Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.

The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

Reliability of beams according to Eurocodes in serviceability limit state

D. Honfi*, A. Mårtensson and S. Thelandersson

Division of Structural Engineering, Lund University, Box 118, SE-221 00 Lund, Sweden

Abstract

To achieve a relatively consistent probability of failure for structural elements, most design

codes apply reliability based code calibration process. Such approaches commonly focus on

the strength of the structural members, which is related to the ultimate limit state (ULS).

However in the design of beams the performance of the structural elements is often limited by

the serviceability requirements, which are related to the serviceability limit state (SLS) using

different load combinations than applied in the ultimate limit state.

The current study aims to investigate the reliability for serviceability design for flexural

members made of different materials (steel, concrete and timber) according to the

specifications of the Eurocodes. Second-order reliability method (SORM) is applied to

determine the reliability index for different design situations for beams subjected to bending.

The probabilistic models of basic variables for time invariant analysis have been taken from

the JCSS Probabilistic Model Code. The characteristic, the frequent and the quasi-permanent

combination of actions are investigated and compared. The differences in service reliability

for different materials are discussed. The results show that there are differences between the

achieved reliability indices in the serviceability state between different materials and that for

the given load combinations in the Eurocode the reliability index is often below the one given

in the code.

*
Corresponding author. Tel.: +46 46 222 73 55; fax: +46 46 222 42 12.
E-mail addresses: daniel.honfi@kstr.lth.se (D. Honfi), annika.martensson@kstr.lth.se (A. Mårtensson)
sven.thelandersson@kstr.lth.se (S. Thelandersson)
1

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
Keywords: Eurocode; Serviceability; Reliability

1. Introduction

Previous surveys showed that most structural defects are related to serviceability conditions

rather than strength, thus serviceability problems are economically really important and

should be avoided [1]. Serviceability problems relate to deflections of horizontal members,

horizontal displacements of structures and vibrations. The main type of serviceability non-

compliances are excessive floor and roof deflections which may cause damages to adjacent

structural or nonstructural elements (crushing, cracking); gaps below partitions; dishing of

floors; jammed doors and windows; slanting furniture; ponding; water/moisture penetration;

damage to services; or simply be aesthetically annoying or give the feeling of being unsafe.

Several studies were carried out to investigate the consistency of Eurocodes in terms of

probability of failure for different structural elements and materials [2,3]. However these

investigations – such as the reliability based code calibration process – commonly focused on

the ultimate strength of the structural members. The characteristic values of actions and the

combination factors – which are used in serviceability limit states as well – are mainly

developed and optimized for the ultimate limit state. Other studies investigated the

consistency of the American and Australian design codes for serviceability [1,4] but no such

research was made referring the European structural standard family. Furthermore previous

investigations considered the deflection limits as deterministic variables. The present paper

aims to determine the reliability of Eurocode specifications for deflections considering a

structural member – made of different materials (concrete, steel and timber) – subjected to

bending considering the deflection limits themselves also as random variables, giving a new

aspect to service reliability. It must also be said that although this paper is restricted to

deflections of beams thereby omitting the important field of vibrations, it is very often the

2
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

deflection limits that are used in order to give an estimate of the risk for vibrations in practical

design situations. It is also obvious that if the system effects would be considered the

reliability of the structural system could be increased, but then consideration was to be taken

to a number of elements and also to the connections between elements. This would increase

the uncertainty of the modeling thereby complicating the analysis. The intent here is to show

the discrepancies between different materials and the uncertainties that influence the results

with the aim to identify crucial points for future studies.

2. Design according to Eurocode

2.1. Load combinations

The first part of the Eurocodes [5] defines the combinations of actions to be taken into

account in the relevant design situations. Three different load combinations are defined for the

serviceability limit states, which should be appropriate for the serviceability requirements and

performance. The code also distinguishes between irreversible and reversible limit states.

