wind has on a
structure depends on the scale of the gust and the dimensions of the structure. Large structure will
never be entirely enveloped by a single gust. Small structures or small parts of a structure can be
enveloped by a single gust, resulting in a higher and more fluctuating wind loads.

NEN 1991-1-4 finally also multiplies the wind load with the dynamic response (cd). In chapter 5 the
possible dynamic effects are described. The American ASCE 7 standard does not specify a factor for
dynamic response but states that: When dynamic effects occur, resulting in loads perpendicular to the
wind direction the use of wind tunnel testing is advised.

When comparing all formulas used in the methods in the four standards, the real difference is made
by the complexity of the determination of the wind pressure distribution and the aerodynamic
coefficient. The newer standards contain a more extensive library of data, for more situations and
cross sections types resulting in a more accurate wind load. But when using simple cross sections, the
resulting wind load will only differ because of difference in wind pressures, as shown in figure 15.

Table 19 shows an overview of all the resulting wind loads from all standards for the crane hotel.
Table 20 compares all maximum and minimum CdA, F and M values and compares the boom in
feathering positions with the boom perpendicular position.

59
Table 19 Wind load crane hotel, overview of all three standards

Table 20 Comparison of the maximum and minimum values between the parameters and crane
orientations.

60
The difference between NEN 13001, Eurocode 1 and NEN 2018 is partly created by the use of two
wind areas or terrain categories. The crane is positions in the centre of Amsterdam, but near the river
so there is a mix in wind areas and terrain categories, depending on the wind direction. Figure 35
shows the position of the crane.

Figure 35 Position of the crane hotel in Amsterdam, North-east of the river Het IJ (source: Google
Maps)

The maximum difference in CdA values is small (3-9%), compared to all other values. Changing the
position of the boom increases the CdA significantly with 25-32%.

The differences in wind load F are enormous, 43-50%. This is caused by the usage of multiple wind
areas and terrain categories. Changing the position of the boom increases the wind load with 34-
40%. The difference in wind load between both positions is more than 25-32% because the wind
pressure increases with height and the difference in CdA between both positions is made at an
average height of 44m.

The differences in wind moment M are of the same magnitude as the wind load F, 42-52%. The
difference between both boom positions has grown even more to 64-78%. The increased wind load
changes at the highest point and therefore strongly changed the moment M.

NEN 13001 and NEN 2018 are crane specific standard with safety factors for stability between 1.16-
1.4. Because the Eurocode is a standard for building, with humans, the safety factor is even higher,
1.5.

61
Table 21 Comparison of the maximum and minimum values between the parameters and crane
orientations. With all equal wind pressure and safety factor form: Eurocode 1, WA I and TC III.

Table 22 Comparison of the maximum and minimum values between the parameters and crane
orientations. With all equal wind pressure and safety factor form: Eurocode 1, WA II and TC II.

Table 20 provides the same CdA data as table 18 and 19, but with one wind pressure and safety
factor, form Eurocode 1: Wind Area II and Terrain Catergory II, Safety factor 1,5. Table 21 again
compares all maximum and minimum CdA, F and M values and compares the boom in feathering
positions with the boom perpendicular position.

It shows that for one position the differences between the CdA, F and M values have reduced
dramatically. But the difference between both feathering and perpendicular boom position remain.

62
4. Design wind load
In chapter three the procedure to determine the wind load is explained. But when designing a crane
the wind load is not the only load which acts on the cranes structure. Its always a combination of
loads.
In-service its at least: Own weight of the structure + hoisting load + accelerations
+ in-service wind load + etc.
Out-of-service its at least: Own weight of the structure + out-of-service wind load +
etc.

All loads are combined to getter in a load case, but not all loads have an equal contribution to the
total load.

NEN-EN 1990, Eurocode 0, defines load cases and multiplication factors for NEN-EN 1991-1-4. All
possible loads or actions are classified by their variation in time in three groups: Permanent actions
(G), Variable actions (Q) and Accidental actions (A). In-service wind loads can be permanent actions
but are mostly variable actions. Out-of-service wind loads are variable actions, but in case of
hurricanes it can very well be an accidental action.

The main representative of the wind load action (F) is the characteristic value of the wind load action.
This value is a constant when the variance is small. But with large variability the characteristic value
will contain a maximum upper value and a minimum lower value. For variable actions (Q) an upper
value is used with an intended probability of not being exceeded during a specific reference period.
For wind loads this is a 50 years return period.

When verifying a structure for ultimate limit states the product of the variable action (Q) and the
multiplication factor is used. Multiple ultimate limit states have to be investigated. For cranes loss of
static equilibrium (EQU) is very important.

The nation annex of the Eurocode 0 gives the following factors


factors for wind loads on buildings 0 = 0 1 = 0,2 2 = 0
Design values for EQU Q,1 = 1,5

So when checking a structure for its ultimate limit states the wind load is reduced by the factor. But
when checking the structure for stability the wind load in increased in the unfavorable direction.

NEN-EN 13001 is a standard specific for crane design so in this standard the load combinations are
even more focused on crane design than in the more general Eurocode standards and there for
described more detailed. In NEN-EN 13001 all possible loads are categorized as regular, occasional
(in-service wind loads) and exceptional (out-of-service wind loads) loads. In the load combinations
both the ultimate limit states and the stability have their own factors to be added to the loads.

63
NEN-EN 13001 gives the following factors for ultimate limit states
Occasional loads Partial safety factor p = 1,22 for in-service wind loads, load
combination B
Partial safety factor p = 1,16 for in-service wind loads, load
combination C
Exceptional loads Partial safety factor p = 1,16 for out-of-service wind loads, load
combination C
Load combination B Overall safety factor f = 1,34
Load combination C Overall safety factor f = 1,22

NEN-EN 13001 gives the following factors for rigid body stability
Occasional loads Safety factor = 1,16-1,22 for in-service wind loads, load
combination B
Safety factor = 1,0-1,16 for in-service wind loads, load combination
C
Exceptional loads Safety factor = 1,1-1,16 for out-of-service wind loads, load
combination C

NEN 2022 is a national standard from 1976 dealing only with crane stability. For non-mobile cranes in-
service wind and out-of-service storm wind are described. For in-service wind no factor is added to
the wind load. But for out-of-service storm winds a factor of 1,2-1,4 is multiplied with the wind load. If
the crane is equipped with stabilizers, which without the crane is also stable and they can be fitted
within 5 minutes, a factor of 1,2 may be used. If the crane is only stable with the stabilizers or there
are none a factor of 1,4 must be used.

