CIE4160 Textbook February 2019
CIE4160 Textbook February 2019
CIE4160 Textbook February 2019
Prestressed concrete
February 2019
Faculty of Civil Engineering and
Geosciences
This book is basically the English translation of the existing book “Voorgespannen
Beton”, that has been used already for a number of years as background material for the
courses.
This second English version is an update from the Dutch code NEN 6720 (VBC) to a
Eurocode (EN 1992-1-1). In spite of all efforts there might be inconsistencies, specially
since (national annexes to) the Eurocodes are regularly updated. The authors are grateful
to everybody who brings errors to their notice and for any constructive criticism.
February 2019
Contents
1. Introduction
1.1 The principle of prestressing
1.2 Application of the prestressing force
1.3 Curved tendon profiles
1.4 Prestressed concrete versus reinforced concrete
1.5 Literature
8. Shear
8.1 Introduction
8.2 Shear behaviour of elements not containing prestressing
8.2.1 Reinforced elements without shear reinforcement
8.2.2 Reinforced elements with shear reinforcement
8.3 Prestressed elements without shear reinforcement, serviceability limit state
8.4 Prestressed elements loaded in shear, ultimate limit state
8.4.1 Elements without shear reinforcement
8.4.2 Structures with shear reinforcement, prestressed with straight
prestressing tendons
8.4.3 Structures with shear reinforcement, prestressed with draped tendons
8.4.4 Influence of ducts in the cross-section
8.5 Vertical prestressing of the web
8.6 Literature
1. Introduction
1.1 The principle of prestressing
Concrete is a building material that is strong in compression, but relatively weak in tension.
Therefore, the occurrence of tensile stresses in the structural design makes that the designer
has to take action. There are two ways to address this:
embed another material strong in tension in the concrete, in which case it is denoted as
reinforced concrete (cracking is not prevented, but crack widths should be limited);
compensate for the tensile forces, which can be achieved by arching or prestressing.
This course focuses on the principle of prestressing.
Because of the relatively low tensile strength of concrete, a reinforced concrete beam will
start to crack at an outer fibre at a relatively small load. In the opposite outer compressive
fibre, the compressive strength of the material is then far from reached. After cracking, the
longitudinal reinforcement takes over the tensile force (fig. 1.2). Although the beam is
cracked, it can sustain the load provided that a sufficient amount of steel reinforcement is
applied. Provided that the crack widths in the serviceability limit state (SLS) are sufficiently
limited, the expected service life time is not affected.
εct
εcb
Thus, structures in reinforced concrete should meet crack width requirements. Moreover, the
structure should meet SLS deflection requirements and should have sufficient resistance at the
ultimate limit state (ULS). Usually this is not a problem: by a rational choice of the
dimensions of a structural element (e.g. height h lmin 35 for floor slabs), deflection
requirements are met implicitly. In addition, the longitudinal reinforcement required to resist
the bending moment at the ULS is in most cases sufficient to meet crack width requirements.
However, a number of factors exist that limit the application of reinforced concrete:
1. The load bearing capacity of an element cannot be increased unlimitedly by still further
increasing the amount of reinforcement. For a reinforcement ratio of over 0,02 (depending
on the strength class of the concrete and the type of reinforcing steel), the compressive
strength of the concrete is governing at ULS and the reinforcing steel will not reach its
yield strength. As a result, an undesired brittle fracture of the element can occur. In
addition, it often becomes practically difficult to apply such a high amount of reinforcing
steel in the cross-section of the element.
2. When the span increases, the bending moment caused by the selfweight increases more
than proportional ( 18 qG l 2 ). The ratio between the variable load that can be applied and the
permanent load becomes more and more unfavourable. Furthermore, problems may arise
with regard to limiting the deflections below certain limit values.
Due to the variable load, compressive stresses are generated at the top side of the beam and
tensile stresses at the bottom side. This stress distribution has to be superposed on the stress
distribution that is already present (fig. 1.3c). The result is shown in fig. 1.3d. Again, tensile
stresses (bottom side) and compressive stresses (top side) must be limited.
Fig. 1.3 Stresses in the cross-section at midspan, as a result of the prestressing force (P),
permanent loading (G) and variable loading (Q)
rigid support
a structural member
jack
A A
b
prestressing steel (bar or cable) jack
c sheath element
Fig. 1.4 Various prestressing methods: (a) prestressing between rigid supports; (b) pre-
tensioning; (c) post-tensioning
Method 2: Pre-tensioning
Tendons are positioned, stressed and anchored before the concrete is cast in the mould (part
A-A of Fig. 1.4b). Once the concrete has sufficiently hardened, the ends of the tendons are
gradually or instanteneously released from the external fixing points. From this moment on,
the concrete element is prestressed. The ends of the tendons are cut off. The prestressing force
is transferred from the tendons to the concrete over a certain distance (the transmission length
lpt2) by the bond stresses between concrete and tendon. The magnitude of the transmission
length lpt2 depends on tendon type and its cross-section, surface profile (roughness) and initial
stress (lpt2 = 300 - 700 mm). This method is very suitable for the pre-fabrication of elements.
The beam from fig. 1.3 is prestressed to such a level that the tensile stresses in the critical
cross-section caused by the permanent and variable load (bottom fibre at midspan) are
compensated for by compressive stresses. However, this way of prestressing generates tensile
stresses at the top side of the beam near the supports, which are not compensated for by the
bending moment caused by the permanent load (fig. 1.5: cross-section B-B).
Therefore, at the supports, the anchorages of the tendons should be positioned higher, for
instance within the kern area of the cross-section. Then, no tensile stresses occur. If the centre
of gravity of the anchorages is outside the kern area, tensile stresses occur. These stresses
should not exceed a certain limit value. In most cases, the centre of gravity of the tendons at
the support is chosen such that it coincides with the centre of gravity of the concrete cross-
section. The tendon now is no longer straight, but has a curved shape. When a parabolic
tendon profile is used (fig. 1.6), the stresses resulting from the prestressing force have the
same profile as the stresses caused by a uniformly distributed load.
qG
stresses in
section A-A
by Pm by qG by qQ
stresses in
section B-B
by Pm only
by Pm only
Because of the curvature of these tendons, not only forces Pm act on the end faces of the
beam, but also forces perpendicular to the tendon profile. This distributed load acts in upward
direction and is called the curvature pressure (qp).
In most cases, not one but several tendons are applied, see fig. 1.7. The tendons are then
positioned such that their overall centre of gravity more or less coincides with the parabolic
profile as shown in fig. 1.6. The tendons are spread over the end faces to provide space for the
anchorages and to have a more uniform distribution of the (often high) anchor forces.
The pre-tensioning principle can be applied too (fig. 1.4). To obtain a similar distribution of
the prestressing as presented before, the tendons are kinked as shown in fig. 1.8. Note that this
is only possible when using tendons that are suited to be bent, for instance strands (not bars).
In most cases, a number of strands is not raised at their ends but is (almost) parallel to the
beam axis. This also has the advantage that a steel tensile tie (see section 8.4.3) is present at
the bottom side over the full length of the beam.
The slope of the tendon near the support has an additional positive effect: the upward directed
vertical component of the prestressing force reduces the shear force caused by the loads,
which results in a reduction of the shear force and, if required, less shear reinforcement
(stirrups).
Fig. 1.9 Test loading in the U.S. in 1954; span 9,2 m, thickness of the prestressed slab
50 mm, with a cast topping of 100 mm
It is of course also an option to increase the span when using the same structural height. The
absence or strong reduction of crack formation in SLS has advantages with respect to water
tightness and corrosion resistance. Therefore, offshore structures are often prestressed.
As an example, fig. 1.10 shows the construction of the F3 P1 offshore platform, designed for
the F3-field in the North Sea. This structure is built up with cells, which are prestressed at the
water retaining side. Thanks to the absence of cracks, the stress cycles in the steel are
relatively small. Therefore, fully prestressed concrete has a very high resistance against
fatigue.
Besides the advantages mentioned, prestressing has the disadvantage that it is often
expensive. The tendons themselves and, especially, the anchorages are more expensive than
the traditional reinforcing steel. Furthermore, the tensioning and grouting activities are quite
labour intensive.
Prestressed and reinforced concrete both have their own advantages and disadvantages. For
each application, investigation has to reveal which of the two options is the best.
Until recently, a fundamental difference existed regarding the design of reinforced concrete
and prestressed concrete. A structure in reinforced concrete was often designed with respect
to the ultimate limit state (ULS): when the dimensions of the concrete cross-section and the
reinforcement met the requirements at the ULS (i.e. a failure criterion; strength), then in most
cases the conditions with respect to the serviceability limit state (SLS; e.g. crack width
control, deflections) were met automatically.
For prestressed concrete, on the other hand, the design was based on the SLS. The design was
performed such that the concrete tensile and compressive stresses for the most unfavourable
load combination in SLS were within certain limits and the magnitudes of the corresponding
deformations were checked. Only afterwards, it was checked whether the requirements at the
ULS were met. The ULS requirements were most often met and it appeared that SLS
requirements were governing.
However, several decades ago it became clear that this clear separation between designing in
reinforced concrete on one hand and prestressed concrete on the other, not always resulted in
the most economical solution. Situations can occur in which a combination of both design
systems gives better results. This combination is denoted as partially prestressed concrete.
The designer then has to be aware that a reduction of prestressing will result in crack
formation at the SLS. This aspect requires special attention because prestressing steel is more
corrosion sensitive than reinforcing steel. Therefore, in most cases extra steel reinforcement is
applied to restrict the crack width.
In several countries, the application of “partial prestressing” has already been incorporated in
daily engineering practice. In the Netherlands, designers were very reluctant to apply partially
prestressed concrete, also because of the lack of specific regulations, codes and design rules.
With the introduction of TGB 1990 “Loads and deformations” (NEN 6702) in 1990, and the
“Regulations for concrete - Structural requirements and calculation methods” (NEN 6720,
also known as VBC 1990 and VBC 1995), the existing distinction between reinforced
concrete and fully/limited/partially prestressed concrete almost disappeared. As stated, this
was thanks to the introduction of a uniform design method for both reinforced and prestressed
concrete.
In a design according to NEN 6702 / NEN 6720, the prestressing has to be modelled as a load
applied on the structure. By doing so, a gradual transition between reinforced concrete on one
hand and fully prestressed concrete on the other, with all possible intermediate options, is
acquired. This integral approach is often referred to as “structural” concrete, no longer using
“reinforced” or “prestressed” concrete [1.2]. This appoach is used in EN 1992-1-1 too.
Fig. 1.11 Prestressed structure for the subway ring line in Amsterdam
1.5 Literature
1.1 Edwards H., “The innovators of prestressed concrete in Florida”, Journal of the PCI,
Special Issue for the FIP-Conference in Stockholm, June 6-10, 1982.
1.2 Bruggeling A.S.G., “Constructief beton”, Cement 1987, no. 1, 2, 3, 4, 6, 7, 9, 10, 12.
Bruggeling A.S.G., “Science into practice”, Heron, Vol. 32, 1987, no. 2.
Reinforcing steel was used in the first attempts to prestress structures. A calculation will
now demonstrate whether reinforcing steel is a suitable material to exert a prestressing
force on a structure for a long period of time.
17,5 mm
17,5 mm
14,1 mm
When applying a modulus of elasticity of the concrete Ec =35·103 N/mm2, the elastic
shortening of the concrete (subscript c) caused by releasing the prestressing force is
approximately:
12
lce 10 103 3, 4 mm
35 10 3
Creep of the concrete (plastic deformation under constant loading; subscript c) causes an
additional time-dependent shortening, which is about 2 to 3 times the elastic deformation,
i.e.:
lcc 8,5 mm
Shrinkage of the concrete (shortening due to drying of the concrete; subscript s) results in
a shortening on long term of about 0,3·10-3, or:
The total shortening caused by elastic deformation, creep and shrinkage is about 3,4 + 8,5
+ 3,0 = 14,9 mm. This implies that only 17,5 – 14,9 = 2,6 mm of the original steel bar
elongation of 17,5 mm remains. As a result, the initial tensile stress in the steel of
350 N/mm2 is reduced to (2,6/17,5) · 350 = 52 N/mm2, and the concrete stress σc is down
from -12 N/mm2 to (52/350) · -12 = -1,8 N/mm2 (compression).