Irreversible serviceability limit states are those where some consequences of actions

exceeding the specified service requirements will remain when the actions are removed.

Contrarily at reversible limit states no consequences of actions exceeding the specified service

requirements will remain when the actions are removed.

The characteristic combination is defined as:

∑G
j ≥1
k, j + Qk ,1 + ∑ψ 0,i Qk ,i
i >1
(1)

where Gk,j denotes the characteristic value of the jth permanent action (i.e. the mean value if

the variability of G can be considered as small), Qk,1 is the characteristic value of the leading

variable action (i.e. the 98% fractile for a reference period of one year), Qk,i is the

characteristic value of the ith variable action and ψ0,i is the factor for combination value of a
3

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
ith variable action. The characteristic combination is normally used for irreversible limit

states.

The frequent combination is defined as

∑G j ≥1
k, j + ψ 1,1Qk ,1 + ∑ψ 2,i Qk ,i
i >1
(2)

where ψ1,1 denotes the factor for frequent value of a leading variable action and ψ2,i is the

factor for quasi-permanent value of the ith variable action. The frequent combination is

normally used for reversible limit states.

The quasi-permanent combination

∑G j ≥1
k, j + ∑ψ 2,i Qk ,i
i ≥1
(3)

The quasi-permanent combination is normally used for long-term effects and the appearance

of the structure.

2.2. Deflections

The vertical deflections of horizontal structural elements should be limited to avoid

deformations that affect appearance/comfort/functioning of the structure or that cause damage

to finishes or non-structural members.

wc
w1
w2 wtot
wmax
w3

Fig. 1. Definition of vertical deflections of a beam.

The definition of vertical deflections is shown in Fig. 1, where wc is the precamber in the

unloaded structural member; w1 is the initial part of the deflection under permanent loads of

the relevant combination of actions; w2 is the long-term part of the deflection under
4
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

permanent loads; w3 is the additional part of the deflection due to the variable actions of the

relevant combination of actions; wtot is the total deflection as sum of w1, w2, w3; wmax is the

remaining total deflection taking into account the precamber.

If the functioning or damage of the structure or to finishes, or non-structural members (e.g.

partition walls, claddings) is being considered, the verification for deflection should take

account of those effects of permanent and variable actions that occur after execution of the

member or finish concerned, see [5] A1.4.3 (3). However the code does not say anything

about which load combination should be used. In terms of the previous definition of

deflections it is a limitation of the incremental deflections w2+w3. If the comfort of the user, or

the functioning of machinery are being considered, the verification should take account of the

effects of the relevant variable actions, see [5] A1.4.3 (6). This is a limitation of w3, however

the code doesn't indicate what the word relevant means in this context. In case of the

appearance of the structure is being investigated, the quasi-permanent combination should be

used, see [5] A1.4.3 (4). It is not stated but the investigation is usually carried out for the

remaining total deflection wmax. Long term deformations due to shrinkage, relaxation or creep

should be considered where relevant, and calculated by using the effects of the permanent

actions and quasi-permanent values of the variable actions, see [5] A1.4.3 (6). This basically

means the quasi-permanent combination.

2.3. Deflection limits

The next question is: what are the limiting values of the previously defined deflections? And

how should they be calculated? Which load combinations should be used? These prescriptions

vary from material to material and are given in the appropriate parts of EC. It must be

emphasized that although limit values in some cases are given in the codes, the limit values

differ between materials, there is an uncertainty about which load combinations that should be

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
adopted in the calculations and the values given in the codes differs between different

countries. Information about the background of these values is also lacking.

2.3.1. Steel

In case of steel structures Eurocode 3 [6] states that the limits for vertical deflections

according to Fig. 1 should be specified for each project and agreed with the client and notes

that the National Annexes (NA) may specify these limits. However the values given in the

NAs are only suggested values and there are no compulsory rules given.

