Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Available online at www.sciencedirect.

com
Available online at www.sciencedirect.com
ScienceDirect
ScienceDirect
Available online at www.sciencedirect.com
Structural Integrity Procedia 00 (2022) 000–000
Structural Integrity Procedia 00 (2022) 000–000 www.elsevier.com/locate/procedia
ScienceDirect www.elsevier.com/locate/procedia

Procedia Structural Integrity 45 (2023) 28–35

17th Asia-Pacific Conference on Fracture and Strength and the 13th Conference on
17th Asia-Pacific Conference
Structural on and
Integrity Fracture and
Failure Strength
(APCFS 2022and
& the
SIF 13th
2022)Conference on
Structural Integrity and Failure (APCFS 2022 & SIF 2022)
Implementation of a nominal stress approach for the fatigue
Implementation of a nominal stress approach for the fatigue
assessment of aluminium naval ships
assessment of aluminium naval ships
Teresa Magoga*, Seref Aksu, Karl Slater
Teresa Magoga*, Seref Aksu, Karl Slater
Defence Science and Technology Group, Department of Defence, 506 Lorimer Street, Fishermans Bend 3207, Australia
Defence Science and Technology Group, Department of Defence, 506 Lorimer Street, Fishermans Bend 3207, Australia

Abstract
Abstract
Fatigue analysis is an important part of the structural design of weight-optimised naval ships, because they are often constructed
from
Fatiguealuminium
analysis alloys or high tensile
is an important part ofsteel. Due to the
the structural low fatigue
design strength of these
of weight-optimised naval metals,
ships,the welded
because joints
they are are vulnerable
often constructedto
cracking. Furthermore,
from aluminium alloys naval
or highships cansteel.
tensile be required
Due to the to operate
low fatiguein demanding
strength ofenvironments,
these metals, and remain joints
the welded in service longer than
are vulnerable to
assumed
cracking.during design. Thus,
Furthermore, navalship
shipsdesigners
can be and in-service
required managers
to operate require efficient
in demanding fatigue evaluation
environments, and remainapproaches.
in serviceOne industry-
longer than
accepted
assumed and relatively
during design. quick
Thus, method, documented
ship designers in design
and in-service codes, to
managers assessefficient
require the fatigue strength
fatigue of welded
evaluation joints isOne
approaches. the industry-
nominal
stress
acceptedor S-N
and curve approach.
relatively quick However, there are uncertainties
method, documented in design codes, associated withthe
to assess using a design
fatigue code;
strength ofifwelded
all of the conditions
joints for its
is the nominal
applicability
stress or S-Nare notapproach.
curve met, the analyst
However,must reliably
there interpret theassociated
are uncertainties code. Further,
with different specifications
using a design forofthe
code; if all theuse of S-N curves
conditions for its
leads to different
applicability fatigue
are not met,life
theestimates.
analyst mustIn this paper,interpret
reliably a refinement of theFurther,
the code. nominaldifferent
stress approach for joints,
specifications typical
for the aluminium
use of S-N curves
welded
leads toship details,
different is proposed.
fatigue The goal
life estimates. Inisthis
to inject
paper,rigour and practicality
a refinement into thestress
of the nominal fatigue analysisfor
approach of naval
joints,ships.
typicalThe
of refinement
aluminium
process leverages
welded ship both
details, in-serviceThe
is proposed. maintenance information
goal is to inject long-term strain
rigour and practicality measurements
into the of aof56naval
fatigue analysis m aluminium
ships. Thepatrol boat.
refinement
The sensitivity
process of the
leverages bothpredicted
in-servicefatigue life of welded
maintenance detailsand
information to the choice strain
long-term of S-Nmeasurements
curve, availableof afrom
56 m a structural
aluminiumdesignpatrolcode,
boat.
and
Thestress parameter
sensitivity of theextraction
predictedisfatigue
investigated. Finally, details
life of welded recommendations
to the choice foroffuture work are
S-N curve, provided.
available from a structural design code,
©and2023
stressThe Authors.
parameter Published
extraction by ELSEVIER
is investigated. B.V. recommendations for future work are provided.
Finally,
© 2023 The Authors. Published by Elsevier B.V.
This
© 2023is anTheopen accessPublished
Authors. article under the CC BY-NC-ND
by ELSEVIER B.V. license (https://creativecommons.org/licenses/by-nc-nd/4.0)
This is an open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review
This is an under
open responsibility
access article of
under Prof.
the
Peer-review under responsibility of Prof. Andrei Kotousov Andrei
CC Kotousov
BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review
Keywords: under
Fatigue life; responsibility of Prof.marine-grade
naval ship performance; Andrei Kotousovaluninium alloys; S-N curves
Keywords: Fatigue life; naval ship performance; marine-grade aluninium alloys; S-N curves

* Corresponding author. Tel.:+61(3) 9344 3185.