It can be concluded that reinforcing steel is not suited to apply a prestressing force.
To apply prestressing successfully, the shortening of the concrete has to be kept small
compared to the elongation of the prestressing steel used. This can be achieved by
applying steel with a much higher tensile strength and strain at failure. Therefore,
prestressing steel has a tensile strength of 900 to 2000 N/mm2.
The steel type ranges from Y1030H to Y1230H, in which the number indicates the
nominal tensile strength in N/mm2. The bars are made from steel that is hot rolled into
bars (code H). Subsequent processing (e.g. accelerated cooling, cold stretching,
additional tempering) might be required to achieve the required mechanical properties.
The bars can be plain (code P) or ribbed (code R) (fig. 2.2).
Fig. 2.2 Ribbed (left) and plain steel (right) prestressing bars
Ribbed bars have better bond properties and offer the advantage that they can be cut and
anchored at any position, or can be extended by coupling. The modulus of elasticity is
205 GPa for bars that are either rolled only or as rolled stretched and tempered. For bars
that are as rolled stretched only, the E-modulus is significant lower and is about 165 GPa.
The possible combinations of bar diameters and steel strength according to EN 10138-4
are given in table 2.1
1100
1230
15,0 177 R
20,0 314 R
25,5 511 P
26,0 531 P P
Y….H 26,5 552 R R
32,0 804 P, R P, R
36,0 1018 P, R P, R
40,0 1257 P, R P, R
P = plain bar, R = ribbed bar
Ep = 205 GPa for as rolled and as rolled stretched and tempered bars
Ep = 165 GPa for as rolled and stretched bars
Note:
EN 1992-1-1 cl. 3.3.6 presents an E-modulus of 205 GPa for bars and wires; 195 GPa for
strands.
The modulus of elasticity of bars and wires is about 205 GPa. Properties of wires
according to EN 10138-2 are listed in table 2.2.
1670
1770
1860
3,0 7,07 x
3,2 8,04 x
4,0 12,57 x
5,0 19,63 x x
6,0 28,27 x
6,9 37,39 x
Y….C 7,0 38,48 x
7,5 44,18 x
8,0 50,27 x
9,4 69,4 x
9,5 70,88 x
10,0 78,54 x
Ep = 205 GPa
1770
1820
1860
1960
2060
2160
wires d [mm]
5,2 13,6 x x x
6,5 21,1 x x
Y….S3 3 6,8 23,4 x
7,5 29,0 x
6,85 28,2 x
7,0 30,0 x x
9,0 50,0 x x
11,0 75,0 x
Y….S7 7
12,5 93,0 x
13,0 100 x
15,2 140 x x
16,0 150 x x
18,0 200 x
12,7 112 x
Y….S7G
7 15,2 165 x
compact 18,0 223 x
Ep = 195 GPa
The steel type ranges from Y1670 to Y2160 (note: In NL the maximum steel quality
allowed for use is Y1860, except for foundation piles in which higher strengths are
allowed). The number of wires used is part of the code used: Y1860S7 denotes a 7 wire
strand constructed of strands having a tensile strength of 1860 N/mm2. The modulus of
elasticity is about 195 GPa.
Stress-strain relationship
Figure 2.5 shows the stress-strain diagram (σ-ε diagram) for a number of different steel
types. Prestressing steel often has no pronounced yield point. In these cases, as an
alternative for the yield point, the value fp0,1k is used, where fp0,1k is the stress that, after
unloading, causes a permanent deformation (plastic strain) of 0,1 %.
s [N/mm2]
prestressing steel
FeP 1860 steel
Y1770 cold-worked
FeP 1770 cold-workedsteel
steel yield value 0.1
tensile strenght
Y1670 cold-worked
FeP 1670 cold-workedsteel
steel
failure
1500
Y1230
FeP 1230
Y1030
FeP 1030
1000
reinforcing steel
B500B
FeB 500 HK
500
FeB 400 HWL FeB 400 HK
FeC 220
0
0.1% 5 10 15 20
[%]
The - relationships given in fig. 2.5 cannot be used directly in the calculation of the
bending moment resistance in ULS (the failure bending moment). Therefore, EN 1992-1-
1 uses schematised - diagrams (fig. 2.6). The position of the slope discontinuity in
this diagram is chosen such that the actual curve of fig. 2.5 is described quite accurately.
Note:
The subscript k refers to 'characteristic' value. This is a lower bound value derived from
test results, for instance from tensile tests on a prestressing bar or compressive tests on
concrete cubes. The characteristic value of the tensile strength of the prestressing steel fpk
follows from the characteristic value of the maximum force resisted by the steel and its
cross-sectional area. The design value of the strength follows from the characteristic
value, divided by a material factor γs.
Note that the NL National Annex to EN 1992-1-1 prescribes γs = 1,15 for reinforcing
steel and γs = 1,1 for prestressing steel (EN 1992-1-1 table 2.1N).
The value of fp0,1k can be calculated from EN 10138 by using Fp0,1k and Sn. According to
EN 1992-1-1 fig. 3.10, it is also allowed to assume that fp0,1k = 0,9 fpk. From this
assumption it follows that fpd = fp0,1k / γs = 0,9 fpk / 1,1.
The NL National Annex states that εud = 0,9 εuk. According to EN 10138 the minimum
value of εuk = 3,5 %.
Often, 7-wire strands from steel type Y1860S7 are used. Their nominal cross-sectional
area is often 100 or 140 mm2 (identification diameter Ø13,0 and Ø15,2 mm,
respectively), see Table 2.4.
The prestressing force depends on the maximum stress allowed after anchorage of the
prestressing steel (according to EN 1992-1-1 cl. 5.10.3: the minimum value of 0,75 fpk =
0,75·1860 = 1395 N/mm2 and 0,85 fp0,1k = 0,85·0,9·1860 = 1423 N/mm2 (Y1860)).
During stressing EN 1992-1-1 cl. 5.10.2.1 allows at the anchorage side a maximum stress
that is the minimum value of 0,8 fpk = 0,8·1860 = 1488 N/mm2 and 0,9 fp0,1k =
0,9·0,9·1860 = 1507 N/mm2 (Y1860).
When the jack has an inaccuracy of less than 5%, it is allowed to overstress to 0,95 fp0,1k
= 0,95·0,9·1860 = 1590 N/mm2 (Y1860) (Note that overstressing is not allowed
according to the Dutch National Annex. This is included in table 2.4).
Note that prestressing bars are not allowed to be overstressed.
Ducts are installed in the formwork or mould and the concrete is cast. As soon as the
concrete has developed sufficient compressive strength, the tendons (bars, wire bundles
or bundles of strands) are installed, tensioned and anchored. Shortly after installing the
anchorages, the open space between the tendons and the ducts is injected with a special
grout to protect the very corrosion sensitive prestressing steel. Finally, for corrosion
protection, the anchorages are covered, for instance with concrete.
Note:
According to EN 1992-1-1 cl. 5.10.2.2 the minimum concrete strength fcm(t) required to
be allowed to start stressing (stepwise per individual tendon) is 50% of the minimum
required compressive strength for full prestressing as given in the ETAG certificate of the
prestressing system. From this minimum value fcm(t) to the strength given in the
certificate, the prestressing force of a tendon can be stepwise increased from 30% of the
maximum tendon force up to the full 100% of maximum tendon force. Additionally,
during stressing, the concrete compressive stress caused by prestressing and other loads
should not exceed 0,6 fck(t). In the case of pre-tensioned steel, this value may be
increased to 0,7 fck(t).
The ducts for the tendons are created in the concrete using special thin-walled steel
sleeves (ducts, sheaths) as shown in fig. 2.7. The wall thickness of these ducts is about
0,2 to 0,4 mm. The ducts are produced in lengths of about 6 m and extended by
couplings. The couplings are sealed to prevent undesired leakage because this might lead
to problems during prestressing or injecting of the prestressing elements. The prestressing
is carried out by hydraulic jacks, examples of which are shown in figs. 2.8 and 2.9.
Injection should be carried out using moderate pressures; a too high pressure might lead
to pushing away (push-out) of the concrete cover on the sleeve, the generation of splitting
cracks in the concrete, or the penetration of grout into empty sleeves close to the duct
being injected.
During the injection of grout, the entrapped air accumulates at high positions in the
sleeve. Vent tubes are installed at these positions to release the air. This enables a
complete injection of the sleeve. A poorly performed injection with inclusion of air may
eventually lead to corrosion damage. The costs of repairs carried out later will often by
far exceed the original construction costs.
Several techniques exist for the anchorage and coupling of prestressing steel. A number
of often applied methods is discussed.
step 1:
Before installing the jack, an accessory (A) is
attached to enable accurate centring of the jack.
(A)
jack
wedges (B)
Step 2:
When the jack is stressed, the conical wedges
(B) are loosened.
travel
Step 3:
During tensioning, the prestressing tendon is
pulled out of the structure.
wedges
Step 4:
When the desired tendon force is reached, the
wedges are pressed hydraulically.
Step 5:
The oil is drained form the jack, thus
decreasing the jack force. The jack is removed
after full release of the force.
(a) (b)
Wedges
Upsetted
heads
Nut
Fig. 2.12 Several methods to attach the bars, wires and strands to the anchor plate
In split-wedge anchorages, three-parted or four-parted conical wedges are used, the inside
of which has fine and sharp serrations and the outside is smooth. The principle of this
type of anchorages is illustrated in fig. 2.12 (first row) and fig. 2.13.
The BBRV system realizes the anchorage by mechanical upsetting of the wire ends (fig.
2.12 second row). As presented before in fig. 2.11, anchoring can be achieved by nuts as
well (fig. 2.12 last row).
anchor plate
three-parted
conical w edges
Because large concentrated forces are transmitted through the anchor plates to the
concrete, the concrete compressive stresses usually will exceed the standard allowable
values. The application of spiral reinforcement (fig. 2.14) creates a volume in which an
external confining pressure is generated. This allows the concrete to resist these high
compressive stresses (EN 1992-1-1 cl. 3.1.9). The spiral reinforcement is an integral part
of the prestressing system and is also presented in the ETAG certificate of the system.
In addition, splitting reinforcement is applied, in order to prevent the generation of
splitting cracks in the concrete because of the introduction of large concentrated forces,
(also see Chapter 10).
Fig. 2.14 Spiral reinforcement to resist the high compressive stresses in the zone where
the prestressing force is transferred from the steel to the concrete
a b
For the construction of offshore structures with sliding formwork, a different type of
blind anchorage is applied for the prestressing in vertical direction (fig. 2.16c). The types
of anchorage shown before would be very unpractical in this case, because the blind
anchors would have to be cast in concrete at the bottom of the structure. This implies that
the tendons are installed already at the start of the sliding process, which causes a lot of
inconvenience during construction. This is overcome by initially installing only the
sleeves during the casting process. After the positioning of the prestressing anchorage on
top of the sleeve, the complete assembled prestressing element is lowered and the first
few meters at the bottom of the sleeve are injected. After hardening of the grout, the
prestressing elements are anchored by bonding. Bond is improved by applying plugs at
the lower ends of the strands. After tensioning the prestressing element, the rest of the
sleeve is injected. Additional reinforcement to distribute the high local load and to
prevent splitting of the concrete might be required in the transmission zone of the
prestressing force at the blind anchor.
sheath
80m
strand
grout pipe
forced on
anchorage
block
5m
b flat anchorage c special bond anchorage (Dywidag)
Coupling anchors
It might be required to couple prestressing elements, for example during staged
construction. The most basic type is the screw coupling (fig. 2.17).
In this system no grout is injected. The grout is replaced by an anticorrosive agent around
the tendon. This is done during the manufacturing process. The strands are provided with
a layer of protective grease and are in a plastic sleeve (fig. 2.20). The strand now is
protected against corrosion for a long period of time. Attaching a mono-strand to an
anchorage is shown in fig. 2.21.
PE-tube strands
cast anchor
wedges
anticorrosive grease
safety hanger
Thanks to the small cross-section of the tendon and the small required concrete cover, a
relatively large distance can be realised between the tendon and the neutral axis of the
concrete cross-section. This is especially important for thin structural elements such as
floor slabs.
This method of prestressing is mainly used for the industrial production of concrete
elements such as piles, floor and roof slabs and beams for bridges and industrial
buildings. This system is very well suited for the production of standardised elements.