These prescriptions are quite varying and can be different depending on the function (e.g.

accessible/non-accessible roof, floor etc.), the importance (main girder, purlin), the type of the

carried material (plaster, brittle finish, non-brittle finish) or other conditions of the

investigated element. A typical value for wmax is L/250 and L/300 for w3.

2.3.2. Concrete

The deflection limits for concrete structures – given in Eurocode 2 [7] – should also take into

account the nature of the structure, of the finishes, partitions and fixings and the function of

the structure.

The appearance and general utility of the structure may be impaired when the calculated sag

of a beam, slab or cantilever subjected to quasi-permanent loads exceeds L/250, where the sag

is assessed relative to the supports. This criterion represents a limit for wmax.

Deflections that could damage adjacent parts of the structure should also be limited. For the

deflection after construction, L/500 is normally an appropriate limit for quasi-permanent

loads. It means that the damage criterion is considered by limiting w2+w3 here. Of course

other limits may be considered, depending on the sensitivity of adjacent parts.

6
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

2.3.3. Timber

In case of timber Eurocode 5 [8] is a bit confusing, since it gives a different definition of

deflections than was given in [5] (see section 2.2). The components of deflection resulting

from a combination of actions are shown in Fig. 2, where the symbols are defined as follows:

u0 is the precamber; uinst is the instantaneous deflection; ucreep is the creep deflection; ufin is the

final deflection and unet,fin is the net final deflection.

Fig. 2. Components of vertical deflections (timber).

The recommended range of limiting values of deflections for beams with span L is also given

in the code (see Table 1) depending upon the level of deformation deemed to be acceptable.

Table 1
Examples of limiting values for deflections of timber beams on two supports [8].
uinst unet,fin ufin
L/300 to L/500 L/250 to L/350 L/150 to L/300

2.4. Target reliabilities

A measure of reliability is conventionally defined by the reliability index β which is related to

the probability of failure Pf by:

Pf = Φ(− β ) (4)

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
where Φ is the cumulative distribution function of standardised normal distribution. For the

purpose of reliability differentiation, EC [5] establishes reliability classes (RC).

Information about the target reliability index for serviceability limit state can be found only

for class RC2. The target reliability for irreversible (!) serviceability limit states in [5] Annex

C is set to be 1.5 for a 50 years reference period and 2.9 for a 1 year reference period for RC2

structural members i.e. structures with medium consequences of failure (e.g. residential and

office buildings, public buildings).

3. Reliability analysis

3.1. Limit state function

In the present paper serviceability failure is deemed to occur when a deflection exceeds an

allowable deflection limit as a result of flexure. For probabilistic calculations this failure can

be described by a limit state function:

g ( X) = g (δ limit , δ max ) = θ Rδ limit − θ Eδ max (5)

where X = x1,..., xn the vector of basic variables, δlimit is the allowable deflection limit –

considered as a random variable – and δmax is the maximum deflection of the beam for the

reference period. θE is the coefficient expressing the uncertainty of the action effect and θR is

the uncertainty of the resistance model (in this case of the deflection model). The probability

of serviceability failure is the probability of exceeding the serviceability limit state:

p f = P[g ( X) ≤ 0]. (6)

3.2. Models of the basic variables

To calculate the reliability index β from the limit state function the probability distribution

functions of the different design variables are required. They were taken from the JCSS

8
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

Probabilistic Model Code [9] and [10]. The reliability was then calculated applying Second

Order Reliability Method using the structural reliability software Comrel 8.10 [11].

3.2.1. Loads

The dead load is assumed to be normally distributed with a coefficient of variation (COV) 0.1

and a mean value equal to the characteristic value of the action Gk [9],

The modelling of the imposed loads is the most uncertain of the stochastic variables. Different

studies use different coefficient of variations ranging from 0.2 to 0.6 [2,3,12,13].