64
5. Dynamic wind effects
This chapter deals with the dynamic effects which could occur under wind loading. Paragraph 5.1
gives an introduction, paragraph 5.2 briefly introduces the structural factor used in NEN-EN 1991-1-4
and paragraph 5.3 deals with the dynamic wind effects and vibrations, known in physics and
applicable for cranes.

5.1 Dynamic response


A structure is considered stiff if it responds directly to wind loads; the maximum strain is directly
proportional to the maximum load. In a dynamic situation the deflections are no longer directly related
to the load. This can generate resonances, possibly at the natural frequency of the structure. Dynamic
behaviour can happen for three reasons, which are explained in detail in 5.3.

1. Fluctuating nature of the wind velocity, Buffeting. When the frequency of the wind equals
the natural frequency of the structure and enough energy is absorbed, this may lead to
sufficient deflection.
2. When the structures shape generates an oscillatory flow pattern. Because the oscillation is
wind generated, its frequency is a function of the wind velocity. Examples are vortex
shedding.
3. When the motion of the structure increases the wind load its considered Aero-elastic. The
oscillating force is generated by the movement of the structure, which usually occurs at
the natural frequency of the structure. Examples are galloping and flutter.

Figure 36 shows the types of motions which belong to buffeting, vortex shedding, galloping and
flutter. Flutter contains rotational movements but can also be a coupled movement of both translation
and rotation.

Figure 36 Motions of dynamic wind effects

65
A dynamic wind effect can generate extreme forces on a structure but can also generated moderate
forces, but so frequent that these force become important in fatigue. Collapse can very well occur due
to (wind load) fatigue and not overload.

When considering the dynamic response of a structure the natural frequencies and natural mode
shapes need to be known. This can be done using computer finite element analyses of the structures
design.

5.2 Structural factor, NEN-EN 1991-1-4


The NEN-EN 1991-1-4 standard takes dynamic effects into account. The wind load is increased with
one combined constant. The structural factor cscd accounts for the combined effect of the non-
simultaneous action of peak wind pressures over faces of the structure (size effect = cs) and the
vibration of the structure in its fundamental mode due to the action of turbulence (dynamic response
= cd).

The structural factor

( )
kp = peak factor, ratio of the maximum value of the
( )

fluctuating part of the response to its standard deviation


Iv = turbulence intensity
zs = reference height can be found in NEN-EN 1991-1-4:2005
figure 6.1 (see figure 37)
B2 = background factor, allows for the lack of correlation of
pressure over the structure
R2 = resonant response, allows for turbulence in resonance
with the structure

The structural factor may only be calculated with the formula above if the structure corresponds to
one of the general shapes shown in figure 37; Vertical structures, horizontal structures and point like
structures. Also only the along-wind vibration in the fundamental mode must be significant, and this
mode shape needs to have a constant sign.

The national annex NEN-EN 1991-1-4+A1+C2/NB states that NEN-EN 1991-1-2:2005 Annex C must
be used for calculating the structural factor, background factor and resonant response. The applied
structural factor may also not be smaller than cscd 0,85.

66
Figure 37 General shapes of structures covered by the design procedure [NEN-EN
1991-1-4 Figure 6.1]

67
5.3 Wind effects and vibrations
This paragraph deals with: Buffeting, Vortex shedding, Galloping, Divergence and flutter.

5.3.1 Buffeting
With Buffeting the fluctuating nature of the wind velocity generates an equally fluctuating wind
pressure. This fluctuating wind pressure may cause the structure to deflect in the along wind
direction. When this motion reaches the natural frequency of the structure an energy build-up can
lead to severe deflection. When the direction of motion of the structure ( ) is equal to the wind
direction ( ) than the relative velocity with respect to the structure ( ) is reduced and therefor
the wind pressure is reduced. This phenomenon is known as positive aerodynamic damping. When the
motion of the structure is opposing the wind direction the relative velocity ( ) is increased and
therefor also the wind pressure. This phenomenon is known as negative aerodynamic damping. Figure
38 shows an example of aerodynamic damping.

Figure 38 Wind velocity (U) on a deflecting structure (x)

How to avoid Buffeting? When a structure is stiff the amount of deflection will be minimal and the
possible negative aerodynamic damping will be small. When a structure or members of it consists of
large frontal areas it will not be swept by a single gust but always by multiple gusts, reducing the
chance of negative aerodynamic damping. In the Eurocode the size effect (cs) is a factor which looks
at the size of the structure and the resulting possibility of being swept by a single gust.

5.3.2 Vortex shedding


Vortex shedding is an unsteady oscillating flow of vortices which can appear downstream a structure
placed in a laminar flow. Alternating low pressure zones downstream of the structure generates a
force on the structure perpendicular to the wind flow. Vortex shedding may result in structural
vibrations if the frequency of vortex-shedding is equal to the natural frequency of the structure. The
generated force can result in the structure vibrating and in fatigue of the structure. Typically, the
critical wind velocity for vortex shedding is a frequent wind velocity. The critical wind velocity is
reached when the frequency of vortex shedding equals to the natural frequency of the structure.

The repeating pattern in which the vortex shedding occurs is called a Karman vortex street. Shown in
figure 39.

68
Figure 39 Simulated vortex street around a no-slip cylindrical obstruction

When the vortex shedding frequency changes towards the natural frequency of a structure a
phenomenon known as lock-in can occur. The strength of the vortices shed, and the fluctuating forces
are enhanced as energy is build up.