Because the prestressing force is transferred by bond to the concrete over the so-called
transmission length (lpt2 ; EN 1992-1-1 cl. 8.10.2.2 & eq. (8.18)) only wires or strands are
used. The wires have some surface profile to limit the required transmission length. A
pronounced surface deformation would result in high bond stresses which, in turn, might
cause large splitting stresses in the surrounding concrete and even result in crack
formation.
Prestressing beds have lengths up to 200 m. Figure 2.22 shows plants for the
manufacturing of prestressed hollow-core slabs. After the tensioning of the tendons
between the abutments at the ends of the bed, concrete is cast continuously or per section
over the full length of the bed.
After hardening of the concrete, the concrete slab is cut into pieces of the required length
by a circular saw (fig. 2.23). The tendons are cut as well and they slip relative to the
concrete over the so-called transmission length (the ends of the tendons are pulled 0,1 –
0,7 mm into the concrete), and transfer the prestressing force to the concrete by bond (fig.
2.24). Because of the lack of concrete cover at the ends of the tendons, elements that are
applied outside (such as bridge beams) should be protected against corrosion by, for
example, bitumen.
Fig. 2.22 Factory for the production of prestressed prefab elements cast after tensioning
of the steel (prestressing with pre-tensioned steel)
strand
tension force
before
before sawing
sawing
through
steelstress
po
σp,max
saw cut
slipping of the wire bond stress
after after
sawingsawing
through
σp,0
pi
po
transmission length σp,max
stress decrease due to elastic
shortening of the concrete
Fig. 2.24 Stress in strands or wires before and after being cut (prestressing with pre-
tensioned steel)
In section 1.3 it was already mentioned that the positions of the tendons at the ends of the
beam should be such that at the top side no or just minor tensile stresses are generated.
For concrete elements with a small structural height, this can be achieved by straight
tendons. The large structural heights as applied in bridge beams require tendon profiles
that have raised ends (fig. 1.8). A similar effect can be obtained in, for instance, roof
beams by an increased height at midspan (fig. 2.25): at the supports, the prestressing
force is close to the centre of gravity of the cross-section whereas it is considerably lower
in the cross-section in the middle of the beam (namely in the area loaded in tension by
permanent and variable loads).
A concrete structure tends to deform during prestressing. The deformation should not be
restrained by the formwork or rigid supports. If, however, a restraint occurs, the forces
resulting from the restrained deformation should be accounted for properly.
During prestressing, both the prestressing force and the elongation of the tendons are
measured. In advance, their minimum and maximum values should be prescribed by the
designer. This information has to be available at the construction site in a “prestressing
protocol”. This protocol should also contain the order of prestressing in case more
prestressing tendons are applied. It also has to be indicated whether just one or both sides
of the tendons have to be stressed.
The designer should be aware of the fact that already stressed tendons will loose some
prestressing force because of the additional elastic shortening of the concrete caused by
the following stressing of other elements (EN 1992-1-1 cl. 5.10.5.1 & eq. (5.44)).
In section 2.2.2 it was stated that the maximum steel stress allowed after anchorage of the
prestressing steel (EN 1992-1-1 cl. 5.10.3) is the minumum value of 0,75 fpk and
0,85 fp0,1k.
During stressing, it is allowed to use an increased maximum stress to compensate for the
stress losses due to wedge set, friction and elastic deformation of the concrete. EN 1992-
1-1 cl. 5.10.2.1 restricts the stress to the minimum value of 0,80 fpk and 0,90 fp0,1k. When
the jack has an inaccuracy of less than 5%, it is allowed to overstress to 0,95 fp0,1k.
However, the NL National Annex prohibits overstressing.
In section 2.2.1 it was already mentioned that, according to EN 1992-1-1 cl. 5.10.2.2, the
minimum concrete strength fcm(t) required to be allowed to start stressing the tendons
(stepwise per individual tendon) is 50% of the minimum required compressive strength
for full prestressing as given in the ETAG certificate of the prestressing system. From
this minimum value fcm(t) to the strength given in the certificate, the prestressing force of
a tendon can be stepwise increased from 30% of the maximum tendon force up to the full
100% of maximum tendon force.
Additionally, during stressing, the concrete compressive stress caused by prestressing and
other loads should not exceed 0,60 fck(t). In the case of pre-tensioned steel, this value may
be increased to 0,70 fck(t).
Permanent compressive concrete stresses of over 0,45 fck(t) are allowed provided that the
designer accounts for non-linear creep (EN 1992-1-1 cl. 5.10.2.2 (5)).
The design should meet the requirements from EN 1992-1-1 cl. 7.3.1 with regard to
allowed steel stresses or crack widths in SLS (see Chapter 9).
In table 2.5, the most important characteristic values of a number of frequently used
prestressing steel types is collected. The data are from EN 1992-1-1. For more detailed
information on prestressing steel, reference is made to EN 10138.
More information about prestressing systems available in the Netherlands, such as the
cross-sectional area and the composition of the different prestressing elements, the
required ducts, the minimum bending radii, etc, are included in the appendices.
19
18
17
16
15
14
13
12
11
10
suspension column suspension column
9
2A
01
02
03
Compared to reinforced tensile columns, prestressed columns have the advantage that
they can remain uncracked during the service life, and, as a result, deform less. In
reinforced columns, a high reinforcement ratio would be required to limit crack widths.
Figure 3.2 schematically shows a comparison between a prestressed and a reinforced
column subjected to axial tension and provided with one steel bar.
Phase 1: The tensile strength of the concrete has not yet been reached. The column is
uncracked and has a high stiffness.
Phase 2: The column is cracked. The deformation of the steel bar is reduced by the
concrete in between the cracks, which is still active by bond (tension
stiffening; transfer of forces between steel and concrete). The stiffness of the
column decreases with ongoing crack formation.
Phase 3: The reinforcement has reached its yield strength and is in a limit state. The
bond forces transferred by the concrete between the cracks are now of minor
importance since the deformation of the steel bar is governed by the position
where yielding occurs.
1 2 3
Nu 3
2
Nrp
N reinforcement
without bond
1 2
reinforcing bar
Nrep (tension stiffening)
prestressed bar
Nrs
1
rep,p rep,r
uncracked cracked yielding
steel
Fig. 3.2 Axial force - mean strain behaviour of an axially loaded tensile member that
consists of reinforcement only, of reinforced concrete and of prestressed
concrete, respectively
l
c s
l
Fig. 3.3 Centrically reinforced tensile member loaded by an axial compressive force
l
N c Ac Ec (3.1a)
l
l
N s As Es (3.1b)
l
F N c Ns (3.2)
From eqs. (3.1) and (3.2) the following relations for the forces are obtained:
1
Nc F c F (3.3a)
1 e s
e s
Ns F s F (3.3b)
1 e s
Es As
where: e ; s
Ec Ac
Before cutting the tendons, the tensile force in the steel is Pmax and the concrete is free of
stresses. After the release of the force on the abutments, the force becomes a compressive
force that is taken over by the concrete element that contains prestressing steel. In
accordance with eq. (3.3), a part Pmax (1 e p ) of the compressive force Pmax is carried
by the concrete and a part e p Pmax (1 e p ) by the steel.
For the resulting forces Nc in the concrete and Pm0 in the prestressing steel it holds:
1
Nc Pmax
1 e p
N c Pm 0 (H 0!)
e p 1
Pm 0 Pmax Pmax Pmax
1 e p 1 e p
The compressive force in the concrete is in equilibrium with the tensile force in the
prestressing steel. This horizontal force equilibrium follows from the requirement that,
once the tendons are cut, no external force is exerted on the concrete element.
abutment
In section 2.3 it was already mentioned that for prestressing with post-tensioned steel,
two possibilities exist to check if the required force in the prestressing steel is reached.
The tendon force can be read from a load cell or the corresponding elongation can be
measured (see fig. 3.5). Both measurements can also be done at the same time. The
measured elongation is then used to verify the force measurement.
When measuring the elongation, its components should be accounted for. To illustrate
this, the axially prestressed element from fig. 3.6 is analysed.
lp lc
sheath
fixed-end
anchorage
PFmax
po
Initially, the tendon is positioned free of stress in its ducts. Then the tendon is stressed
until the prestressing force Pmax is applied. A hydraulic jack loads the prestressing steel.
The reaction force is exerted on the concrete. Afterwards, the tendon is anchored and the
duct is injected with grout.
Pmax l
lp
Ep Ap
Pmax l
lc
Ec Ac
The distance over which the tendon is pulled out of the concrete is the sum of both
displacements:
l l
l lp lc Pmax
E A E A
p p c c
When measuring the elongation, one should be aware that the tendon must be taut before
any stress can be built up (fig. 3.7). Therefore, in most cases, initially a small part of the
prestressing force is applied (pulling taut), and subsequently, the elongation is measured
for the remaining part of the force to be applied.
no force in
tendon
In the example, prestressing the member did not introduce any additional loading other
than the prestressing load itself. In practice however, often a part of the selfweight and
static load is activated during prestressing. The effect of this action should be included in
the calculation of the elongation. This will be demonstrated by an example with a
suspension column.
The structure from fig. 3.8 is designed such that before prestressing, the selfweight of the
column (Fg1) and the suspended floor slab II and its cross beam (Fg2) are carried by a
temporary structure, e.g. formwork. The prestressing force has to be designed such that
after prestressing a compressive force Fq is present in cross-section A-A. The question is
how to determine the elongation of the prestressing steel to achieve this.
During prestressing, the column shortens and the temporary supporting structure is
gradually unloaded. As soon as Pmax Fg1 Fg2 , the prestressing steel has taken over the
total support reaction. At that moment, the elongation of the prestressing steel is:
l1
F g1 Fg2 l
Ep Ap
Fpo
Pmax
A A
floor I
suspension
column Fg1
floor Fg2
formwork
auxiliary construction
The concrete stress σc in cross-section A-A now is equal to zero. To have a compressive
force Fq in cross-section A-A, the force in the prestressing steel has to be increased by Fq.
The corresponding elongation of the prestressing steel is:
Fq l
l2
Ep Ap
Fq l
l3
Ec Ac
Note that the force in the concrete column is Fg1 + Fg2 (tension) from the selfweight of
the structure plus Fg1 + Fg2 + Fq (compression) from prestressing. The resulting force in
the concrete is Fq (compression).
The total force in the prestressing steel is: Pmax Fg1 Fg2 Fq
The accompanying pulled-out length of the prestressing steel is:
l l
lp l1 l2 l3 Pmax Fq (3.4)
Ep Ap Ec Ac
Fg Fg1 Fg2
selfweight of hanging column
permanent load from floor to be suspended = 500 kN
Fq 300 kN
variable load on the floor
tension side
1500
suspension column
3000
Ac
500
suspended floor
The tension column is constructed from C28/35 concrete and is assumed to be prestressed
with post-tensioned steel. Prestressing takes place after the concrete has reached its 28-
day strength, so that the allowed (initial compressive) stress is σc = -0,6 · 28 =
-16,8 N/mm2 (EN 1992-1-1 cl. 5.10.2.2 (5)). The prestressing steel type is Y1030H. At
ULS the prestressing steel has to be designed to resist the forces caused by a fundamental
load combination with a load factor 1,2 for the permanent load and a load factor 1,5 for
the variable load.
To make sure that the column remains uncracked (high stiffness to limit deformations), a
compressive stress of 2,0 N/mm2 is required when the column is subjected to the
maximum loading according to SLS.
It can be assumed that the time-dependent prestress losses are 15%. The immediately
occurring losses (namely the elastic shortening of the concrete and steel as a result of the
prestressing sequence (the stresses induced by prestressing other elements), the set at the
anchorage and the friction losses) can be neglected.
The following results are required:
the cross-sectional area of the column;
the deformation;
the elongation of the prestressing steel relative to the concrete.
At maximum loading at SLS a (compressive) stress of 2,0 N/mm2 has to be present. The
most unfavourable situation occurs after the stress reduction caused by the time-
dependent prestressing losses, i.e. at a prestressing force of Pm(∞) = 0,85 Pm0 where Pm0 is
the initial prestressing force (after anchoring) and Pm(∞) is the working prestressing force
that is present after all time-dependent prestress losses are taken into account. Because
the immediately occurring losses (elastic shortening resulting from the stressing of
following tendons, the wedge set at the anchorage and the friction losses) are supposed to
be zero, it holds Pm0 = Pmax.