Before the reliability analysis Monte Carlo simulation was carried out to estimate the

distribution of the variable loads with a 1 year reference period for floors with different user

category. The magnitude of the sustained load was assumed Gamma distributed with expected

value µs and standard deviation:

2 2 A0
σ s = σ v,s + σ u ,s κ (7)
A

where A0 is the correlation area, A is the influence area (for beams 2 times the tributary area

AT) and the κ is peak factor (κ=1.4 for beams). The magnitude of the intermittent load is also

assumed to be Gamma distributed with expected value µi and standard deviation:

2 A0
σ i = σ u ,i κ (8)
A

Both loads – sustained and intermittent – are modelled as Poisson processes with intensities λ

and ν respectively. The arbitrary-point-in-time values of the variable loads for different

occupancy types – and other parameters required for the simulation – are given in Table 2.

Table 2
The parameters of the variable loads for floors [9].
A0 µs σv,s σu,s 1/λ µi σu,i 1/ν dp
[m2] [kN/m2] [kN/m2] [kN/m2] [a] [kN/m2] [kN/m2] [a] [d]
9

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
office 20 0.50 0.30 0.60 5 0.20 0.40 0.3 1-3
residence 20 0.30 0.15 0.30 7 0.30 0.40 1.0 1-3
lobby 20 0.20 0.15 0.30 10 0.40 0.60 1.0 1-3
hotel room 20 0.30 0.05 0.10 10 0.20 0.40 0.1 1-3
patient room 20 0.40 0.30 0.60 5 0.20 0.40 1.0 1-3

The upper tail of the simulated results is fitted with a Gumbel distribution representing the

total variable load. Since the characteristic load is defined as the 98th percentile of the annual

load, the following expression is valid:

µQ 1
= (9)
Qk 6
{− ln[− ln(0.98)] − 0.5772} VQ + 1
π

where µQ is the mean value, VQ is the coefficient of variation and Qk the characteristic value

of the variable load Q.

The results of the simulations show that the mean value µQ and the coefficient of variation VQ

of the imposed load change with the tributary area. Eurocode 1 [14] recommends a reduction

factor αA to take into account the changing of the mean value. Fig. 3 shows, how the mean

value to the modified characteristic value changes with increasing tributary area for different

kind of floors.

Fig. 3. Changing of the mean value of the fitted Gumbel distribution with the tributary area.

10
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

From the above investigation it is clear that µQ and the VQ depend on the occupancy and the

tributary area. In the following 4 different design situations will be studied with µQ/Qk equal

to 0.4, 0.5, 0.6 and 0.7 representing the situation with a large tributary area for a lobby, an

office, a residence and a hotel room respectively. The coefficient of variations belonging to

these mean values are 0.578, 0.386, 0.257 and 0.165 respectively. To model the action effect

θE is chosen lognormally distributed with a mean value equal to 1 and COV=0.1 [9].

3.2.2. Material properties

The material models and the applied resistance factors are given in Table 3 for steel, Table 4

for concrete and Table 5 for timber. Since the resistance factor for midspan deflections is not

given in the JCSS code the values for the midspan moment were applied.

Table 3
The simplified probabilistic models of basic variables for steel.
Description X Distribution µX σX
Young’s modulus E Normal En 0.04µX
Moment of Inertia I Normal In 0.03µX
Resistance factor θR Lognormal 1 0.05µX
Table 4
The simplified probabilistic models of basic variables for concrete.
Description X Distribution µX σX
Young’s modulus of reinforciment Es Normal Es,n 0.04µX
Width of the beam b Normal bn-0.003bn 4mm+0.006bn
Height of the beam h Normal hn-0.003hn 4mm+0.006hn
Effective depth d Normal dn 0,02µX
Reinforcement area As Normal As,n 0.02 µX
Concrete compressive strength fck Lognormal fck+2σX 0.17µX
Resistance factor θR Lognormal 1 0.10µX
Table 5
The simplified probabilistic models of basic variables for timber.
Description X Distribution µX σX
Young’s modulus E Normal En 0.13µX
Height of the beam h Normal hn 0.005µX
Width of the beam b Normal bn 0.02µX
Resistance factor θR Lognormal 1 0.10µX