Vortex shedding can create enormous instability. It caused the collapse of three cooling towers at the
Ferrybridge power station in Great Brittan in 1965. Shown in figure 40.

Figure 40 Collapsed cooling tower of the Ferybridge power station, 1965

When the ratio of the largest and smallest crosswind dimension of a structure, both taken in the plane
perpendicular to the wind, exceeds a factor 6, vortex shedding should be investigated.

Vortex shedding need to be investigated, according to NEN-EN 1991-1-4, Annex E.1

vm = mean 10min. wind velocity


Vcrit,I = critical wind velocity, for vibration mode i.

69
The critical wind velocity is defined a function of the natural frequency of the structure and the
Strouhal number. The Strouhal number is a dimensionless number describing oscillating flow
mechanisms. It depends on the cross-section shape and on the Reynolds number. NEN-EN 1991-1-4,
Table E.1 and figure E.1 present the Strouhal numbers for different cross-sections. Figure 41 shows
the Strouhal number for a boxed cross-section.

Figure 41 NEN-EN 1991-1-4, Figure E.1, Strouhal number for rectangular cross-sections
with sharp corners

The critical wind velocity, for bending vibration mode i.

b = the reference with of the cross-section at which vortex

shedding occurs
ni,y = the natural frequency of the considered flexural mode i
of cross-wind vibration
St = Strouhal number

The critical wind velocity for vortex shedding depends strongly on the size of the structure or its
member and the stiffness. The Strouhal number is always in the range of 0,06 to 0,18. For the crane
hotel vcrit,I > 37m/s resulting in a minimal natural frequency of 4,5Hz for a square boxed section of
1x1m.

The extent in which vibrations caused by the wind flow play a part, depends on the ratio between the
structural mass per unit length times the structural damping and the fluid mass times the square of
the reference with. This is expressed with the Scruton number (Sc). Higher Scruton number result in
smaller displacements of the vibrations. For circular cross sections a Sc > 10 usually indicates a low
amplitude of vibration.

70
In the paper of (Hansen, 2007) the flow and structural characteristics governing vortex-induced
vibrations of structures are described. Results obtained from wind tunnel tests, with different cross-
sections, are compared with the mathematical models presented in NEN-EN 1991-1-4, Annex E.1.
Hansen states that although a great deal of effort has been made in improving analytical models for
predicting vibrations due to vortex shedding, these models are still rather crude. Models should focus
on rare and extreme events, occurring once in 10-50years and generating large amplitudes but also
on frequent events, occurring very often during the expected lifetime of a structure and generating
small amplitudes. Frequent events give a major contribution to the fatigue damage of a structure.

The first approach in NEN-EN 1991-1-4, Annex E.1.5.2 is based on the vortex-resonance model and
provides response estimates, which are larger than frequent and lower than rare events. The second
approach in NEN-EN 1991-1-4, Annex E.1.5.3 is based on the spectral model and takes rare as well as
frequent events into account by including the influence of turbulence in the vibration amplitudes
predicted.

Large-scale turbulence in the atmosphere may be interpreted as a slowly varying mean wind velocity.
Hansen report shows that when the mean wind velocity is 4-5% higher or lower than the critical wind
velocity the generated amplitude of the vibration is reduced by more than 50%, see figure 42. So
large scale turbulence and the resulting high wind velocity variation will have a reducing effect on the
vortex-induced vibrations.

Figure 42 Amplitude built-up after the cylinder has been released in low turbulence flow
(Hansen 2007)

In certain meteorological situations with cold and smooth air flow over a relatively long period of time,
say of 1 hour, some slender steel structures may experience larger vibrations than predicted by the
vortex-resonance model used in the first approach in NEN-EN 1991-1-4, Annex E.1.5.2. This drawback
of the first approach is not presented in the second approach: the spectral model NEN-EN 1991-1-4,
Annex E.1.5.3.

71
How to avoid vortex shedding? Vortex-induced vibrations and movements can be reduced by
sufficient structural damping and by avoiding cross sections which a prone to vortex shedding. Mostly
sections with high Strouhal numbers like: square boxed sections and smooth circular sections. But
when using such cross sections aerodynamic devices can be added to reduce the chance of vortex
shedding. Like guide vanes, splitters, adding protuberances, helical stakes or damping devices. Figure
43 shows splitters and spoilers. Figure 44 shows helical stakes. Helical stakes are fitted on circular
cross sections and produce turbulence with different length scales and hence a less constructive force
on the structure. Figure 45 shows protuberance on multiple structures. Protuberances generate the
same effect as helical stakes. The drawback of helical stakes and protuberances is that they increase
the drag coefficients and drag force and possibly increase buffeting.

Figure 43 Splitter and spoiler to reduce vortex shedding

Figure 44 Helical stakes. Reducing the vortex shedding of circular cross sections

Figure 45 Protuberances. Reducing the vortex shedding of circular cross sections

72
The paper of (T. Zhou, 2011) describes a research on cylinders with and without helical strakes. It is
found that for cylinders with a diameter d=80mm helical strakes can reduce vortex induced vibrations
with 98% and no lock-in is experienced.

The paper of Hansen also states that high turbulence can create high fluctuating wind velocities and
therefor reduce the change of a constant wind at the critical velocity. So turbulence can reduce the
chance on vortex shedding. So downstream members in a lattice structure are less sensitive to vortex
shedding. Unless the distance between the members is identical to the length of the vortexes. In that
case the downstream member will increase the already present vortexes.

5.3.3 Galloping
Galloping is a self-excited vibration and aero-elastic phenomena. Its a single degree of freedom
motion and occurs when there are large amplitude oscillations normal to the mean wind flow, at low
frequencies. Galloping arises when the aerodynamic damping is negative and so becomes a forcing
function. Aerodynamic damping is a form of wind resistance generated by the movement of the
structure itself. Its negative if the structure moves opposite of the wind flow, increasing the wind
load. The structure extracts kinetic energy form the wind and transferring it into motion of the
structure itself. Galloping needs to be initiated by a velocity of the structure perpendicular to the
velocity of the wind.