The density of the concrete is assumed to be 25 kN/m3 . The selfweight of the
column Fg2 = Ac l , where l is the column length (3 m).
Requirement:
A concrete compressive stress of at least -2 N/mm2 in SLS at full loading.
If it is assumed that full SLS loading occurs after the time-dependent prestress losses
have developed, the compressive stress requirement is:
Pm ( ) Fg1 Ac l Fq
c 2 N/mm 2
Ac
0,85 Pm 0 500 Ac 3 25 300
c 2000 kN/m 2
Ac
where Ac is the cross-sectional area of the column in m2 and Pm(∞) is the working
prestressing force in kN.
Requirement:
A not too high concrete compressive stress directly after anchoring the tendons.
The maximum concrete compressive stress occurs when the floor is not yet loaded and
the prestressing force is still at its maximum. According to EN 1992-1-1 cl. 5.10.2.2 (5)
the initial (compressive) stress σc after anchoring has to be limited to -0,6 fck =
-16,8 N/mm2. From this it follows:
Pm 0 Fg1 Ac l
ci 16,8 N/mm 2
Ac
or:
Pm 0 500 Ac 3, 0 25
16800 kN/m 2
Ac
The eqs. (3.5) and (3.6) determine the possible combinations of Ac and Pm0 (fig. 3.10).
The minimum value of the cross-sectional area is Ac = 0,0307 m2 at a minimum initial
prestressing force Pm0 = 1016 kN (Fig. 3.10).
A column cross-section of 0, 25 0, 25 m ( Ac 0, 0625 m 2 ) is chosen.
0.3
Ac (m2)
0.2
0.1
n 3.6
equatio
0.0
5
3.
n
tio
-0.1 qua
e
-0.2
-0.3
-0.4
1000 2000
Fpi (kN)
Pm0 (kN)
The corresponding minimum initial prestressing force follows from expression (3.5) and
reads: Pm0 1094 kN .
The associated cross-sectional area of the prestressing steel follows from the smallest
value of the maximum stresses allowed, see table 2.5:
It is assumed that losses from set at the anchor and friction can be neglected. Therefore,
σpm0 = σp,max.
It is found:
Further, it should be checked whether the column has sufficient reserve against failure
(ULS), i.e.:
Ap
1, 2 Fg 1,5 Fq
1, 2 500 1,5 300 103 1246 mm 2
f pd 843
Note that it is assumed that the bars reach the fpd = fp0,1k / s design strength, not the fpk / s
value, which is reached at a much higher strain. This is a conservative approach.
The SLS is governing. Four Dywidag bars Ø26 mm (see table 2.1) are selected with
Ap 4 531 2124 mm 2 . The duct dimension is Ø32/38 mm (inner/outer diameter).
The suspension column was designed in a global analysis in which the SLS condition at
t = and ULS were checked. The results are now checked in a detailed analysis in which
characteristic consecutive moments in time are looked at. Moreover, the ducts are grouted
after anchoring. This implies that loads exerted before grouting are carried by a column
that has different properties than a column with grouted ducts.
The following data is used, which has been obtained from the global design method:
Prestressing steel
Total cross-sectional area: Ap 2124 mm 2
Concrete
Gross cross-sectional area: Ac 62500 mm 2
Net cross-section before grouting: Ac1 Ac 4 Aduct 62500 4 14 322 59,3 103 mm 2
Net cross-section after grouting: Ac2 Ac Ap 62500 2124 60, 4 103 mm 2
The stress in the prestressing steel at SLS reaches its maximum when the live load is
assumed to be activated directly after stressing and grouting. The time-dependent
prestressing losses are then still zero. According to eq. (3.3), the contribution of the steel
in carrying an additional external load applied to the column is:
e p
Np Fq p Fq
1 e p
where:
e Ep Ec 205 103 / 32, 0 103 6, 41 (Ec from EN 1992-1-1 table 3.1)
and
p Ap Ac2 2124 / 60, 4 103 0, 035 .
The remaining part of the tensile force Fq (i.e. Np – Fq) is carried by the concrete and
results in a reduction of the compressive stress from prestressing:
Fq Fq 300 103
c c (1 p ) (1 0,183) 4,1 N/mm 2
Ac2 Ac2 60, 4 10 3
6, 41 0, 035
N p p Fq Fq 0,183 300 54,9 kN
1 6, 41 0, 035
N p54,9 103
p 26 N/mm 2
Ap 2124
Thus, the maximum value of the initial stress in the prestressing steel should not exceed:
This demonstrates that it might not be correct to prestress the bars in the column up to
their maximum allowable stress, because after applying the live load to the suspension
column, the stress in the bar increases and its maximum allowable stress (773 N/mm2)
might be exceeded.
Time t = 0
The axial compressive force in the concrete has its maximum value when the bars are
prestressed up to Pm0 and anchored. Now, only the permanent load Fg acts on the column.
In this state, with ducts that are not yet injected, the concrete (compressive) stress in
cross-section A-A of fig. 3.9 is:
c1
Pm0 Fg
1587 500 0, 0625 3, 0 25 103
18,3 N/mm 2 16,8 N/mm 2
Ac1 59,3 103
It appears that the bars are stressed to a too high level; the bars should not be prestressed
up to the maximum stress allowed in the concrete. The maximum initial prestressing
force follows from:
c1 16,8 N/mm 2
Pm0 500 0, 0625 3, 0 25 103
Pm0 1501 kN
59,3 103
To summarise:
The bars can be stressed to σp,max = 773 N/mm2 to meet the requirement on the
allowed initial prestressing steel stress before and directly after anchoring.
The initial prestressing steel stress has to be reduced to 747 N/mm2 when the live load
is assumed to be present already at t = 0 (i.e. before time-dependent losses occur).
The initial prestressing steel stress has to be further reduced to maximum 707 N/mm2
to meet the requirement on the initial concrete compressive stress caused by
prestressing.
After grouting of the prestressing ducts, the full live load is applied to the column, which
results in the previously calculated stress changes in the concrete and steel (see eq. (3.3)):
Fq Fq 300 103
c c (1 p ) (1 0,183) 4,1 N/mm 2
Ac2 Ac2 60, 4 10 3
Fq 300 103
p p 0,183 26 N/mm 2
Ap 2124
1 e p
where: c and p
1 e p 1 e p
The stress in the concrete:
Time t
It is assumed that the prestressing force reduces to 85% of its initial value, which is
caused by losses such as shrinkage, creep and relaxation (see Chapter 6). This implies
that the concrete is unloaded by a force 0,15 Pm0 0,15 1501 225 103 N , which follows
from internal force equilibrium in the column. In the unloaded state (i.e. no live load), the
concrete (compressive) force is:
After applying the live load, the concrete (compressive) stress is reduced. The stress
change was calculated before (+4,1 N/mm2). The resulting concrete stress is:
The stress in the steel changes to 85% of its initial value plus the stress increase caused
by the external load:
It can be concluded that during the service life of the structure, the conditions with
respect to the maximum stresses are satisfied.
It appears that the loading sequence influences the initial prestressing force allowed: The
structural engineer has to judge whether the live load can be present already directly after
grouting or will it be present after (part of) the time-dependent prestress losses have
occurred.
Immediately after applying the prestressing force Pm0 and the activation of the permanent
load, the compressive strain in the concrete is:
c1
Pm0 Fg
1501 500 0, 0625 3, 0 25 103
0,53 103 0,53 0 00
Ec Ac1 32 103 59,3 103
Fq 300 103
c2 c 0,817 0,13 103 0,13 0
00
Ec Ac2 32 10 60, 4 10
3 3
During the service life, a time-dependent deformation occurs due to shrinkage and creep
of the concrete, as well as relaxation of the prestressing steel (see Chapter 6). The
elongation of the prestressing steel decreases. By approximation, it holds:
The elongation of the prestressing steel relative to the concrete at t = 0 is fully analogous
to eq. (3.4). One finds:
lp 5000
lp Pm0 c1lc 1501 103 0,53 103 3000
Ep Ap 205 10 2124
3
17, 2 1, 6 18,8 mm
Note that it is assumed that the elastic shortening of the two floors can be neglected.
The example demonstrates that knowledge about the steps in the construction process and
the loading sequence are important to accurately estimate the stresses that develop in the
prestressing tendons and the concrete. A structural engineer should list possible loading
sequences and judge whether they are realistic.
In the case of the suspended column, it appeared that the time at which the live load
might be present, might be governing for the initial prestressing force allowed.
The position of the prestressing steel in concrete structures subjected to bending is mainly
determined by two factors:
1 - The concrete is not allowed to crack at all or just to a small degree, under the most
unfavourable load combinations. Therefore, limiting values are given for the maximum
tensile stress or crack width (also see Chapter 9). This usually leads to a low position of
the prestressing steel at midspan of a beam and a high position over the supports (the
position follows the concrete areas loaded in tension).
2 - It must be possible to anchor the prestressing force adequately. The anchor plates
must have prescribed minimum dimensions to transfer the high anchor forces. In
addition, a minimum mutual distance between the anchors should be applied to prevent
the development of unfavourable stress concentrations. Therefore, it is necessary to
spread out the tendons toward the beam-ends. In many cases, the anchorage issue is
governing in the choice of the type of prestressing tendon. It is often required to enlarge
the end of the concrete element in order to accommodate the required anchors ("end
block") (fig. 4.1).
Fig. 4.1 Prestressing anchors at the end of a beam (side view and longitudinal cross-
section)
As a result of these two factors, the tendons are usually draped (prestressing with post-
tensioned steel) or harped (prestressing with pre-tensioned steel), see also figs. 1.7 and
1.8.
Because of this layout, curvature pressures or forces are generated, which load the beam.
This is illustrated in fig. 4.2.
c.a.
Pm sin Pm sin
Pm cos Pm cos
Fig. 4.2 Calculation of the curvature pressure in a beam with a draped tendon:
a. actual prestressing load on a beam
b. curvature pressure on tendon element having length R·dφ
c. relation between curvature radius R, beam length l and drape f
d. tensile force exerted on the tendon and curvature pressure exerted by
the tendon on the beam
e. schematised prestressing load on the beam
In the example from fig. 4.2, the tendon is a circle segment with radius R and segment
height (also called the drape) f (fig. 4.2a). It is assumed that the prestressing force is
constant over the full length of the tendon. Figure 4.2b shows a small part of the curved
tendon. From the equilibrium equations it follows:
Pm
qp R d Pm d qp (4.1)
R
From fig. 4.2c the following relation between the radius R and the drape f is obtained:
f R R 2 14 l 2
This yields:
f 2 2 f R R 2 R 2 14 l 2 f 2 2 f R 14 l 2
In most cases R >> f, which implies that the term f 2 can be neglected. The relation then
reduces to:
l2
R (4.2)
8f
8 Pm f
qp (4.3)
l2
At the ends of the beam, the force Pm is introduced at an angle θ (also see fig. 4.2d). The
force Pm can be decomposed into a vertical force Pm sin and a horizontal force
Pm cos . In most cases, the angle θ is small. For a beam with a span of 25 m and a drape
f of 0,80 m, the angle θ is 7,5º (0,13 rad). Then it follows that Pm cos 0,991 Pm and
Pm sin 0,13 Pm . It is a good approximation to use a horizontal force introduced at the
beam-ends equal to Pm.
A similar reasoning holds for the curvature pressure qp Pm R . The vertical component
is almost equal to Pm R . The horizontal component is relatively small. It has to
compensate for the difference between the horizontal force Pm cos at the beam end and
the horizontal force Pm at midspan, which holds for a tendon that experiences no friction.
For an angle 7,5º the difference between both is 0, 009 Pm , which implies that the
contribution from the horizontal component of qp can be neglected.
In the analysis of the problem, a total prestressing load as shown in fig. 4.2e has to be
taken into account.
EN 1992-1-1 (and also the Dutch code NEN 6720 (VBC 1995)) uses this approach in
which the prestressing is introduced as a load (EN 1992-1-1 cl. 5.10.1). In the literature,
this method is known as the “equivalent prestressing load method”.