* E-mail teresa.magoga@defence.gov.au
address:author.
Corresponding Tel.:+61(3) 9344 3185.
E-mail address: teresa.magoga@defence.gov.au
2452-3216 © 2023 The Authors. Published by ELSEVIER B.V. This is an open access article under the CC BY-NC-ND license
(https://creativecommons.org/licenses/by-nc-nd/4.0)
2452-3216 © 2023 The Authors. Published by ELSEVIER B.V. This is an open access article under the CC BY-NC-ND license
Peer-review under responsibility of Prof. Andrei Kotousov
(https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review under responsibility of Prof. Andrei Kotousov

2452-3216 © 2023 The Authors. Published by ELSEVIER B.V.


This is an open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0)
Peer-review under responsibility of Prof. Andrei Kotousov
10.1016/j.prostr.2023.05.010
2 Teresa
Author nameMagoga et al.Integrity
/ Structural / Procedia Structural
Procedia Integrity
00 (2019) 45 (2023) 28–35
000–000 29

1. Introduction

Fatigue is an important factor in the structural design and through-life management of weight-optimised ships. Such
ships are typically constructed from marine-grade aluminium alloys or high tensile steel. The reduction in the
scantlings due to the use of these materials has in some cases lead to fatigue cracking.

Nomenclature

D Fatigue damage
FEA Finite Element Analysis
FL Fatigue Life [years]
FLd Design fatigue life [years]
FLm Time from commissioning to discovery of first crack [years]
FLp Predicted fatigue life [years]
K Stress concentration factor
m1 Inverse slope of S-N curve for N ≤ 5x106 cycles
m2 Inverse slope of S-N curve for 5x106 < N ≤ 108 cycles
n Number of stress cycles
N Number of stress cycles to failure
S-N Fatigue resistance (stress range versus no. cycles to failure)
, S Stress range [MPa]
c Reference stress range of fatigue strength at 2x106 cycles [MPa]
max Maximum stress range [MPa]
crest Stress under design crest landing condition [MPa]
hollow Stress under design hollow landing condition [MPa]
 Number of stress ranges identified in a time history

Methods to assess the fatigue strength of welded joints include the nominal stress approach, hot-spot stress
approach, effective notch stress approach, fracture mechanics, and component testing. The first three methods are
categories of the S-N curve concept, which is based on empirical data collected from fatigue tests of common structural
details. If the detail of interest is represented in a standard then use of the nominal stress approach is valid. However,
the selection of a suitable reference detail, and assessment of structures characterised by complex geometry and/or
load combinations, can be difficult (Blagojevic et al. 2002, Collette and Incecik 2006, Shen et al. 2016, Soliman et al.
2015, Tveiten et al. 2007). As such, following an industry recognised code to conduct fatigue analysis can require
interpretation by the analyst. For example, the aluminium structural design code Eurocode 9 (Technical Committee
CEN/TC 250 1999) uses S-N curves based on nominal stresses. If the crack initiation site is a weld toe and the nominal
stresses in the joint are not clearly defined, the hot-spot stress approach is preferable. However, hot-spot S-N curves
must be available. In addition, selection of an applicable detail or S-N curve from a fatigue design code for a complex
structure can be challenging and somewhat subjective (Sielski 2008). Such complexities are reflected in the variability
of outcomes in the literature. Al Zamzami and Susmel (2017) conducted a comparative assessment of different
approaches based on extensive experimental data, finding that use of industry design curves with the nominal stress
approach provides an adequate level of accuracy. In the fatigue analysis of trapezoidal joints in a fast ferry, Garbatov
et al. (2010) assumed that the neighbouring ‘low-gradient’ is the nominal stress but did not provide a quantitative
definition of this stress. In comparison, Soliman et al. (2015) conducted fatigue analysis of a fillet welded detail in an
aluminium catamaran using the hot-spot stress approach because many of the construction details have no direct match
in design guides. Further, there are numerous variants in modelling and procedures for both the nominal stress and
hot-spot stress concepts (Radaj et al. 2009). Tveiten et al. (2007) demonstrated the variability in the stress levels at
highly stressed locations with different Finite Element Analysis (FEA) models and stress extrapolation techniques.
Information relating to fatigue failures and operational profile is valuable for the design, acquisition, and
management of ships, and to help validate fatigue evaluation approaches. The International Association of
30 Teresa
Author Magoga
name et al. / Integrity
/ Structural ProcediaProcedia
Structural00Integrity 45 (2023) 28–35
(2019) 000–000 3