Lightweight and flexible structures are most likely to undergo galloping. Cross sections which are
prone to Galloping are: square sections, D-shaped sections and iced-up lines or guy cable. Galloping
always occurs at the structures natural frequency. The tendency and intensity of structure galloping
increases when the wind velocity increases. Ice may also cause stable cross sections to become
unstable.

The wind velocity at which galloping starts is the onset velocity of galloping (vCG). NEN-EN 1991-1-4,
Annex E.2.2

Sc = Scruton number

aG = galloping instability factor


ni,y = the natural frequency of the considered flexural mode i
of cross-wind vibration
b = the reference with of the cross-section at which vortex
shedding occurs

One should ensure that the onset wind velocity of galloping is at least 25% higher than the mean
wind velocity.

73
When the ratio between the onset wind velocity of galloping and the critical wind velocity for vortex
shedding is within 0,7-1,5 expert advice is recommended.

The cross section of a member determines the galloping instability factor. Table 22 shows such a
table.

Table 23 Factor of galloping instability a0 [NEN-EN 1991-1-4, Table E.7]

When multiple cylinders are coupled classical galloping may occur. The wind velocity at which classical
galloping is equal to the formula for the wind velocity at which galloping starts. When multiple
cylinders are arranged close together, without being connected, interference galloping may occur. A
front member in the structure may generate wakes which can cause galloping to the following
members. The wind velocity at which interference galloping starts is given in NEN-EN 1991-1-4, Annex
E.3
74
aIG = the combined galloping instability factor aIG = 3,0

a = the spacing between the cylinders

Interference galloping can be avoided by coupling the free-standing cylinders. In that case classical
galloping may occur. For multiple cylinders the dimensions, spacing and Scruton number determines
the difference between the offset velocity for coupled and interference galloping. Figure 46 shows an
example of two cylinders including the dimensions, stiffness, Scruton numbers and the instability
factors for coupled and interference galloping.

Figure 46 Example of coupled and interference galloping

Above example shows that for structures with low Scruton numbers (light and not much internal
damping) coupling increases the offset velocity with just 20% but for structures with high Scruton
number (heavy and high internal damping) coupling increases the offset velocity with 170%.

The thesis of (Larsen, 2011) deals with vortex shedding and galloping. His research focused on a
rectangular cross section with a height/depth ratio of 1:2. This specific cross section starts galloping
wind velocity 70% lower than NEN-EN 1991-1-4 predicts. So if a structure is critical for galloping a
wind tunnel test is advised.

How to avoid galloping? Just like with vortex shedding the Scruton number and the natural
frequency strongly contribute to the offset wind velocity for galloping. So a high Scruton number and
a high natural frequency increase the offset velocity. So a high internal damping, mass, natural
frequency and with of the structure have a positive contributions.

Interference galloping can be avoided by coupling the multiple cylinders. But the example above
shows that not all structures will strongly benefit form coupling.

For lines and wires galloping cannot be completely prevented because of the small internal damping
of cables and lines, but galloping can be reduced. This can be done by coupling multiple lines together
by mechanical devices. Figure 47 shows multiple mechanical devices.

75
Figure 47 Mechanical coupling of wires, to reduce galloping

5.3.4 Divergence and flutter


Divergence and flutter are instabilities that occur for flexible plate-like structures above a certain
critical wind velocity which involve rotations of the structure. Its an oscillation of two or more degrees
of freedom. Plate-like structures are boxed section of which the ratio between the characteristic
diameter divided by the wind depth is smaller than twenty-five per cent.

Divergence and flutter are caused by the deflection of the structure. This deflection changes the wind
load on the structure, which again changes the defection. Depending on the stiffness and damping of
the structure this can be a stable or unstable deflection/motion. Divergence and flutter should be
avoided. The Eurocode 1 also states in Annex E.4 that the critical divergence velocity should be at
least twice the mean wind velocity.

Divergence is a steady aero-elasticity which can grow exponentially. Where the change in
aerodynamics and loads increases the forces resulting in a positive feedback process, bringing the
structure back to its equilibrium. But in some cases the increased loads can deflect the structure to its
limits and failure.

Flutter is a dynamic aero-elasticity, containing two dimensional body rotations causing an oscillating
motion. The change in aerodynamics and loads increases the forces, resulting in vibration. When
vibrating at a structures natural mode it can result to failure. When the damping of the structure is
insufficient an energy build-up in the system results in increasing amplitudes of the vibration and a
self-exciting oscillation. Flutter lead to the destruction of the original Tacoma Narrows Bridge.

76
Figure 48 shows the difference between divergence and flutter. The numbers provide the order in
which the photos have been taken. Divergence is a rotational phenomenon in which a structure is
rotated until the lack of stiffness and damping leads to failure. Flutter is an oscillating motion which
also leads to failure when the structure lacks stiffness and damping.

Figure 48 The oscillating motion of flutter and the rotational motion of divergence

The Tacoma Narrow Bridge collapsed on November 7, 1940 under a wind velocity of just 68km/h. A
torsional vibration mode, the second torsional mode of the bridge, twisted the roadway to such
excessive deflections and stresses it collapsed, see figure 49. The critical flutter wind velocity in this
case was definitely not higher than the maximum mean wind velocity. Vortex shedding did not cause
the collapse of the bridge, the frequency of the destructive mode was 0,2Hz and this was not a
natural mode of the bridge.

Figure 49 Tacoma Bridge November 7th, 1940

77
Braniff Flight 542, a Lockheed L-188 Electra, disintegrated in mid-air resulting in a fatal crash on
September 29, 1959. The left wing experienced flutter. When the magnitude of the flutter grew the
frequencies decreased to 3Hz creating harmonic coupling. This caused even lager wing vibrations
resulting in the left wing disintegrating from the airplane. Northwest Orient Airlines Flight 710, also a
Lockheed L-188 Electra, crashed on March 17, 1960 by the same cause as Braniff Flight 542.