In this method, the stresses in a cross-section are directly calculated using a cross-
sectional bending moment from prestressing equal to the prestressing force times the
tendon eccentricity in the cross-section considered. Figure 4.3a shows a case in which a
tendon has an eccentricity ep at midspan of a statically determinate beam.
The concrete stresses at the bottom and top side of the beam follow from:
Pm Pm ep
cb (bottom side) (4.4)
Ac Wcb
Pm Pm ep
ct
Ac Wct (top side) (4.5)
centroidal axis
Exactly the same values can be found using the “equivalent prestressing load method”.
To demonstrate this, the following loads are introduced: an axial compressive force Pm
having an eccentricity ep0 at the beam-ends and an upward uniformly distributed
curvature pressure (also see fig. 4.3b):
Pm 8 Pm f
qp
R l2
At midspan of the beam, the bending moment comes from the eccentric axial force at the
beam-ends and the uniform upward load:
1 1 8 Pm f 2
M Pm ep0 qp l 2 Pm ep0 l Pm ep0 Pm f
8 8 l2
M Pm ep
At first sight, the cross-section method seems to be easier to use than the equivalent
prestressing load method. However, this only holds for statically determinate beams. If
the beam is statically indeterminate, the bending moment line from the curvature
pressures no longer directly follows from the eccentricity of the tendon. This is caused by
the statically indeterminate nature of the structure: curvature pressures deform the
structure. Since the structure is statically indeterminate, these deformations cannot
develop freely. As a result, additional reaction forces and/or bending moments develop.
The actual bending moment line from the load follows from statics.
It is, therefore, strongly advised to always use the equivalent prestressing load
method, even in case of a statically determinate structure.
From the previous discussion, the concept “prestressing” can be summarised as follows:
“PRESTRESSING IS PRELOADING”
For the case of prestressing with post-tensioned steel (both with and without bond), the
prestressing force is the force Pm0 exerted on the beam after anchoring.
In prestressing with pre-tensioned steel the prestressing force is the force Pmax exerted on
the abutments just before cutting (unloading) the tendons (in most cases strands). The
previously mentioned concept becomes clear when regarding the term “unloading” as
“loading” by a force having an opposite sign.
yield
stress
prestressed concrete
from prestressing
reinforced concrete
Figure 4.4a shows an axially prestressed element on a prestressing bed. The tendons are
still anchored to the abutments and have a force Pmax. The stress in the prestressing steel
then is p,max Pmax / Ap and the stress in the concrete c 0 .
Due the release of the prestressing force between the abutments, the force is transferred
to the concrete element. Following eq. (3.3) it holds:
Pmax
c
Ac 1 e p
e p Pmax
p
Ap 1 e p
Now, the reinforced element from fig. 4.4b is considered, in which the steel cross-section
As is equal to Ap of the prestressed element. In the unloaded state s 0 and c 0 .
When this element is loaded by an external force Pmax, this force is distributed over the
concrete and steel cross-sections in the same way as in the prestressed element. In both
cases, the stress increase is the same. The only difference is the initial stress p0 in the
prestressing steel, and as a result of this, the reserve in strength relative to the yield
strength (fig.4.4c). This finding is important for the understanding of the concept of
prestressing.
A statically determinate beam is prestressed using a draped tendon (fig. 4.5). The
anchorages of the tendon are in the centre of gravity (centroidal axis) of the cross-section.
Fig. 4.5 Loads on a statically determinate prestressed beam with a parabolic tendon
profile
The moment in the cross-section at midspan caused by the prestressing load (fig. 4.5a) is:
1
M p qp l 2
8
where qp is the uniformly distributed prestressing load from the curvature pressure. Using
relation (4.3) it can be found that:
18P f
M p m2 l 2 Pm f
8 l
This result is as expected because, according to the cross-section method, as shown in fig.
4.5b, the drape f at midspan is exactly equal to the eccentricity of the tendon.
In the past it was often stated that crack formation was not permitted at SLS loading. In
that case, three boundary conditions must be met when determining the prestressing force
and the position of the centre of gravity of the prestressing in the cross-section:
In the unloaded state (i.e. loads acting at the time of tensioning or release of
prestress), the stress in the pre-tensioned compression zone (top side of the structure
from fig. 4.5) should not exceed a specified tensile stress.
At the same time, the stress in the pre-compressed tension zone (bottom side of the
structure from fig. 4.5) should not exceed a specified compressive stress which is
based on fck(t) (the characteristic compressive strength of the concrete at time t when
it is subjected to the prestressing force ) (EN 1992-1-1 cl. 5.10.2.2 (5)).
At maximum load, the tensile stress in the pre-compressed tension zone should not
exceed a certain value.
A distinction is made between full, limited and partial prestressing. In case of full
prestressing, no tensile stresses should occur at all. In case of limited prestressing, small
tensile stresses are allowed, whereas partial prestressing enables crack formation in SLS.
A discussion on partial prestressing is given in Chapter 12.
zct
kt
kern area e
x kb
F
zcb
The effects of the geometry of the cross-section are now analysed. For that purpose, the
concept “kern area” has to be explained first (see fig. 4.6).
The kern area of a cross-section is the part of the cross-section where an axial
compressive force has to be applied to avoid the occurrence of tensile stresses in the
cross-section.
The position of the bottom point of the kern area is determined by making the stress in
the top fibre of the cross-section equal to zero:
F F e zct
ct 0
Ac Ic
I c zct Wct
e kb (4.6)
Ac Ac
For the position of the upper point of the kern area, the stresses at the bottom side of the
cross-section are set to zero. Solving a similar expression for the eccentricity results in:
I c zcb Wcb
e kt (4.7)
Ac Ac
ct
F
kt
kb
F
cb
a b c
Fig. 4.7 Stress distribution from an axial compressive force for different lines of
action
The stress distribution can simply be determined as soon as the positions of the outer
points of the kern area are known. When the axial force is applied in the upper kern area
point, the stress cb at the bottom fibre is zero (fig. 4.7a), and when the load is applied in
the lower kern area point, the stress ct at the top side of the cross-section is zero (fig.
4.7b).
For loads applied at arbitrary positions in the cross-section, the stress distribution can be
calculated using the position of kern area points. In case of a compressive force F having
an eccentricity e relative to the centre of gravity, it is found:
F e kt F e kb
cb ; ct
Wcb Wct
In the following analysis the option “fully prestressed concrete” is considered. The
statically determinate beam from fig. 4.3 has a tendon drape ep. From the “prestressing is
preloading” approach, it followed that the midspan bending moment from prestressing is
Pmep. The axial compressive force Pm at centroidal axis (neutral axis) level is present over
full beam length. In the midspan cross-section the combination of the force Pm and the
bending moment Pmep is replaced by a force Pm that has an eccentricity ep (fig. 4.3a &
fig. 4.7).
The governing situation for the top fibre of the beam occurs when it has its permanent
load G only. Because tensile stresses are not allowed, the requirement is:
Pm M g Pm ep M Pm ep kb
ct g 0
Ac Wct Wct Wct Wct
or:
Mg
e p kb M g Pm e p kb (4.8)
Pm
The governing situation for the extreme bottom fibre occurs at full loading G + Q:
Pm M g M q Pm ep M g M q Pm ep kt
cb 0
Ac Wcb Wcb Wcb Wcb
or:
Mg Mq
ep kt M g M q Pm ep kt (4.9)
Pm
Equations (4.8) and (4.9) provide the information to determine the zone in which the
prestressing force must be applied, see fig. 4.8.
kt
kb
admissible area
Fig. 4.8 Area in which the line of action of the prestressing force should be situated
From this analysis, a number of interesting conclusions can be drawn with respect to the
effectiveness of the shape of the cross-section.
The minimum moment required from static loading G follows from equation (4.8):
M g,min Pm ep kb
This implies:
“The greater the distance between the prestressing steel and the lower kern area point,
the greater the bending moment capacity for static loading”.
From eqs. (4.8) and (4.9) the maximum bending moment allowed from the live load Q is
obtained:
M g M q Pm ep kt (from equation (4.9))
This implies:
“The greater the height of the kern area in the cross-section, the greater the moment
capacity for live loads”.
In fig. 4.9, a number of shapes having the same cross-sectional area are compared. The
prestressing force is kept constant.
1 2 3 4
Fig. 4.9 Data for several cross-sections; bending moment resistance for a prestressing
force Pm = 1/2 Ac σcp
From this comparison, it can be concluded that cross-sections that contain one
compression flange and have a high permanent load (e.g. selfweight) can sustain only
relatively small variable loads. On the other hand, symmetrical cross-sections are well-
suited for resisting low permanent loads combined with high variable loads (for example
crane beams).
The most obvious conclusion from fig. 4.9 is that it is the most economic to maximise the
height of the section. However, the following aspects make that this is often not the case:
There is a lower limit for the thickness of the web, because of the required minimum
concrete cover on the web reinforcement (e.g. stirrups and longitudinal web
reinforcement) and the required shear capacity (of the concrete struts loaded in
compression).
The beams should have a specific minimum stiffness in the lateral direction to
prevent lateral torsional buckling to occur (See EN 1992-1-1 cl. 5.9 1(P): ‘Lateral
instability of slender beams shall be taken into account where necessary, e.g. for
precast beams during transport and erection, for beams without sufficient lateral
bracing in the finished structure etc’).
In beams without a bottom flange and having a small web thickness, the prestressing
elements must be arranged vertically. This raises the centre of gravity of the
prestressing, which has an unfavourable effect on the effectiveness of the
prestressing (e.g. reduction of the drape).
In order to minimise the total construction costs, it might be desired to keep the total
structural height as small as possible.
For partially prestressed concrete different criteria hold. Therefore, for partial prestressed
concrete, especially at low levels of prestressing, some caution is required with respect to
the previous conclusions.
Fig. 4.10 Prestressing with pre-tensioned steel: situation before (a) and after (b) the
tensioning
Before the prestressing force is transferred to the beam, the stress in the steel is σpmax =
Pmax / Ap. The element shortens due to the transfer of the prestressing force to the
element. In the steel a compressive force is generated. This force is assumed to be -ΔPel
and must be added to the prestressing force Pmax. This implies that a force Pmax - ΔPel is
exerted on the net concrete cross-section. This causes the following shortening of the
concrete at the steel level:
Pel x
lp
Ap Ep
Because it should hold that lc lp , the extra compressive force transferred to the steel
is:
e p f
Pel Pmax (4.10)
1 e p f
where:
Ac ep2 E Ap
f 1 ; e s ; p
I c Ec Ac
1
Pmax Pel Pmax (4.11)
1 e p f
Example
A hollow-core slab is prestressed by seven strands Ø12,5 mm (Ap = 93 mm2/strand),
Y1860S7 (fpd = 1522 N/mm2; table 2.5). The concrete strength class aimed at is C35/45,
but is only C20/25 at the time when the prestressing is applied to the concrete.
The dimensions of the hollow-core slab are given in fig. 4.11.
Geometrical data:
From table 2.5 it can be read that the maximum steel stress in SLS must be
σpm0 < 1395 N/mm2. During prestressing it is allowed to use σpmax = 1488 N/mm2. A
further increase to σpmax = 1590 N/mm2 is allowed by EN 1992-1-1 cl. 5.10.2.1 (2)
provided that the jack force is accurately measured (note: The NL National Annex does
not allow this increase).
The increase of the steel stress to σpmax can compensate for the stress losses caused by
wedge set when anchoring the strands, by friction and elastic shortening of the concrete.
Young’s modulus of concrete strength class C35/45 having reached a class C20/25
strength (EN 1992-1-1 table 3.1 & eq. (3.5)):
0,3 0,3
f (t ) 20 8
Ecm (t ) cm Ecm 34 10 29,9 10 N/mm
3 3 2
f cm 35 8
Es 195 103
e 6,52
Ec 29,9 103
Ap 651
p 4,34 103
Ac 150 10 3
and:
p0 1 0, 048 pmax
p0 1395
pmax 1465 N/mm 2
1 0, 048 1 0, 048
So, it is not possible to make full use of the stress increase to σpmax = 1488 N/mm2 since
this will cause a too high σp0 (restricted to 1395 N/mm2).