Classification Societies (International Association of Classification Societies 2010) use ‘damage experience’, or the
number and location of cracks related to the fleet, as the main source of information for maintenance planning.
Similarly, the presence or absence of a crack at a construction detail can be interpreted as a sample of the corresponding
fatigue life (Groden and Collette 2017). However, use of in-service load and response data combined with survey
reports to update service life predictions has been limited (Hifi and Barltrop 2015). Fricke et al. (2002) compared
fatigue analysis approaches applied to a pad detail on a containership. Though this study highlighted the variability in
the results, a weakness was that the design life rather than the actual fatigue life was used as the baseline.
This paper proposes a refinement of the nominal stress approach for joints typical of aluminium welded ship details.
The refinement process leverages maintenance records and stress spectra derived from long-term strain measurements
acquired from a hull monitoring system onboard a 56 m naval ship, of marine-grade aluminium alloy construction.
The sensitivity of the predicted fatigue life of welded details to the choice of S-N curve, available from a structural
design code, and stress parameter extraction is investigated.

2. Method and Materials

Welded details of interest on a naval High Speed Light Craft, shown in Table 1, are studied. The ship was
constructed from aluminium alloy 5083-H321/H116 for plating, and 6082-T5/T6 for rolled sections.

2.1. In-Service Information

Structural fatigue failures were observed at common locations across the fleet. These observations provide an
opportunity to combine different sources of data and analysis approaches (Magoga et al. 2019).
The maintenance records of the fleet were reviewed to obtain the time from commissioning to first discovery of a
crack (FLm). Table 1 presents the fleet-wide average and standard deviation of FLm for each detail of interest,
normalised by that of ID-1. Cracking was first discovered at ID-2 followed by ID-1, ID-4, and ID-5. At the time of
the analysis, no defects had been observed at ID-3.
The standard deviation ranges between 9.3% and 31%. This variation is considered reasonable, as maintenance
data often suffers from issues with accuracy and completeness (Hodkiewicz and Tien-Wei Ho 2016). In addition, the
variation of FLm can be attributed to difficulties with accessing some parts of the structure, inconsistent defect
reporting, and operational variability across the class.

2.2. Finite Element Analysis

The commercially available FEA package MAESTRO (DRS Defense Solutions 2013) is used to calculate the
stresses in the ship structure. The global model of the ship is relatively coarse, though represents the overall stiffness
adequately. Fine mesh models of the details of interest are embedded into the global model. The mesh density is
approximately one thickness (t) of the material, which provides acceptable stress resolution and is commonly adopted
for the evaluation of stresses at welded joints (Hobbacher 2008). Linear elasticity is assumed.

2.3. Fatigue Life Analysis

The nominal stress approach to calculate the fatigue life of a welded structural detail is the most commonly
practiced in the maritime industry (Collette and Incecik 2006, Horn et al. 2009, Soliman et al. 2015), and is employed
in the present study. This method uses the stress acting on the location to be assessed, neglecting the stress
concentration arising from both the structural configuration and the weld. These effects are inherently taken into
account in the S-N curves.
Eurocode 9 (Technical Committee CEN/TC 250 1999) provides S–N curves relating nominal applied cyclic stress
ranges S to the corresponding number of cycles to failure N. The curves are based on estimates of the mean and
standard deviation, assuming a normal distribution, of observed logarithmic cycles for given logarithmic stress values.
These statistics are used to obtain a characteristic regression line for a probability of survival of approximately 97.7%
from the mean. This is less than 80% of the corresponding mean strength value, and allows for wider variations in
Teresa Magoga et al. / Procedia Structural Integrity 45 (2023) 28–35 31
4 Author name / Structural Integrity Procedia 00 (2019) 000–000