How to avoid flutter? Flutter and divergence are rotational deflections so the number one step is to
avoid flutter is to provide the structure with enough rotational stiffness. So that the structure cannot
be rotated to such a point that the wind load keeps increasing until failure.

78
6. Wind tunnel research
Aerodynamic coefficients or wind loads may also be determined with wind tunnel experiments. NEN-
EN 1991-1-4 and ASCE 7 advices to use wind tunnel testing when their analytical method is
insufficient. This can be the case when dynamic effects are expected to occur and not adequately
covered in the standard, or when the structure has an unusual shape.

When using a wind tunnel one must always use a scale model of the structure, or parts of the
structure. The wind flow inside the wind tunnel must always represent a real life wind flow, so must
also contain a boundary layer. The forces, vibrations, pressures etc. on the model structure must be
measured using sensors and a setup. The setup itself determines which parameters and which
behaviour can be measured.

Wind tunnel research can also as a tool for determine the stability of a crane. This could lead to
adding extra stowing devices, like rail clamps ore tying down the crane. In September 2003 super
typhoon Maemi destroyed multiple container cranes in South Korea. Figure 50 shows the demolished
cranes.

Figure 50 The demolished cranes in South Korea, after super typhoon Maemi in 2003

79
6.1 Scale model
Because cranes are often very large structures one must use a scale model for wind tunnel testing.
Figure 51 shows an example. To make sure the scale model and the results are realistic and
applicable for the full scale structure some rules must be checked.

Figure 51 Example of a scaled crane model inside a wind tunnel

The flow around a cross section is determined by the Reynolds number. The flow around a scaled and
full scale cross section is identical when both have the same Reynolds number.

v = wind velocity

D = characteristic diameter
= kinematic viscosity

With a scale model the characteristic diameter of all members is reduced. So when working with a
scale 1:50 model the wind velocity should be increased 50 times compared to the full scale mean wind
velocity. This is impossible and not necessary. The aerodynamic coefficient of many cross sections is a
function of the Reynolds number and so is the aerodynamic coefficient. But the aerodynamic
coefficient is constant for certain Reynolds number ranges. So its not needed to scale up the wind
velocity with the same factor as the dimensions are reduced, but one should only check that the
Reynolds number generated by the wind tunnel flow is the same range as the full scale model.

Besides the Reynolds number there are many other parameters which should correspond between the
full scale and scale model, especially when testing dynamic behaviour, values like the stiffness and
damping should correspond. These parameters are known as non-dimensional response/pressure
coefficients. They are all associated with the flow and the structure. Examples are
Turbulence intensity, the ratio between inside the wind tunnel and in the real world.
Cauchy number, a ratio between the elastic forces and the inertial forces.
Density, the ratio between the density of the structure and the air density.
Jensen number, ratio between the length scale factors of the structure and the atmospheric
boundary-layer simulation.
Froude number, ratio between the wind velocity and the wave propagation velocity.
Strouhal number, ratio between the frequency of vortex shedding and the wind velocity.
80
When making a scale model crane, not always all details are scaled. Minute structures, like stairs,
cables, railings etc. my not be included in the scale model. The report of (Sang-Joon Lee, 2008)
shows that drag forces and pitching moments coefficients of a scale model without minute structures
can be decreased with 8-20% and 12-33%, respectively, compared with detailed crane models which
include minute structures.

6.2 Wind flow


A correctly constructed and dimensioned model and a well-chosen wind velocity is still not enough to
ensure a proper result. Not only the velocity but also the turbulence intensity and boundary layer must
be generated properly.

When determine the wind forces and dynamic response the turbulence level and boundary layer of
the wind flow are important. A change in both can result in a change in de measured forces and
response. The report of (A. Scarabino, 2005) shows the result on drag coefficients in a smooth flow
and in a turbulent flow. Wind tunnels can produce smooth flow, the results form wind tunnel
experiments are therefore not always representative of the real wind loads on structures. In the real
world turbulence can be generated by surrounding structures ore landscape. Turbulent flows can
reduce the drag coefficients. Making drag coefficients determined in smooth flow higher and more
conservative.

A boundary layer can be created by adding surface roughness to the floor of the upstream part of the
wind tunnel, shown in figure 52 and contains turbulence. The length of the upstream part determines
the developed height of the boundary layer. The developed height should exceed the height of the
model and envelop the model completely. One real life effect inside the boundary layer cannot be
achieved in wind tunnel testing: the Ekman Spiral. Always ensure that the blockage ratio is below 5%,
to prevent significant increases or decreases of the wind velocity around the model and wind
pressures on the model. In the report of (Jong-Hoon Kang, 2008) the difference in wind load between
smooth flow and two types of atmospheric boundary layers is shown. Is shows that the wind load is
reduced in ABL compared to smooth flow, but the overturn moment is increased. The report also
points out that the aerodynamic force coefficients have nearly constant values, when the Reynolds
number is beyond Re = 5x105.

Figure 52 Boundary layer creation in a wind tunnel, Holmes (2001)

81
6.3 Support system
When working with a scale model of a structure there are three options for the support system of the
model. Figure 53 shows illustrations of them.

1. Rigid model on which multiple surface pressure tapes are placed. The wind pressure
distribution over the model can be determined.
2. Force balance model which is connected to a balance. The overall forces and
moments can be determined
3. Aero-elastic model which is dynamically coupled to the support. This allows
determining the aerodynamic forces and the dynamic response.

Figure 53 Support system for wind tunnel testing with scale models. left: pressure taps, center:
force balance, right: dynamically coupled.

The support system of the model structure determines the possibility to measure the aero elastic
forces and the resonant response of the model structure. A base-pivoted support system allows for
the model structure to vibrate. A base balance is a stiff support system, where the aero elastic
properties of the real structure are not modelled. A base balance support system can reduce the
amount of wind tunnel testing time by a large factor. When working with an aero-elastic model the
behaviour of the model should agree with the behaviour of the real structure.