The initial prestressing force and the losses due to the elastic shortening are:
1
Pmax Pel Pmax 0,952 954 908 kN
1 e p f
The maximum compressive stress in the pre-compressed tensile zone (bottom fibre,
assumed to be at a distance h / 2 from the centroidal axis):
1 ep 12 h 3 1 60 12 200
cb Pmax Pel 908 10 6
13,9 N/mm 2
150 10 690 10
3
Ac Ic
When applying the prestressing force to the slab (during the release of the stress in the
strands at the abutment), the selfweight of the slab will be activated; the slab bends
upwards because of the eccentric prestressing load, and, as a result, will be supported
only at its ends. From that moment on, the bending moment caused by the selfweight is
activated.
However, at the ends of the slab the effect of the selfweight is zero, so that the previously
calculated concrete stress applies at these sections.
EN 1992-1-1 cl. 5.10.2.2 (5) requires that the initial concrete compressive stress is less
than:
c 0, 70 f ck (t ) 0, 70 20 14, 0 N/mm 2
where fck(t) is the characteristic cylinder compressive strength at the moment the
prestressing is applied to the concrete. This requirement is just met. In most cases, the
slabs are prestressed quite quickly after casting (for example at C20/25). After continuing
hardening of the concrete, a higher strength class will be reached (for example C45/55).
When the prestressing force is exerted on the post-tensioned steel, this force does not act
on the composite cross-section, because there is no bond yet between the prestressing
steel and the concrete. Due to the prestressing, the beam will deform, which activates the
selfweight as a load, see fig. 4.12.
centroidal axis
formwork formwork
When at the manometer of the prestressing jack a force Pmax is indicated, the stresses in
the concrete at midspan, at the bottom and at the top are:
Pmax M g Pmax ep
Bottom: cb (4.12a)
Ac Wcb Wcb
Pmax M g Pmax ep
Top: ct (4.12b)
Ac Wct Wct
where Wcb and Wct are the section moduli of the concrete cross-section relative to the
bottom and top fibre respectively, and Mg is the moment caused by activating the
selfweight of the structure by upward bending.
To make sure that the intended stress in the prestressing steel is actually applied, not only
the force in the jack is registered, but also the elongation of the tendon relative to the
concrete surface is measured.
To calculate the expected elongation of the tendon, the beam is supposed to be divided
into small segments (fig. 4.13a,c). The x-axis is horizontal; the s-axis follows the tendon
profile.
1
Ec
Fig. 4.13 Calculation of the elongation of the tendon relative to the concrete surface
It should be noted that prestressing losses occur as a result of friction between the tendon
and the duct, see fig. 4.13b. The calculation of these losses is dealt with in section 4.4.3.
The elongation of the tendon relative to the concrete surface is the elongation of the
prestressing steel itself (lp > 0) and the shortening of the concrete (lc < 0) at the level
of the tendon. When introducing the (+)-sign for the elongation of the steel and the (-)-
sign for the shortening of the concrete, the relative displacement of the tendon (lp,rel) is:
The shortening of the concrete at the level of the prestressing steel in an arbitrary
segment (fig. 4.13c), is:
1 ep2 ( s ) 1
cp Pmax ( s ) cos( ( s ))
Ac I c Ec
In most cases the contribution of the third integral from eq. (4.16) is relatively small and
can be ignored:
1 1
l
1 1
lp,rel
E A E A Pmax ( x) dx
E A E A mm
P l (4.17)
p p c c 0 p p c c
If the elongation of the prestressing steel during prestressing is not checked, this may lead
to damage to the structure, as will be shown in the following example.
During the prestressing of a beam (fig. 4.14), flexural cracks occurred at the top, although
the manometer indicated that the intended level of prestressing was not yet reached. It
appeared that the damage was caused by a rather flexible formwork system. For that
reason the full selfweight of the structure was not activated immediately during
prestressing, but was gradually activated. This resulted in a too high upward load and, as
a result, unexpected high tensile stresses at the top of the beam, which resulted in the
observed cracks.
If one would have calculated and controlled the elongation of the prestressing tendon, this
deviation would already have been discovered during prestressing. It would be clear that
the measured elongation relative to the concrete (eq. (4.16)) did not correspond with the
magnitude of the prestressing force, as indicated by the manometer (a relatively too high
elongation). This can be explained using eq. (4.16): the last integral should have been
small, but was relatively high since a part of the selfweight was not active yet (a too small
Mg).
This example underlines the meaning of the proposition:
“Prestressing is deforming”
4.4.2 Prestressing losses caused by elastic deformation during the prestressing of a
member with more than one prestressing tendon
When applying a prestressing force, not only the beam will shorten, but also the tendons
that were tensioned before. So, before injecting the ducts, already a part of the tendons
have lost part of their stress. This is illustrated by fig. 4.15.
Fig. 4.15 Cross-sectional and side view of a prestressed beam with four prestressing
tendons
The beam shown in fig. 4.15 has a cross-sectional area Ac = 1,0·106 mm2, a modulus of
elasticity Ec = 30·103 N/mm2 and a prestressing force per tendon Pm = 875 kN (Ap =
902 mm2).
When discussing the relative elongation of the prestressing tendon during prestressing
(eq. (4.16)), it was already pointed out that the last integral can be ignored in most
practical situations. Therefore, the effect of bending of the beam will be ignored in this
case and only the axial deformations are taken into account. If a prestressing tendon is
stressed to a force Pm, the shortening of the concrete is:
Pm l
lc
Ec Ac
The total shortening lpel of the first prestressing tendon caused by the prestressing of the
following (n - 1) tendons is:
Pml
lpel (n 1)
Ec Ac
Ep Ap
Pel (n 1) Pm (4.18)
Ec Ac
The total loss for n tendons is calculated by summing up the losses of all individual
tendons. Note that each tendon has a different number of tendons prestressed after the
tendon considered is stressed: n - 1 ; n – 2; n – 3 ................... n - n (= 0). The mean value
of all these individual factors is (n – 1) / 2; their sum is n (n – 1) / 2. The total force loss
is:
n (n 1) Ep Ap
Pel Pm (4.19)
2 Ec Ac
In EN 1992-1-1 cl. 5.10.5.1 & eq. (5.44) the average loss per tendon is presented:
(n 1) Ep Ap
Pel Pm
2 Ec Ac
In this example four prestressing tendons are applied. The losses, calculated using eqs.
(4.18) and (4.19) are listed in table 4.1.
Table 4.1 Losses per prestressing tendon caused by stressing of the following
tendons
tendon loss Pm losses
[kN] [%]
1 15,8 1,8
2 10,5 1,2
3 5,3 0,6
4 - -
total 31,6 0,9
During the prestressing of post-tensioned steel, especially for curved tendons, a part of
the prestressing force is lost due to friction between the tendon and the duct. Therefore,
the prestressing force is not constant along the length of the structure, see for example
fig. 4.13b.
Figure 4.16a shows a beam with a curved tendon profile. A detail is shown in fig. 4.16b.
Pm ( x)
qp
R
Using a friction coefficient for the interface between tendon and duct, the frictional
force is:
Pm ( x)
qp
R
In fig. 4.16c the same part of the tendon having a length R θ is shown again. From
equilibrium considerations it follows that:
Pmax ( x)
Pμ ( x) R
R
Pμ ( x)
P max ( x)
it follows that:
Using the boundary conditions Pmax(x) = Pmax and θ = 0 for x = 0 it is found that:
anchor
Pm ( s)
Pm ( s ) Pm ( s )
Pm ( s )
duct
tendon Pm ( s ) Pm ( s )
Substituting eq. (4.21) in eq. (4.20) and replacing Pmax(x) by Pmax(θ) results in:
Pmax ( )
ln
Pmax
or
In this equation θ is the total angular rotation (in radians) of the tendon, calculated from
the location of the prestressing jack. The coefficient of friction depends on the
prestressing system. For prestressing strands 0,10 – 0,24; for bars 0,33 – 0,65
(EN 1992-1-1 table 5.1).
Measurements of the frictional losses in practice demonstrated, however, that eq. (4.22)
mostly results in too low values for the losses. This is caused by the actual angular
rotation being larger than assumed in the calculation.
Therefore, a distinction is made between the “intentional” angular rotation θ and the
“unintentional” angular rotation k, which is caused by the so-called “Wobble-effect”.
The Wobble-effect is caused by unintended curvatures of the prestressing duct, and the
deflection of a “straight” tendon between its supports, see fig. 4.17. Prestressing systems
used in practice have an angle k due to the Wobble effect that is in the range of 0,005 –
0,01 rad/m (EN 1992-1-1 cl. 5.10.5.2 (3)).
The values for and k follow from documentation of the prestressing system used.
Ø duct
at prestressing:
stressing the tendon in the duct
Ø duct
tendon stretches
Fig. 4.17 Origin of the Wobble-effect by the angular rotation k according to Theile
[4.1]
In general, the following expression is used to determine the frictional losses (EN 1992-
1-1 eq. (5.45)):
Here k is the angular rotation (in rad) per unit of length caused by the Wobble-effect, and
x is the distance from the location of the prestressing jack to the cross-section considered.
Sometimes it is doubted whether combining the intended angular rotation θ and the
unintended angular rotation k is correct. From a physical point of view it is probable that
there is a certain interaction between k x and θ. For large intended angular rotations θ the
unintended angular rotation k x should be almost 0.
Therefore Walter, Utescher and Schreck [4.2] proposed not to use the sum of θ and k x
but to use only the largest one. That would imply:
Equation (4.25b) was found to be only valid for curved tendon profiles which meet the
condition:
1
R (4.25)
k
where R is the radius of curvature of the tendon and k is the angular rotation due to the
Wobble-effect (in rad/m).
When anchoring using wedges (fig. 4.18), some wedge slip (wedge set) occurs when
releasing the jack. The wedges slip until the anchorage force is fully developed. This
results in a loss of prestress as shown in fig. 4.18. The slip of strands or wires is generally
in the range of 5 – 15 mm.
In this expression both ΔP2 and lw are unknown. The stress reduction in the prestressing
tendon caused by friction is assumed to be linear. Therefore, the following expression
holds:
pμ,m
P2 2 Ap lw (4.27)
x
where Δσpμ,m/Δx is the mean stress reduction in the tendon caused by friction over the
length lw. (Note: Units, for example, N/mm2/mm). Expression (4.26) can now be
rewritten as:
Ap pμ,mlw2
w
Ep Ap x
Fig. 4.18 Loss of prestress due to (wedge) slip of the anchor at the tensioned side
Now, the length over which wedge set influences the tendon force is known:
w Ep
lw
pμ,m
x
The force loss at the position of the anchor is calculated using expression (4.27). The
expression derived can not only be used when the friction is constant over the length lw,
but also in case of a profile with a changing radius of curvature or kinks. In the latter
case, it is mostly required to iterate to find a solution: after an initial estimation of length
lw, Δσpμ,m/Δx can be calculated and, as a result of this, lw is recalculated. If the calculated
and the estimated values are almost identical, the required distance lw is found.
moduli (W). Its implication for practice is illustrated for the cross-section shown in fig.
4.19.
The cross-section is prestressed with tendons consisting of strands Ø12,9 mm, having a
total cross-sectional area Ap = 3600 mm2. The diameter of the prestressing ducts is
60 mm. If only the concrete in the cross-section is regarded, the following cross-sectional
properties are calculated:
For the composite cross-section, having a concrete strength class C35/45, the following
results are obtained:
I i I c Ac ( zi zc )2 ( e 1) Ap ( zi zp ) 2 I p
7,87 102 0, 443 0, 0212 4,97 3, 60 103 0,5562 0 8, 44 102 m 4
In such calculations mostly only the properties of the concrete section are used, not
taking into account the influence of prestressing and reinforcing steel. This is a
conservative simplification. In the case considered, the concrete stresses calculated using
the composite (fictitious) section (caused by M) are about 7% smaller than those
calculated using the concrete cross-section only.
The fictitious cross-section is not often used in design, since all dimensions must be
known in detail. It might be used in a detailed design check, using a standard computer
program.
I-shaped cross-sections are often applied for smaller spans, especially if the variable load
is significantly higher than the permanent load. As a result of the large height of the kern
area, relatively high variable loads can be sustained at a relatively small prestressing
force.
Box girder beams (fig. 4.20), which can in fact be calculated as being T- or I beams, are
mostly applied for large to very large spans. Furthermore, they are especially suited for
carrying torsional moments.