production than can be expected in a set of fatigue specimens. This is relevant because the welding procedure in a
shipyard is different from that in a controlled laboratory (Kecsmar and Shenoi 2004). Most of the S-N curves comprise
three slopes: m1 is the inverse slope of the S-N curve for N ≤ 5x106 cycles, m2 is inverse slope for 5x106 < N ≤ 108
cycles, and any stress cycles below the cut-off limit (108 cycles) are assumed to be non-damaging. The detail categories
are expressed in terms of the reference value of fatigue strength at 2x106 cycles (c) and m1.

Table 1: Description, photograph, schematic, reference strain gauge, load direction, FLm normalised by FLm at ID-1, and possible analogous
Eurocode 9 details for each welded detail of interest
Ref. strain FLm/ Eurocode
ID Description Photograph Schematic gauge; FLm at ID- 9 details
direction 1 (c-m1)
CROSS-SECTION 7.4.3
(32-3.4)
Pillar (hollow tube)
joined to an end 1.0  7.5
1 A; vertical
plate by a bevel-butt 45 25% (18-3.4)
circumferential weld
9.3
(12-3.4)

7.2.1
Butt weld between a
(50-4.3)
relatively thick B; 0.72 
2
insert and thinner longitudinal 31%
7.2.2
deck plating
(40-4.3)

11.3
Built-up beam N/A –
(36-3.4)
comprised of web C; no
3
plating of different t1 longitudinal defect
t2 5.5
thickness reports
(40-4.3)

7.6
(36-3.4)
Longitudinal plating
joined to transverse B; 1.2  9.1
4
plating via a double longitudinal 9.3% (28-3.4)
fillet weld
9.2
(25-3.4)
11.3
(36-3.4)
Butt weld in girder
D; 1.7  7.2.1
5 flange, between
t1 longitudinal 10.3% (50-4.3)
different thicknesses

t2
7.2.2
(40-4.3)

The fatigue relationship for endurance between 10 5 and 5x106 cycles, where Ni is the number of cycles to failure
for the ith stress range i, is defined by Equation 1:

∆𝜎𝜎𝑐𝑐 𝑚𝑚1
𝑁𝑁𝑖𝑖 = 2 × 106 ( ) (1)
∆𝜎𝜎𝑖𝑖
32 Teresa Magoga et al. / Procedia Structural Integrity 45 (2023) 28–35
Author name / Structural Integrity Procedia 00 (2019) 000–000 5

The fatigue relationship between 5x106 and 108 cycles is defined by Equation 2:
𝑚𝑚2
∆𝜎𝜎𝑐𝑐 𝑚𝑚2 2 𝑚𝑚1
𝑁𝑁𝑖𝑖 = 5 × 106 ( ) ( ) (2)
∆𝜎𝜎𝑖𝑖 5

In Eurocode 9, the modified nominal stresses should be used when there are gross geometrical effects present in
the vicinity of the initiation site. Gross geometrical effects include cut-outs, re-entrant corners, and gross changes in
stiffness at junctions between open or hollow sections. For example, Fig. 1 reveals a significant stress concentration
at detail of interest ID-1 (circumferential weld at the top of a pillar) under the wave crest landing condition. There is
no clear nominal stress in the pillar. Therefore, it is judicious to use the modified nominal stress.

max

min

Fig. 1. Distribution of normal stress under wave crest landing condition at detail of interest ID-1

The ratio between max at the detail of interest to that at the reference location is denoted by K, in Equation 3:

(𝜎𝜎𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 −𝜎𝜎ℎ𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜 )𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 𝑜𝑜𝑜𝑜 𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 (∆𝜎𝜎𝑚𝑚𝑚𝑚𝑚𝑚 )𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑𝑑 𝑜𝑜𝑜𝑜 𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖


𝐾𝐾 = (𝜎𝜎 = (∆𝜎𝜎 (3)
𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 −𝜎𝜎ℎ𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜𝑜)𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙 𝑚𝑚𝑚𝑚𝑚𝑚 )𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙

The factor K is applied to i in Equations 2 and 3. Eccentricity in the butt weld of detail of interest ID-2 is also
taken into account according to Det Norske Veritas (2010).
To calculate the fatigue damage at a detail of interest, a reference stress spectrum is translated to the detail of
interest. To do this, the stresses at the strain gauge location and the detail of interest are determined from Finite
Element Analysis (FEA) under a unit load. Two quasi-static load cases are considered that represent the design Wave
Crest Landing and Wave Hollow Landing conditions defined in DNV High Speed Light Craft rules (2011).
The fatigue damage (D) can be predicted using cumulative damage theory (Miner 1945), where  denotes the
number of stress ranges identified in a stress time history and n i denotes the number of cycles at the ith stress range:
𝑛𝑛𝑖𝑖
𝐷𝐷 = ∑𝑖𝑖=1 (4)
𝑁𝑁𝑖𝑖

The fatigue life (FL) in years is the ratio of the service life in years to the fatigue damage:

𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆 𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙 (𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦)


𝐹𝐹𝐹𝐹(𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦) = (5)
𝐷𝐷

Approximately 4500 hours of strain measurements at different locations on the ship are available (Magoga et al.
2016). The strains are assumed to be the responses to the dominant wave loads applied to the ship structure. The strain
gauge locations and stress spectra and are shown in Fig 2.a and b, respectively.
Teresa Magoga et al. / Procedia Structural Integrity 45 (2023) 28–35 33
6 Author name / Structural Integrity Procedia 00 (2019) 000–000

Fig. 2. (a) Profile view schematic of patrol boat, showing strain gauge locations, b) stress spectra at strain gauge locations

3. Analysis and Results

Different Eurocode 9 details are considered in the fatigue life analysis of the details of interest (listed in Table 1),
based on consistency of category, weld type, and geometry. For example, Eurocode 9 detail 7.4.3 is potentially
applicable to detail ID-1 because the category (butt weld) and geometry (hollow section) are consistent. However, the
weld type is not consistent; ID-1 features a partial penetration weld whereas Eurocode 9 detail 7.4.3 is a full penetration
weld. Alternatively, Eurocode 9 detail 7.5 may be appropriate because it is a partial penetration weld though the
geometry differs to that of detail ID-1.
Fig. 3 displays FLp normalised by FLm (given in Table 1) for details of interest ID-1, ID-2, ID-4, and ID-5 based
on different Eurocode 9 details. As FLm is not applicable at ID-3, FLp normalised by the design life FLd is presented
(Fig. 4). For each detail of interest the stresses are recovered (and in turn fatigue life calculated) at increasing distance
between the element centroid and the weld intersection (d/t).

Fig. 3. d/t versus FLp/FLm at detail of interests ID-1, ID-2, ID-4, and ID-5

Fig. 4. d/t versus FLp/FLd at detail of interest ID-3


34 Teresa Magoga et al. / Procedia Structural Integrity 45 (2023) 28–35
Author name / Structural Integrity Procedia 00 (2019) 000–000 7

The results in Fig. 3 suggest that when d/t approximates one, there is reasonable prediction of the fatigue life
relative to the historical data when the Eurocode 9 detail judged most applicable is used. Based on this observation,
FLp/FLm for each detail of interest is presented in Table 2. Also presented are the values of FLp normalised by FLm at
detail of interest of ID-1.

Table 2: K, and FLp/FLM at d/t=1, and FLm/FLM for each detail of interest
ID K Eurocode 9 Detail FLp/FLm at d/t = 1 FLp/FLm at ID-1
1 3.3 7.4.3 0.9 0.9
2 5.0 7.2.1 1.0 0.7
3 1.7 11.3 4.0* (*FLp/FLd) 13.0
4 3.5 7.6 0.9 1.1
5 2.7 11.3 1.1 2.0