11

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
3.2.3. Deflection limits

A critical part of the analysis is modeling the uncertainties of the deflection limits. There

exists very little information about this in literature, primarily due to the fact that it is very

hard to measure and that the acceptance of deflections depend a lot on the situation. Hossein

and Stewart [15] describe a probabilistic model for deflection limits. They consider two types

of damaging deflections: perception damage and partition wall damage. The former is related

to the total deflections (wmax) while the latter to the incremental deflections, where

incremental deflection is that part of the total deflection that takes place after the construction

of partition walls (w2+w3). The parameters of the probabilistic model of the deflection limits

are given in Table 6. Since measuring the incremental deflections is quite difficult, the field

measurements are usually restricted to detect the total deflections thus the proposed model for

partition wall damage corresponds to total deflection values. The authors in [15] suggest that

the incremental deflections may be around 80-85% of total deflections. Hence in the current

probabilistic study the limit for incremental deflection is multiplied by 1.25 to make them

comparable with the total deflection.

Table 6
The simplified probabilistic models of basic variables for the deflection limits.
Description X Distribution µX σX
Perception damage wmax Lognormal 0.0077L 0.42µX
Partition wall damage wmax Gamma 0.0054L 0.57µX

4. Results

To investigate the effect of the variable actions a load ratio χ is defined representing the ratio

of the variable load to the total load:

Qk
χ= . (10)
Gk + Qk

12
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

The partial safety and combination factors applied in the calculations are the ones given by

Eurocode: γG=1.35, γQ=1.5, γM0=1.0, ψ0=0.7, ψ1=0.5 and ψ2=0.3 (for office, domestic and

residential floors).

4.1. Steel

In case of steel beams the design equation for the remaining total deflection is formulated as:

5 qL4 L
wmax = ≤ (11)
384 EI 250

and for the deflection from the variable actions as:

5 qL4 L
w3 = ≤ (12)
384 EI 300

q is the uniformly distributed load − calculated from the appropriate load combination −, L is

the span, E is the modulus of elasticity and I is the second moment of inertia.

Fig. 4 presents the reliability index (with a reference period 1 year) for a steel beam with

varying load factor for the characteristic load combination. The values are below the target

reliability index (β=2.9) and are very low for small values of variable loads. Furthermore the

higher the value of µQ/Qk the higher is the reliability.

Fig. 4. Reliability indices for a steel beam (wmax=L/250, characteristic combination).


13

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
Fig. 5. Reliability indices for a steel beam for w3 (characteristic combination).

For w3 the load ratio χ does not influence the results, since the deflection is given by the

variable loads only. However, applying stricter limits increases the reliability – as shown in

Fig. 5. Increasing µQ/Qk ratio of the variable load decreases the reliability. One should note

that the target value is given for irreversible limit states only, thus values far below the dashed

line do not necessarily mean that the reliability is unacceptable.

5.2. Concrete

The remaining total deflection of a simply supported concrete beam considering creep and

cracking of the cross section is checked as:

5 qL4 5 qL4 L
wmax = ζ + (1 − ζ ) ≤ (13)
384 Ec ,eff I 2 384 Ec ,eff I1 250

and the incremental deflections is checked as:

5 qL4 5 qL4 L
w2 + w3 = ζ + (1 − ζ ) ≤ (14)
384 Ec ,eff I 2 384 Ec ,eff I1 500

where I1 and I2 are the moment of inertia of the uncracked and fully cracked cross-section

respectively, ζ is the distribution coefficient – taking account of the degree of cracking.

The effective modulus of elasticity Ec,eff is defined as:

14
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

Ecm
Ec , eff = (15)
1 + ϕ (∞, t 0 )

where Ecm is the secant modulus of elasticity of concrete, φ(∞,t0) is the final creep coefficient

and t0 is the time at initial loading.