Rectangular beams and slabs are only applied in the case of small spans.
Other requirements and demands can, of course, also have impact on the choice of the
optimal shape of the cross-section.
Fig. 4.20 Examples of box girder bridges with one (a) and two (b) boxes and practical
dimensions
Figure 4.21a shows the cross-section of a traffic bridge. The shape of the precast beams
makes that no scaffolding/formwork has to be used when casting the concrete in-situ.
After casting, a solid slab bridge is obtained.
Figure 4.21b shows a partially precast bridge deck. For carrying the load, the lower
flanges are actually not required. Nevertheless they are applied to provide the viaduct
with sufficient resistance against transverse collision loads caused by passing vehicles,
see fig. 4.22.
The ratio between the depth of the cross-section and the span (h / l) ranges from l/20 (for a
solid cross-section) to 1/35 (for an I-beam or a box girder).
cast-in-situ
bridge deck concrete prestressed T-girder
b
possible
impact load
bearing block
A bridge beam has a span of 21,5 m. Apart from the selfweight qg of the structure itself,
there is an additional permanent load qg = 2,5 kN/m’ (from the bridge deck and the
parapets) and a variable load qq = 12 kN/m’. The concrete strength class is C35/45 and
the environmental class is XD3. In this example full prestressing is applied, which
implies that in SLS no tensile stresses in the concrete should occur.
qq = 12 kN/m
qg = 2,5 kN/m
zct
zcb
c 0, 6 f ck c 0, 6 35 21 N/mm 2
Note that if the concrete compressive stress is permanently over 0,45 fck, non-linear creep
has to be accounted for.
The maximum stresses allowed in the prestressing steel Y1860 are given in table 2.5:
Initial stress after anchoring pm0 < 1395 N/mm2
Stress during prestressing p,max < 1488 N/mm2 (no overstressing)
In this example it is assumed that the prestress losses caused by elastic shortening of the
concrete are compensated for by overstressing. This will, therefore, not be dealt with in
this example. Prestress losses caused by shrinkage and creep of concrete are also not
dealt with. Creep and shrinkage are discussed in Chapter 6. Tendon friction is first
neglected; it will be introduced in detail in a following section.
As a start, it is assumed that the total time-dependent losses are 20%. This implies that
the prestress is pm∞ = 0,8 · 1395 = 1116 N/mm2. The prestressing force is now
determined such that the requirement on "no concrete tensile stress" is met. This criterion
is mostly governing in the design. Afterwards, it will be checked whether the conditions
with regard to the concrete compressive stresses are met as well. Furthermore, it will be
checked whether the maximum tensile stress in the prestressing steel meets the specified
limits.
In this example the main item is the control of the concrete stresses in this statically
determinate beam. In the first step, prestressing is regarded as an external load, see fig.
4.5a.
Fig. 4.25 Loads as a result of prestressing a tendon that has a parabolic profile
The load qp introduces a parabolic bending moment line, having a maximum at midspan:
M p 18 qp l 2 P ep
The concrete stresses at top (ct) and bottom (cb) are now calculated:
Pm Pm ep
ct
Ac Wct
Pm Pm ep
cb
Ac Wcb
selfweight:
ME,g = 1/8 · 8,2 · 21,52 = 474 kNm
static load:
ME,g = 1/8 · 2,5 · 21,52 = 144 kNm
total permanent loading:
ME,g = 474 + 144 = 618 kNm
variable load:
ME,q = 1/8 · 12,0 · 21,52 = 693 kNm
Note:
It is assumed that the variable load is caused by a crowd of people. Then, ψ1 = 0,75 (EN
1990 defines the frequent and quasi-permanent load parts of a variable load).
For the stress in the concrete at the top fibre of the cross-section, the construction stage is
governing. In that case only the selfweight of the beam and the full initial prestressing
force Pm0 (directly after prestressing, so without time-dependent prestress losses) are
present. The condition is:
0
0,328 0, 092 0, 092
474
ep 0, 28 (4.28)
Pm0
At the bottom fibre, the highest tensile stress occurs if the maximum moment is applied
on the beam, whereas at the same time, the time-dependent prestress losses have fully
developed. At a working prestressing force Pm,∞ = 0,8 Pm0 the result is:
Pm, Pm , ep M E,freq
cb 0
Ac Wcb Wcb
or
0,8Pm0 103 0,8 Pm0 10 ep 1138 103
3
cb 0
0,328 0, 0703 0, 0703
which results in the condition:
1423
ep 0, 21 (4.29)
Pm0
In fig. 4.26 the relation between the prestressing force Pm0 and the eccentricity ep is
shown. The shaded area shows combinations of Pm0 and ep allowed from the stress
requirements.
In fig. 4.24 it can be estimated that the maximum possible eccentricity of the prestressing
steel to the centre of gravity of the cross-section is ep,max = 0,44 m. This extra limit is also
shown in the diagram using a horizontal dashed line. This is an additional limit to
possible combinations of ep and Pm0.
Fig. 4.26 Relation between the initial prestressing force and the eccentricity ep
From economy reasons the smallest possible prestressing force is chosen. In combination
with ep = 0,44 m, eq. (4.29) results in Pm0 ≥ 2189 kN and eq. (4.28) gives Pm0 ≤ 2963 kN.
For an initial prestress p0 = 1395 N/mm2 (Y1860) the required cross-sectional area of the
prestressing steel Ap = 2189·103 / 1395 = 1569 mm2. If four tendons are applied, each
tendon should have a cross-sectional area of at least 1569 / 4 = 392 mm2. In this example
tendons, each of them containing 5 strands Ø13,0 mm (Ap = 500 mm2) are chosen (total
Ap = 2000 mm2). The duct diameter is 45 mm.
Figure 4.27 shows the four tendons in the cross-section. The tendons are positioned such
that their centre of gravity is 125 mm from the outermost concrete fibre. Mean tendon
eccentricity is ep = 441 mm (zcb = 0,566 m).
ep =
441
Now, the amount of prestressing steel is known and stress checks are performed. The
calculations were performed not taking into account friction losses. Tendon friction is
now dealt with in detail.
Already during prestressing, the stress in a tendon varies along the length of the beam.
This is caused by the friction losses. In the calculation a parabolic profile of the
prestressing tendons was assumed. Furthermore, it was assumed that the prestressing
force is applied at the centre of gravity of the gross cross-section at the beam ends. The
drape of the tendons then is f = 0,441 m (the ep from fig. 4.27). The layout of the tendons
can be described by the relation (fig. 4.28):
4f 2
y 2
x 0, 0038 x 2 where f is the drape.
l
If the beam is prestressed from one side, the prestressing force varies according to the
expression (EN 1992-1-1 eq. (5.45)):
where x is the distance from the position where the jack is applied (active end) to the
cross-section considered. It is assumed that the coefficient of friction is = 0,15 and that
the angular rotation caused by the Wobble effect is k = 0,005 rad/m’.
The inclination of the tendon can be calculated in any cross-section from its layout,
according to
dy 8f x
2 x (4.31)
dx l R
x
R
For a number of points along the beam the magnitude of the prestressing force has been
calculated using the equations (4.30) and (4.31) (see table 4.2)
Table 4.2 Calculated prestressing force in a number of points along the beam
x θ k Pm0(x) / Pm0(x=0)
0 0 0 1,00
1
/4 l 0,041 0,0269 0,99
1
/2 l 0,082 0,0538 0,98
3
/4 l 0,123 0,0806 0,97
l 0,164 0,1075 0,96
At an initial stress (after anchoring) p0 = 1395 N/mm2 at the position of the anchor, the
steel stress at midpan is 0,98 · 1395 = 1367 N/mm2.
This prestressing force meets the requirement found before, namely that:
2189 kN ≤ Pm0 ≤ 2963 kN.
The mean prestressing force is Pm0 = 2734 kN. During prestressing only the moment
caused by the selfweight ME,g = 474 kNm is activated. The compressive stresses at the
top and bottom fibre at midspan are checked:
Top:
Pm0 Pm0 ep M E,g
ct
Ac Wct Wct
Bottom:
First, the concrete stress at the outer tensile fibre at maximum load for t = is calculated.
The most critical situation occurs when the stress losses due to shrinkage and creep have
developed. To simplify the calculation, it is assumed that the force Pm,∞ = 0,8 Pm0 =
2187 kN acts directly after prestressing. So, the cross-sectional forces due prestressing
(P) and selfweight (qg) must be calculated using the properties of the “pure concrete”
cross-section, whereas those for the additional static load (qp) and the variable load (Q)
should be calculated using the fictitious composite cross-section.
As a result, for the tensile stress at the bottom fibre of the cross-section it is found that:
Because of the small differences between the gross (pure) concrete cross-section and the
fictitious (composite) cross-section, in calculations often only the gross cross-section is
used.
Once the area within which the prestressing force has to be applied is known, the tendon
profile can be determined, see fig. 4.29. The limitations for the tendon layout usually
follow from the requirement that a certain tensile stress (here ct = 0 N/mm2) is not
exceeded.
Fig. 4.29 Area within which the tendon profile has to be to prevent concrete tensile
stresses to occur
At a section x where MEd,g(x) and MEd,freq(x) act, the following conditions (requirement:
no concrete tensile stresses) apply:
M E,g ( x) Wct
ep ( x) (4.30)
Pm0 Ac
and
M E,freq ( x) Wcb
ep ( x) (4.31)
Pm, Ac
As an example, with the aid of the equations (4.30) and (4.31) it is found, that at x = 1/4 l,
where ME,g = 3/4 · 474 = 356 kNm and ME,freq = 3/4 · 1138 = 854 kNm, the following
condition must be met:
0, 25 m ep ( 1 4 l ) 0, 42 m
Wcb W
ep (0) ct
Ac Ac
For x = 0 this implies that the prestressing force should be applied between the most
upper and the most lower kern area points, see also eqs. (4.6) and (4.7). This refers, of
course, to the resultant of the prestressing tendons. One often denotes this as the
“fictitious” or “imaginary” tendon. In practice, the tendons are spread as much as
possible at the beam ends, in order to position their anchorages appropriately in the
structure.
Fig. 4.30 Prestressing operation for a large concrete beam (trough girder) in the
Hanzelijn (railway line) near Kampen (NL)
4.8 Literature
4.1 Theile, V.: The influence of prestressing on the SLS in prestressed concrete
structures. Dissertation, TU Darmstadt, Germany, 1986 (in German)
4.2 Walter, R., Utescher, G., Schreck, D.: Preliminary assessment of prestress losses
due to deformation restraint. Deutscher Ausschuss für Stahlbeton, No. 282, 1977
(in German)
Subsequently, the tendon layout as shown in fig. 5.2 is considered. This layout shows
discrete changes of direction (kinks) at the supports, whereas in between the supports (in
the spans) the cables are straight.
The prestressing tendon exerts forces on the beam only where its direction is changed
(kinked), or where it is anchored. This implies that, in the case considered, the
prestressing loads are directly transmitted to the supports. The support reactions
counterbalance the lateral prestressing forces. Therefore, the effect of prestressing in the
beam is the same as for the case of centrically prestressing shown in fig. 5.1: the concrete
is uniformly stressed in compression over the full height of its cross-section.
Now the beam as shown in fig. 5.3, having a curved tendon profile, is regarded. Note that
also in this case, there are kinks at the intermediate supports.
If the layout of the tendon would have been like the one from fig. 5.2 (linear tendon parts
with kinks at the intermediate supports; dashed lines in fig. 5.3a), there would be no
bending moments in the beam as a result of prestressing.
However, the tendon profile is not linear and kinked, but curved and kinked. The loads
from the curved tendon profile parts in between the supports are uniformly distributed
equivalent prestressing loads qp:
8 Pm f
qp
l2
The figures 5.3b and 5.3c show, qualitatively, the loads on the tendon and on the
concrete. The magnitude of the uniformly distributed prestressing loads on the beam are
shown in fig. 5.3d (points loads at anchors and kinks; uniformly distributed loads in the
spans).
Fig. 5.3 a. Beam with curved and kinked profile of the prestressing tendon
b. Forces exerted on the prestressing tendon
c. Forces exerted on the concrete
d. Practical representation of the equivalent prestressing load
In practice, it is not possible to choose the layout of the prestressing tendon such that
there are abrupt changes of direction (kinks), like the ones shown in fig. 5.2 and fig. 5.3.