4. Discussion

It is proposed that the nominal stress parameter can be extracted by averaging the normal stresses recovered at a
distance of 1t from the weld/intersection, substantiated by: Good agreement between the predicted fatigue life values
and those derived from the maintenance records - ranging between 90% and 110% - in Table 2, and correlation
between the order of predicted and in-service fatigue failures (the order of FLp/FLm at ID-1 in Table 2 matches the order
of FLm/FLm at ID-1 in Table 1).
The results in Fig. 3 and 4 demonstrate the variability in fatigue life estimates with selection of Eurocode 9 S-N
curve, which was also shown by Aksu et al. (2015). Whilst the proposed application of the nominal stress approach is
reliant on the appropriate selection of the Eurocode 9 detail, the requirement for expert judgment in the design and
through-life management of structures is not new. Maljaars and Vrouwenvelder (2014) noted that despite reasonable
accuracy of the load measurements, expert judgement is often required to account for uncertainty in the applied loads
in the structural reliability assessment of bridge structures. Similarly, sound judgement based on the analyst’s practical
experience is needed to define the long-term seaway loads that sufficiently represent the fatigue demand that the ship
structure will experience, and to design certain welded details (Horn et al. 2009, Wang 2010). The reliance on the
appropriate selection of the S-N curve is need not be a limitation. Over a ship’s life-cycle its structural configuration,
and analytical models, can be expected to change. These changes lead to updating of the variables incorporated into
the structural predictions (Hifi and Barltrop 2015, Magoga et al. 2019).
The fatigue life values of the details of interest inferred from in-service information (presented in Table 1) should
be understood to be optimistic as it relies on ‘average behaviour’ and a limited number of samples. A crack most likely
propagated to a critical length (requiring rectification) some time before being reported. At the same time, use of the
Eurocode 9 S-N curves is conservative because they correspond to the mean life curve minus two standard deviation
from the experimental data. Therefore, it is recommended that the analyst be aware of the different sources of
(non)conservatism in the parameters required as input to a fatigue analysis. To build confidence in the proposed
modified stress approach, Monte Carlo concepts for statistical estimations may be useful. In addition, there is merit in
investigating Bayesian methods that address the small number of cracking reports (Beer et al. 2013, McNeish 2016).
The presented work contributes to the ship structures community as it expands on research to evaluate the fatigue
life of one detail on an aluminium vessel (Soliman et al. 2015), in-service data (strain monitoring and defect reports)
is utilised, and a modified nominal stress extraction process has been detailed to extend the guidance provided by
Eurocode 9. Whilst use of techniques such as fracture mechanics could be pursued, significant investment in obtaining
the necessary input information (crack growth rate data, initial flaw size) is needed.

4. Conclusion

Both naval ship designers and in-service managers require reliable fatigue evaluation approaches that are
sufficiently accurate but also time-efficient. The selection of an applicable detail or S-N curve from a fatigue design
code can be challenging, particularly when both the geometry and stress field are complex. Also, the determination of
the stress parameter is not always clear. Therefore, this paper proposed a refinement of the nominal stress approach
for joints typical of aluminium welded ship details, which extends upon the guidance provided by Eurocode 9. The
Teresa Magoga et al. / Procedia Structural Integrity 45 (2023) 28–35 35
8 Author name / Structural Integrity Procedia 00 (2019) 000–000

refinement leverages both in-service maintenance reports and stress spectra derived from long-term strain
measurements acquired from a hull monitoring system onboard a 56 m naval ship, of marine-grade aluminium alloy
construction. Therefore, the validity of the approach is established via comparison to in-service information (hull
survey reports).

Acknowledgement

This paper was developed with the support of the Australian Department of Defence.