82
7. Numeric example
Chapter 7 shows a numeric example of the in the Netherlands applicable standards for a simplified
crane. Paragraph 7.1 described the crane, paragraph 7.2 deals with the standards and paragraph 7.3
compares the results. Paragraph 7.4 and 7.5 presents the result of a paper in which results from wind
tunnel measurements are compared to results from NEN 13001 and NEN 2018.

7.1 Numeric example crane


Figure 54 shows the container which has been simplified and used in this example, shown in figure 55
is the frontal view.

Figure 54 Container crane, numeric example Figure 55 Simplified container crane, numeric
example

The U-shape in figure 55 consists of three members. Two vertical members (I and II) of: 2mx2mx25m
and one horizontal member (III) of: 2mx3mx30m.
d1 = 2,00m l1 = 30,00m d2 = 3,00m l2 = 25,00m depth = 2,00m

In this chapter just one U-shape is considered and a constant wind pressure of q = 500N/m2.

83
7.2 Results numeric example
This paragraph gives the results using the crane described in paragraph 7.1.

7.2.1 NEN-EN 13001-2:2011


The relative aerodynamic length of both members is found in NEN-EN 13001-2 Table A.1. (see figure
20)
= 1,85 with Table A.1 # 4 using the formula:
lo = 25,00m 1,50m = 23,50m and d = 2,00m
= 1,78 with Table A.1 # 7 using the formula:
lo = 30,00m 2,00m = 28,00m and d = 3,00m

The solidity ratio for all members

The aerodynamic slenderness of both members

Figure 56 shows the determination of the reduction factors for the members
I and II = 0,78
III = 0,76

Figure 28 shows a boxed cross section member with sharp edges.


c0, I and II = 2,0 c0, III = 2,13 (linear interpolation between b/d=1 and b/d=0,5)
2
AI and II = 47,0m AIII = 84,0m2

The wind load on the simplified crane


( )

The simplified crane from figure 55 consists in reality of two U-shapes, who are 20,0 meters behind
each other. Annex A.4 of NEN-EN 13001-2 deals with multiple members behind each other. The
shielding factor () depends on the solidarity and the ratio a/d, where a = distance between the
members, d = characteristic dimension. (see figure 24)
= 0,999 for and

84
The wind load on two U-shapes
( )

Figure 56 Reduction factors of both member of the example crane, NEN-EN 13001-2 Figure A.1

85
7.2.2 NEN 2018:1983
The aerodynamic coefficients of both member types are determined using NEN2018 table 22, shown
in table 23.

Member I and II h/d = 1 l/b = 11,75 Cf = 1,59 A = 47,0m2


III h/d = 1,5 l/b = 9,33 Cf = 1,63 A = 84,0m2

Wind load on the simplified, single U-shape crane

The shielding factor for the second U-shape frame, at a distance of 20,0 meters = 1.

The wind load on two U-shapes

Table 23 NEN 2018, table 22, force coefficients, example crane

86
7.2.3 ISO 4302:1981
For the force coefficients of the types of members linear interpolation is necessary. Shown in table 24.

Member I and II b/d = 1 l/b = 11,75 Cf = 1,59 A = 47,0m2


III b/d = 1,5 l/b = 9,33 Cf = 1,63 A = 84,0m2

Wind load on the simplified, single U-shape crane

The shielding factor for the second U-shape frame, at a distance of 20,0 meters = 1.

The wind load on two U-shapes

Table 24 ISO 4302:1981, table 2, force coefficients, example crane

87
7.2.4 NEN-EN 1991-1-4:2005
The direct method is used for this example. Figure 57 and 58 show the graphs used to determine the
coefficients.

Member
I and II b/d = 1 cf,0 = 2,1 r =1 = 21,7 =0,78 A = 47,0m2
III b/d = 1,5 cf,0 = 1,87 r =1 = 26,6 =0,81 A = 84,0m2

The structural factor cscd of this simplified crane is not calculated. A value of 1 is used.

[ ( ) ]

[ [( ) ] [( ) ]]

Figure 57 NEN-EN 1991-1-4, Figure 7.23, force coefficient, example crane

Figure 58 NEN-EN 1991-1-4, Figure 7.36, end-effect factor, example crane

88
7.3 Comparison of the results
The example crane is this example consists of a simple and common cross section. All standards
provide aerodynamic coefficients for this cross section. All dough the newer standards use a more
intense method the results are for all standards identical. This is mainly caused by the simple cross
section. The newer standards contain an extensive library of cross sections with adequate
aerodynamic coefficients, which are often a function of the shape ratio and Reynolds numbers.

NEN-EN 1991-1-4 does not contain a shielding factor, which is necessary for crane design. Using no
reduction caused by the shielding effect leads to a strongly overestimated wind load.

89
7.4 Wind tunnel vs. NEN 13001 vs. NEN 2018
In the papers (Bos, Wind loading on crane structures, 2002) and (Bos, Calculations Wind loading on
Crane Structures, 2003) the results of a wind tunnel test on a scales lattice structure jib (see figure
59) have been compared with calculations on the same jib using NEN 2018 and NEN 13001.

Figure 59 Wind tunnel tests on a model crane structure, J.F. Eden 1985

Figure 60 shows all dimensions of the jib sections which have been tested in the wind tunnel. Two
types of sections have been tested: the typical jib section and the high solidity jib section. Both at a
scale of 1:12.

Figure 60 Dimension of the jib section tested in the wind tunnel (J.F. Eden 1985)

From both jib sections the aerodynamic coefficients and projected area have been calculated with NEN
2018 and NEN 13001-2. Table 25 shows the calculation of the typical jib section for NEN 13001-2. For
the calculations the dimensions of the scale model have been used. This has been done because the
real dimension will lead to higher Reynolds numbers. For NEN 13001-2 this can lead to higher
aerodynamic coefficients for curtain cross sections.

90
Table 25 Calculation of AC NEN 13001-2, typical jib section

Table 26 shows all AC values, wind velocity and the resulting wind load. The peak wind load from the
wind tunnel experiment has also been included.
Table 26 Overview of the resulting wind load on the jib sections

The differences between the standard are minimal in this example. But infect there is a difference
between the newer NEN 13001 and the older NEN 2018.