Fig. 6 presents the reliability index (with a reference period 1 year) for a concrete beam with

varying load factor for the characteristic load combination. The results are very similar to

those calculated for a steel beam, however in case of concrete the long-term effects are more

of interest, thus the quasi-permanent combination has to be investigated. Nevertheless

considering the quasi-permanent combination the effect of increasing the variable load is

totally different than for the characteristic combination as shown in Fig. 7. Increasing χ the

reliability decreases especially if the µQ/Qk value is high (e.g. residence, hotel room). That

means that if the dimensions of the beam are governed by the deflection criteria, the limits

will be exceeded with a quite high probability. If we think about the damage criterion it can

be a problem, but on the other side it should be mentioned that the consequences of exceeding

a deflection limit with just a little is usually different than exceeding an ultimate limit state.

Fig. 6. Reliability indices for a concrete beam (characteristic combination, wmax=L/250).

15

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
Fig. 7. Reliability indices for a concrete beam (quasi-permanent combination, wmax=L/250).

Fig. 8 presents the situation for the frequent combination. If the ratio of µQ/Qk is high, the

reliability is quite consistent. With higher µQ/Qk of the variable load (e.g. hotel room) β

decreases with increasing load ratio, however the effect is not that significant like in case of

the quasi-permanent combination. In case of concrete the incremental deflection (w2+w3)

consist the effect of the permanent load, thus the load ratio χ influences the results. Fig. 9

shows that the behaviour is very similar to those experienced for wmax, although the β values

are somewhat higher and do not fall below zero. The results presented in this case are

calculated considering the quasi-permanent combination.

Fig. 8. Reliability indices for a concrete beam (frequent combination, wmax=L/250).

16
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

Fig. 9. Reliability indices for a concrete beam (quasi-permanent combination, w2+w3=L/500).

5.3. Timber

For timber the deflection limits applied and how they were calculated are as follow:

- for the instantaneous deflection:

5 L4 L
u inst = (Gk + Qk ) ≤ (16)
384 E 0, mean I 400

- for the final deflection:

5 L4 L
u fin =
384 E 0,mean I
[ ]
Gk (1 + k def ) + Qk (1 + ψ 2Q k def ) ≤
200
(17)

- for the net final deflection:

5 L4 L
u net , fin =
384 E 0,mean I
[ ]
Gk (1 + k def ) + Qk (1 + ψ 2Q k def ) − u 0 ≤
300
(18)

where kdef is the creep factor depending on the type of the wood-based material and the

service class. This definition is not equivalent to calculating the deflections from the quasi-

permanent load combination! In case of the variable load it includes the rare and the quasi-

permanent part of the variable action and hence leads to a significantly higher load level than

in case of concrete for instance.


17

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
Although it is not indicated in EC5 but from the EC0 prescription it seems to be logical to use

an additional criterion as suggested in [16]. For the deflection from the time dependent part of

the permanent loads and the variable loads (w2+w3):

5 L4 L
u 2, fin =
384 E 0,mean I
[ ]
Gk k def + Qk (1 + ψ 2Q k def ) ≤
400
(19)

In case of timber all of the previously defined deflections contain the effect of the permanent

loads, thus the effect of load ratio is of interest in all cases. Fig. 10 shows the reliability

indices for the different deflections considering the deflection limit values given in section

2.3.3 with µQ/Qk=0.5 (e.g. office foor) for a reference period 1 year. For all 4 cases the

reliability decreases by increasing the ratio µQ/Qk. This is significantly different from what is

observed for concrete and steel. The reason is the different interpretation of the deflections

and the related load combinations. It is interesting how close the reliability indices for unet,fin

and u2,fin are to each other, even though the former is based on the damage perception

criterion, while the latter on the partition wall damage criterion. The calculations are made

assuming no precamber. Note that if that is the case ufin and unet,fin are the same and since the

latter has a stricter limit the former cannot be critical.

Fig. 10. Reliability indices for a timber beam (µQ/Qk=0.5).