References

Aksu, S., Magoga, T. and Riding, B. (2015) Analysis of HMAS GLENELG’s Onboard Structural Monitoring Data. In: Pacific 2015 International
Maritime Conference, Sydney, Australia
Al Zamzami, I. and Susmel, L. (2017) On the accuracy of nominal, structural, and local stress based approaches in designing aluminium welded
joints against fatigue. International Journal of Fatigue 101 2017/08/01/ 137-158
Beer, M., Ferson, S. and Kreinovich, V. (2013) Imprecise probabilities in engineering analyses. Mechanical Systems and Signal Processing 37 (1-
2) 4-29
Blagojevic, B., Domazet, Z. and Kalman, Z. (2002) Productional, Operational, and Theoretical Sensitivities of Fatigue Damage Assessment in
Shipbuilding. Journal of Ship Production 18 (4) 185-194
Collette, M. and Incecik, A. (2006) An approach for reliability-based fatigue design of welded joints on aluminium high-speed vessels. Journal of
Ship Research 50 (1) 85-98
Det Norske Veritas (2010) Fatigue Assessment of Ship Structures, Classification Note No. 30.7. Høvik, Norway,
Det Norske Veritas (2011) Rules for Classification of High Speed, Light Craft and Naval Surface Craft. Høvik, Norway,
DRS Defense Solutions (2013) MAESTRO 11.0.0. Stevensville,
Fricke, W., Cui, W., Kierkegaard, H., Kihl, D., Koval, M., Mikkola, T., Parmentier, G., Toyosada, M. ,Yoon, J.-H. (2002) Comparative fatigue
strength assessment of a structural detail in a containership using various approaches of classification societies. Marine Structures 15 1-13
Garbatov, Y., Rudan, S. and Guedes Soares, C. (2010) Fatigue assessment of welded trapezoidal joints of a very fast ferry subjected to combined
load. Engineering Structures 32 800-807
Groden, M. and Collette, M. (2017) Fusing fleet in-service measurements using Bayesian networks. Marine Structures 54 38-49
Hifi, N. and Barltrop, N. (2015) Correction of prediction model output for structural design and risk-based inspection and maintenance planning.
Ocean Engineering 97 114-125
Hobbacher, A. (2008) Recommendations for fatigue design of welded joints and components. XIII-1539-96 / XV-845-96, Paris, France,
International Institute of Welding
Hodkiewicz, M. and Tien-Wei Ho , M. (2016) Cleaning historical maintenance work order data for reliability analysis. Journal of Quality in
Maintenance Engineering 22 (2) 146-163
Horn, A., Andersen, M., Biot, M., Bohlmann, B., Mahérault-Mougin, S., Kozak, J., Osawa, N., Jang, Y., Remes, H., Ringsberg, J. ,van der Cammen,
J. (2009) TCIII.2: Fatigue and Fracture. In: Proceedings of the 17th International Ship and Offshore Structures Congress. Vol. 1. Seoul, South
Korea, Seoul National University 211-287
International Association of Classification Societies (2010) Common Structural Rules for Bulk Carriers and Oil Tankers. London, United Kingdom,
Kecsmar, J. and Shenoi, R. A. (2004) Some notes on the influence of manufacturing on the fatigue life of welded aluminum marine structures.
Journal of Ship Production 20 (3) 164-175
Magoga, T., Aksu, S., Cannon, S., Ojeda, R. ,Thomas, G. (2016) Comparison between Fatigue Life Values Calculated Using Standardised and
Measured Stress Spectra of a Naval High Speed Light Craft. In: 13th International Symposium on the Practical Design of Ships and Other
Floating Structures. Copenhagen, Denmark
Magoga, T., Aksu, S., Cannon, S., Ojeda, R. ,Thomas, G. (2019) Through-Life Hybrid Fatigue Assessment of Naval Ships. Ships and Offshore
Structures 14 (7)
Maljaars, J. and Vrouwenvelder, A. C. W. M. (2014) Probabilistic fatigue life updating accounting for inspections of multiple critical locations.
International Journal of Fatigue 68 2014/11/01/ 24-37
McNeish, D. (2016) On Using Bayesian Methods to Address Small Sample Problems. Structural Equation Modeling: A Multidisciplinary Journal
23 (5) 2016/09/02 750-773
Miner, M. (1945) Cumulative damage in fatigue. Journal of Applied Mechanics 12 159-164
Radaj, D., Sonsino, C. M. and Fricke, W. (2009) Recent developments in local concepts of fatigue assessment of welded joints. International
Journal of Fatigue 31 (1) 2009/01/01/ 2-11
Shen, W., Yan, R., Barltrop, N., Liu, E. ,Song, L. (2016) A method of determining structural stress for fatigue strength evaluation of welded joints
based on notch stress strength theory. International Journal of Fatigue 90 87-98
Sielski, R. (2008) Research Needs in Aluminum Structures. Ships and Offshore Structures 3 (1) 57-65
Soliman, M., Barone, G. and Frangopol, D. M. (2015) Fatigue reliability and service life prediction of aluminum naval ship details based on
monitoring data. Structural Health Monitoring 14 (1) 3-19
Technical Committee CEN/TC 250 (1999) Eurocode 9: Design of aluminium structures. Brussels, Belgium, British Standards
Tveiten, B., Wang, X. and Berge, S. (2007) Fatigue assessment of aluminum ship details by hot-spot stress approach. Transactions - Society of
Naval Architects and Marine Engineers 115 31-49
Wang, Y. (2010) Spectral fatigue analysis of a ship structural detail – A practical case study. International Journal of Fatigue 32 310-317

You might also like