91
7.5 NEN 13001 vs. NEN 2018
In the papers (Bos, Wind loading on crane structures, 2002) and (Bos, Calculations Wind loading on
Crane Structures, 2003) not only wind tunnel tests have been compared. But also the wind load on a
mobile crane and a harbor crane both with a solidity of = 33% or = 51%. Both cranes in the
wind condition: in- and out-of-service. In this comparison the differences between both standards
appear. The aerodynamic coefficients in NEN 13001 come from an extended and more detailed library
than found in NEN 2018 this causes a difference for higher Reynolds numbers. Another difference
between both standards is found in the wind pressure. In-service both standards use similar wind
pressure but out-of-service NEN 13001 uses much higher wind pressures, as shown in figure 6, 7 and
15. Table 27 and 28 show the wind pressure, effective aerodynamic coefficient of the total crane and
the wind load.

Table 27 Overview of total calculated wind force according to both standards for a typical jib
section ( = 33%)

Table 28 Overview of total calculated wind force according to both standards for a high solidity
section ( = 51%)

In table 27 and 28 it shows that for low wind pressures and Reynolds numbers the differences are
small. But for the out-of-service wind condition the differences are much larger. The wind pressure
from NEN 13001 is >16% higher in all cases. Because the high Reynolds numbers reduce the effective
aerodynamic coefficients of the harbor crane the wind load differs between -8% to +16%.

So generally results of both standards (7.4 + 7.5) show good similarity with the wind tunnel results
and with each other for low Reynolds numbers. Differences between the standards occur at high
Reynolds-numbers. NEN-EN 13001 has an extended shape library of aerodynamic coefficient
compared with NEN 2018, which results in better aerodynamic coefficients. In the out-of-service
condition the difference in wind pressure between the standards is 20-80%, depending on the wind
area chosen and the height. This enormous difference is due to the increase in knowledge in the
almost 30 years between both standards.

92
8. Conclusion
When designing a crane the use of standards is obligatory. The standards describe all occurring loads
and load combinations resulting in a safe and durable construction. The wind load is one of the loads
a structure must withstand. In all standards the wind load is calculated using the basic physic formula,
shown in paragraph 1.4.

All standards used in this report show a similar method used to determine the wind load. The
difference is made in the modelling of the wind pressure out of the basic wind velocity and the
turbulence intensity. The method of determine the aerodynamic coefficients and projected area differs
in the amount of used parameters and detail.

How is the wind pressure determined and modelled?


The wind pressure in all standards is modelled as a static wind pressure which represents the pressure
of a dynamic and constantly changing wind flow. Distinction is made in some standards between an
in-service wind load and an out-of-service wind load. The out-of-service wind load increases with
height. The bases of the wind pressure is the fundamental wind velocity, which is measured at 10m
height above flat and open ground and averaged over a certain time, 3 seconds or 10minutes and
comes with a certain return period, often 50 years. This fundamental wind velocity depends on the
topographic location of the structure and the presence of nearby structures and landscape. In the
newer standards the resulting basic wind velocity is increased with a factor which adds the dynamics
of a turbulent wind flow. Finally the mean wind velocity is multiplied with the air density resulting in
the wind pressure, as a function of the structures height.

( )

( ) ( )

( ) ( )

How are the shape coefficients of the structures member determined?


The shape coefficients of the structure determine the aerodynamic drag they create. In the older
standards these constants where very rough estimates and all cross section and members where dealt
with in a single table of aerodynamic coefficients. In the newer standards this has changed
enormously. For a certain cross section a basic aerodynamic coefficient is given, for some cross
sections this basic coefficient is even a function of the Reynolds number or wind angle. Their position
within the structure determines the possibility of reducing this basic aerodynamic coefficient.
Depending on the slenderness, solidity and roundness of the corners.

( ) ( )

All standard allow wind tunnel testing or CFD as well, to determine aerodynamic drag coefficients.

93
How is the wind load on a crane determined?
The final wind load on a crane is the multiplication of the wind pressure with the aerodynamic
coefficient and the projected area. The projected area is the frontal area which is hit by the wind
flow. When multiple members of a structure are downstream of each other, they can provide shielding
from the wind. This shielding effect reduces the wind pressure. They amount of shielding depends on
the ratio of the spacing between the members and the height of the members.

( ) ( )

For design calculations on a crane the wind load is not the only load. Its combined with the own
weight, hoisting load, accelerations etc. Each load has its contributions to the total load situation.
When looking at ultimate limit states the contribution of the wind load is reduced by a safety factor
below one. But when looking at rigid body stability the influence of the wind load is increased with a
safety factor above one.

( )

Which dynamical effects are possible to occur with large structures?


The possible dynamical effects are not intensively dealt with in all standards. The occurrence of
dynamical effects mainly depends on the structures dimensions, stiffness and natural frequencies. The
cross sections dimensions can strongly determine the sensitivity of a structure to start behaving
dynamically. Cranes commonly have a natural frequency below 1Hz and contain slender steel
members. Therefore cranes can be sensitive to dynamical behaviour, especially buffeting and vortex
shedding. Avoiding critical cross sections and adding spoilers, helical stakes, coupling etc. can strongly
increase the offset wind velocity for dynamical behaviour. Wind tunnel research with an aero-elastic
model is a useful and advised tool to investigate dynamical behaviour.

Wind loading standards are often based on extensive research, but they remain simplified models of
wind loading. So great accuracy cannot be expected. Wind tunnel and CFD research can be a valuable
tool for determine the wind load and dynamic behaviour of a crane model. When a crane consists of a
complex structure and/or is surrounded by many structures causing aerodynamic interference
between the members standards can result in an extremely conservative wind load, so wind tunnel
testing can be an option. Make sure that when using a wind tunnel test the model, the wind flow and
the setup are constructed correctly and all dimensionless scaling parameters are checked.