18
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

5. Conclusions

The results of the current paper show that the reliability in SLS is not consistent for different

load ratios and is below the target value (β=2.9), which is given in the code for irreversible

serviceability limit states. In general we can say that the reliability of serviceability of

Eurocodes seems not to be always consistent − since different load levels are used for

different materials − and should be improved.

With regard to the different materials we can conclude that:

• In case of steel the reliability increases with higher portion of variable loads in the

total load which is a favorable situation, since steel structures are usually light and

carry more variable load than self-weight.

• Considering concrete beams in the quasi-permanent load combination the probability

of exceeding the given deflection limit is quite high especially for higher variable

loads, which can cause problems for innovative light-weight concrete structures.

• For timber members the EC prescriptions to calculate the long-term deflections differ

from those given for concrete which results in different effect of increasing the

variable loads to the reliability for serviceability limit state.

There is obviously a need for further investigation in this field, this analysis indicates that

some of the main objectives with future studies should be to discuss relevance of limit values

and also what are the relevant reliability indices in serviceability. System effects must also be

discussed since all structural elements are parts of larger system. The latter point is, however,

difficult and to perform such analysis sound knowledge about the individual elements is

needed.

19

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
References

[1] Stewart MG. Optimization of serviceability load combinations for structural steel beam

design. Structural Safety 1996;18(2): 225-238.

[2] SAKO 1999. Basis of design of structures – proposals for modification of partial safety

factors in Eurocodes. Joint committee of NKB and INSTA-B. NKB Committee and work

reports, 1999:01 E.

[3] Gulvanessian H, Holicky M. Reliability based calibration of Eurocodes considering a steel

member. JCSS Workshop on Reliability Based Code calibration; 2002.

[4] Stewart MG. Serviceability reliability analysis of reinforced concrete structures. Journal of

Structural Engineering 1996;122(7): 794-803.

[5] EN 1990 Eurocode – Basis of structural design. Europe´en de Normalisation, Brussels,

Belgium; 2002.

[6] EN 1992-1-1 Eurocode 2 – Design of concrete structures: General rules and rules for

buildings. Europe´en de Normalisation, Brussels, Belgium; 2002.

[7] EN 1993-1-1 Eurocode 3 – Design of steel structures: General rules and rules for

buildings. Europe´en de Normalisation, Brussels, Belgium; 2005.

[8] EN 1995-1-1 Eurocode 5 – Design of timber structures: General rules and rules for

Buildings. Europe´en de Normalisation, Brussels, Belgium; 2005.

[9] Model Code. Joint Committee of Structural Safety, JCSS; 2001. Available from:

http://www.jcss.ethz.ch.

[10] Sorensen JD, Svensson S, Dela Stang B. (2005) Reliability-based calibration of load

duration factors for timber structures. Structural Safety 27(2): 153-169.

[11] COMREL Version 8.10. RCP GmbH; 2008. www.strurel.de.

[12] Galambos TV, Ellingwood B. Serviceability limit states: Deflection. Journal of Structural

Engineering 1986;112(1) 67-84.


20
Published in Engineering Structures, Volume 35, February 2012, Pages 48-54.
The final publication is available at ScienceDirect via https://doi.org/10.1016/j.strusafe.2014.03.004

[13] Melchers RE. Structural Reliability: Analysis and Prediction. 2nd Edition, John Wiley

and Sons, Chichester; 1999. ISBN 0 4719 8771 9

[14] EN 1991-1-1 Eurocode 1 – Actions on structures. Europe´en de Normalisation, Brussels,

Belgium; 2002.

[15] Hossain NB, Stewart MG. Probabilistic models of damaging deflections for floor

elements. Journal of Performance of Constructed Facilities 2001;15(4): 135-140.

[16] Blass HJ. Timber Engineering STEP 1: Basis of design, material properties, structural

components and joints; 1995.

21

© 2012. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/

You might also like