All standards look alike, but still differ. The growth of world trade will force the number of standards
to reduce. More consistency in standards will develop more detailed and worldwide applicable
standards. Because each part of the world has its own climate and resulting wind velocity,
temperature, air density etc. a worldwide standard can contain much more data and detail in

94
determine the wind pressure. The knowledge on wind dynamics and the resulting wind loads is a
constantly developing field, in which wind tunnel and CFD testing provides more and more insight. A
worldwide standard can increase and accelerate the knowledge on wind loads on structures. For
manufactures a worldwide standard would mean that a crane design can be sold all over the world,
for a certain maximum wind level. This can reduce the cost of recalculation the crane design for each
countys own national standard.

95
References
(sd). Opgehaald van Universit de Liege: http://www.tdee.ulg.ac.be/doc-30.html
A. Scarabino, J. M. (2005). Drag coefficients and Strohal numbers of a port crane boom girder section.
Journal of Wind Engineering and Industrial Aerodynamics 93, 451-460.
ASCE, T. C.-I. (2011). Wind Loads for Petrochemical and Other Industrial Facilities. Reston, Virginia:
American Society of Civil Engineers.
Bos, I. W. (2002). Wind loading on crane structures. Delft: Faculty Mechanical Engineering and Marine
Technology, section Transport Technology and Logistics, TU Delft.
Bos, I. W. (2003). Calculations Wind loading on Crane Structures. Delft: Faculty Mechanical
Engineering and Marine Technology, Section Transport Technology and Logistics, TU Delft.
Bouw, N. (2011). Bouw Jaarboek 2011. Delft: Nederlands Normalisatie-instituut.
CEN. (2005, April). Eurocode 1: Actions on structures - Part 1-4: General actions - WInd actions [NEN-
EN 1991-1-4:2005].
Cook, N. (2007). Designers' Guide to EN 1991-1-4 Eurocode 1: Actions on structures, general actions
part 1-4. Wind actions. London: Thomas Telford Publishing.
Engineers, A. S. (2010). Minimum Design Loads for Buildings and Other Structures [ASCE/SEI 7-10].
Reston, Virginia, USA: American Society of Civil Engineers.
Han, D.-S. H.-J. (2011). The difference in the uplift force at each support point of a container crane
between FSI analysis and a wind tunnel test. Journal of Mechanical Science and Technology
25 (2), 301-308.
Hansen, S. O. (2007). Vortex-induced vibrations of structures. Bangalore, India: Structural Engineers
World Congress.
Holmes, J. D. (2001). Wind Loading of Structures. London: Spon Press.
J.F. Eden, A. B. (June 1985). A new approach to the calculations of wind forces on a vertical and
horzontal latticed structures. In The structural Engineer Volume 63A.
J.Verschoof, I. (1999). Cranes. John Wiley and Sons Ltd.
Jenkinson, A. (1955). The frequency distribution of the annual maximum (or minimum) values of
metrological elements. Quarterly Journal of the Royal Metrological Society 81, 158-171.
Jong-Hoon Kang, S.-J. L. (2008). Experimental study of wind load on a container crane located in a
uniform flow and atmospheric boundary layers. Engineering Structures (30), 1913-1921.
Larsen, M. (2011). Vortex Shedding and Galloping on a Rectangular Cross Section. Lyngby, Denmark:
Technical University of Denmark, Department of Civil Engineering.
Lawson, T. (2001). Building Aerodynamics. London: Imperial College Press.
Louisiana State University, H. C. (sd). Hurricane Engineering. Opgehaald van
http://www.hurricaneengineering.lsu.edu/
Normalisatie-instituut. (1976, March). Cranes. Stability (safety against turning over) [NEN 2022].
Normalisatie-instituut. (2007, August). Technical principles for building structures - TGB 1990 -
Loadings and deformations [NEN 6702].
Normalisatie-instituut. (2011, December). Eurocode: Basis of structural design [NEN-EN
1990+A1+A1/C2].

96
Normalisatie-instituut. (2011, December). National Annex to NEN-EN 1990+A1+A1/C2: Eurocode:
Basis of structural design [NEN-EN 1990+A1+A1/C2/NB].
Normalisatie-instituut, N. (1983, October). Cranes, Loads and combinations of loads [NEN 2018].
Normalisatie-instituut, N. (2011, April). Crane safety - General design - Part 2: Load actions [NEN-EN
13001-2 (en)].
Normalisatie-instituut, N. (2011, December). National Annex tot NEN-EN 1991-1-4+A1+C2: Eurocode
1: Actions on structures - Part 1-4: General actions - Wind actions [NEN-EN 1991-1-
4+A1+C2/NB].
Noselli, D. B. (2011). Experimental evidence of flutter and divergence instabilities induced. Journal of
the Mechanics and Physics of Solids 59, 2208-2226.
Peng Huang, Y.-j. W. (2006). Wind Tunnel Test and Numerical Simulation of Mean Wind Loads on a
Container Crane. The Fourth International Symposium on Computiona lWind Engineering
(CWE2006), (pp. 893-896). Yokohama.
Perry, K. C. (2002). Guide to the use of the Wind Load Provisions of ASCE 7-98. Reston, Virginia:
American Society of Civil Engineers.
Sang-Joon Lee, J.-H. K. (2008). Wind load on a container crane located in atmospheric
boundarylayers. Journal of Wind Engineering and Industrial Aerodynamics 96, 193-208.
Seong-Wook Lee, R.-W. A.-S.-H.-J. (2007). The Structural Stability Analysis of a Container Crane
according to the Boom Shape using Wind Tunnel Test. Key Engineering Materials Vol. 347,
365-372.
Snbjrnsson, J. T. (2002). Full and model scale study of wind effects on a medium-rise building in a
built up area. Trondheim, Norway: Norwegian University of Science and Technology.
Standardization, I. O. (1981, May). Cranes - Wind load assessment [ISO 4302].
T. Zhou, S. M. (2011). On the study of vortex-induced vibration of a cylinder. Journal of fluid and
Structures 27, 903-917.

97

You might also like