For the curvature of the prestressing steel a minimum radius is prescribed, in order not to
damage the prestressing steel. This inevitable local anti-curvature of the prestressing steel
causes a downward load. This load (qp = Pm / R) acts only over a small length, but is
relatively high due to the high curvature (1 / R) of the tendon. In the design stage, the
local downward load at the supports is often not regarded, but it cannot be ignored in an
accurate calculation.
The equivalent prestressing load, corresponding to the layout of the prestressing tendon
shown in fig. 5.4a, is represented in fig. 5.4b. Note that the actual layout is used; the
kinks at the intermediate supports are replaced by downward curved parabolas.
Fig. 5.4 Equivalent prestressing loads for upward and downward curvatures of the
tendon
The equivalent prestressing load method can also be applied in the case of non-prismatic
members. When the system line of the structure (the line through the centres of gravity of
the subsequent cross-sections) has a nonlinear layout, additional forces are generated,
resulting from the axial force in the cross-section. As a result of the angle between two
subsequent cross-sections (cross-sections system line), the axial force has a component
perpendicular () to the system line, see fig. 5.5 (q R dφ in fig. 5.5a and 2 N sin(1/2φ) in
fig. 5.5b).
q
N
N
c.a. N
N
c.a.
d
N.d = q.R.d
q.R.d d 2.N.sin 12
q= N
R
(a) (b)
Fig. 5.5 Load perpendicular to the system line of a non-prismatic structure, as a result
of a normal compressive force (c.a. = centroidal axis)
At a kink in the system line, a force 2 N sin(1/2φ) (fig. 5.5b) occurs in the direction of the
bisectrice of the corner between the adjacent parts of the system line.
The vertical component of this force is 2 Pm sin(1/2α) cos(1/2α) = Pm sinα; the horizontal
force component is Pm (1 - cosα)
In the case of a curved system line, a load arises, which is inversely proportional to the
radius R of the system line (qp = Pm / R) (fig. 5.6c,d).
If such a structure is not calculated using a frame analysis program (which automatically
generates forces provided that the structural geometry is correctly schematised, see
section 5.2) such loads are dealt with in the same way as the equivalent prestressing load
following from the tendon curvature.
prestressing tendon
system line
As an example, the structure sketched in fig. 5.7a is now discussed. The following load
cases are distinguished:
1. Equivalent load qp as a result of the curvature of the prestressing tendon (fig. 5.7b).
2. Two vertical forces F, one vertical force 2F and two horizontal forces Pm (1-cosα)
caused by the axial compressive force Pm as a result of a kink (angle α relative to the
horizontal axis) in the system line (fig. 5.7c).
3. Moments Pm ep as a result of the eccentric introduction of the prestressing load at the
ends of the structure (fig. 5.7d).
4. A load Pm divided in a horizontal (centric) and vertical force component (fig. 5.7e).
Up to now, the effect of tendon friction was not regarded. In a more accurate calculation
this should be regarded as well and the equivalent loads resulting from prestressing and
kinks and curvatures in the system line have to be adapted. In most cases, however, the
results are sufficiently accurate if the mean value of the prestressing force is used.
For an accurate determination of the layout of the tendons, the calculation of the friction
losses, the elongation of the tendon at prestressing and the equivalent load by
prestressing, it may be useful to express the tendon profile (layout) using a mathematical
function. To make the calculation of the tendon layout and its derivatives as simple as
possible, a parabolic cable layout is chosen, typically for beams and slabs.
R1
R1
Pm(1-cosα)
Pmsinβ
Pmcosβ Pmcosβ
In section 5.1 it was shown in which way the prestressing load and the forces caused by a
curved or a kinked system line, can be applied as loads on a statically (in)determinate
structure. For the structures shown in figs. 5.6 and 5.7 it will be shown in this section
how those forces must be dealt with when a frame-analysis program is used.
The first example refers to the structure shown in fig. 5.6a. In this example the friction
losses are provisionally left unconsidered. In fig. 5.8a it is shown how the system line of
the structure can be schematised for a frame-analysis. Since account has been taken of the
kinked shape of the systemline, the prestressing load should only be applied at the ends of
the beam (at the anchors), see fig. 5.8c.
system line
stiff connection
The second example is the structure shown in fig. 5.6c. Also in this case friction losses
are not considered. In fig. 5.9b it is shown how the system line of this structure can be
stiff connection
The third example concerns the structure shown in fig. 5.7. The friction losses will now
be considered. The schematisation of the system line is shown in fig. 5.10b, and is similar
to that of fig. 5.8b. The force Pm in the prestressing steel will be gradually reduced by the
effect of tendon friction. The development of the prestressing force Pm is represented in
fig. 5.10c. The load on the structure as a result of tendon friction and wedge set is
represented in fig. 5.10d. Figure 5.10e shows the prestressing loads at the beam ends. The
prestressing loads from the curved tendon profile are given in fig. 5.10f.
The influence of the friction (fig. 5.10d) is not only reflected by the distribution of the
equivalent prestressing load perpendicular to the structure, but also as a uniformly
distributed axial load along the structure. Also in this example it is not necessary to apply
the apparent force separately at the kinks in the system line.
system line
distributed load in
longitudinal direction
from friction and wedge
set
Fig. 5.10 Example of a structure having a kinked system line and a curved tendon
profile; the influence of friction losses
Pm 3
Pm e p l1 l2
( Pm H )l 2
Ml M 2 l2
1
1
1
R (5.2)
2 EI1 EI1 EI1 EI 2 24 EI 2
where I1 and I2 are the moments of inertia of the column and the beam, respectively, and
R is the radius of the curvature of the prestressing tendon.
l2
Furthermore R 2 (eq. (4.2)), where f is the drape of the tendon (fig. 5.11a).
8f
R
a l2 b
f
2
1 1
l1
c d
PFmp Pm
Fp
M M
ep ep
d2
H H
d1
After combining the equations (5.1) and (5.2), a relation is obtained for the ratio between
H and Pm:
3 2 f 3ep 2
H 2 d2 2d 2
(5.3)
Pm 3
3 2
4 2
2 8 3 4
where:
For = 0,075, f / d2 = 0,7 and ep / d2 = 0,3 the diagram shown in fig. 5.12 is derived,
where the ratio H / Pm is presented as a function of d2 / d1 and l2 / l1.
For d2 / d1 > 1,0, the prestressing losses in the beam due to the action of the columns is
less than 20%. However, the equivalent prestressing load, which is a result of the tendon
curvature, is not influenced. If in the case of very stiff columns only a small part of the
prestressing load would act as axial compression, whereas the equivalent prestressing
load by the tendon curvature would still be able to play a major role in carrying the loads
on the beam (fig. 5.13).
Fig. 5.13 Action of the prestressing as an equivalent load on the beam of a frame
having very stiff columns
“Creep” can be defined as the increase of the deformation with time under a sustained
constant load (fig. 6.1a).
“Shrinkage” is the shortening of the concrete occurring without the influence of any load
(fig. 6.1b), which is caused by the drying of the material (the chemical shrinkage
occurring during the hardening of the concrete is not considered).
“Relaxation” is closely related to creep. However, in case of relaxation, the deformation
of the material is kept constant and the initially present stresses decrease with time (fig.
6.1c).
c
= constant
deformation
due to creep
a co
elasticdeformationby
elastic deformation by
applying
applying a load
load
t (time)
c
b
deformation due
to shrinkage
t (time)
c
decreaseofoftension
degrease compression
due to relaxation
c l = constant due to relaxation
t (time)
s
t (time)
Fp
b
t (time)
instantaneous support
displacement (t = t 0)
c
t = t
d
counter-acting moment
Mp t = t0
Fig. 6.2 Some examples of the effects of the long-term behaviour of concrete
Figure 6.2b shows a beam which is prestressed without bonding. The prestressing
element can be considered as a spring.
* When shrinkage occurs, the beam will shorten.
* In the compressive zone extra shortening due to creep has to be added.
* By this shortening, the tendon will be released, by which the exerted force on the
concrete will decrease.
Since the prestressing has to limit the magnitude of the tensile stresses, respectively the
width of the cracks in the serviceability limit state, or has to balance a fixed part of the
load, it is necessary to account for this loss in prestressing when designing the structure.
Figure 6.2c shows a beam on three supports. A deformation is forced upon the structure
by a sudden imposed displacement of the middle support. This will generate a bending
moment in the structure. However, in this case the time-dependent deformation has a
favourable effect: by relaxation of the concrete, this undesired moment will eventually be
reduced to a relatively small part of its initial value.
Creep of concrete is caused by the deformation of the gel structure and the capillary
stress of the chemically non-bonded water.
Development and ultimate magnitude of creep depend on the following factors:
The climate in which the structure is situated; especially the relative humidity and
temperature are important.
The development of the degree of hydration1, as well as the degree of hydration under
loading. These depend on the age of the concrete under loading, the applied type of
cement and the curing conditions.
The strength class of the concrete2.
The dimensions of the cross-section.
The duration of the loading.
Research has indicated that the creep consists of two components:
A reversible part (also indicated as the visco-elastic part, the retarded elastic part, or
the reversible creep).
A permanent part also indicated as the plastic part.
The development of these contributions as a function of time is quite different. The age of
the concrete hardly influences the reversible part εccr ( c = creep, r = reversible) and
already after a short period of time, its final level is reached. On the other hand, the
permanent contribution εccp ( c = creep, p = permanent) strongly depends on the age of
the concrete, and its ultimate value is reached just only after a long period. For the
simplification of the calculations, and because of the large scatter in the occurring creep,
a separate approach of both contributions is abandoned:
where:
1
The degree of hydration is the extent to which the cement is chemically bonded to water and hardened
cement paste is created.
2
The strength class of the concrete is determined to a large extent by the water-cement ratio (wcr) of the
concrete (amount of water per kg of cement).
Figure 6.3 shows creep curves for two situations, each with a different stress
development as a function of time.
c c
t t
ce ce
elastic strain t t
ccr ccr
reversible strain t t
ccp ccp
plastic strain t t
c (t) c (t)
total strain t t
c
cc (t ) (t , t0 ) (6.2)
Ec
where
Ec = 1,05 Ecm (EN 1992-1-1 cl. 3.1.4 (2); the Young's modulus at the origin of
the stress-strain relationship)
According to Appendix B.1(1) of EN 1992-1-1, the creep function φ(t,t0) can be written
as:
(t , t0 ) 0 c (t , t0 ) (6.3)
where:
βc(t,t0) is the factor that describes the development with time after loading
0,3
t t0
=
H t t0
16,8
β(fcm) is the factor for the effect of concrete strength = (6.4)
f cm
(fcm = fck + 8 N/mm2; EN 1992-1-1 table 3.1)
1
β(t0) is the factor for the concrete age at loading = (6.5)
0,1 t00,2
2Ac
h0 is the notional size of the member in mm =
u
Ac is the cross-sectional area
u is the perimeter in contact with the atmosphere
t is the age of concrete in days
t0 is the age of concrete at loading in days
t-t0 is the non-adjusted duration of loading in days
βH is a coefficient depending on the relative humidity and the notional size
1,5 1 (0, 012 RH)18 h0 250 1500 for f cm 35 N/mm 2
(6.7)
1,5 1 (0, 012 RH)18 h0 250 3 1500 3 for f cm 35 N/mm 2
0,5
35
3
f cm
At a low relative humidity, the potential difference between the moisture content in the
structure and its environment becomes larger, which causes the structure to dry more
quickly. Now two opposite effects are of importance: a strong reduction of the moisture
content in the structure increases the creep, while a structure with a lower moisture
content creeps less. In practice, the influence of the difference in moisture content
between the structure and the environment dominates. This is demonstrated by
expressions (6.6) and (6.7). A low RH combined with a small notional size results in a
high φRH according to eq. (6.6). A high RH combined with a large notional size results in
a high βH according to eq. (6.7) and, as a result, a low βc(t,t0).
With regard to the concrete strength class two aspects play a role. First, concrete having a
high strength also has a high stiffness compared with lower strength concrete. The higher
stiffness results in a smaller creep (eq. (6.4)). Second, concrete having a higher strength is
less permeable and therefore exhibits a slower drying process. This is beneficial from the
point of view of reducing creep deformation (eqs. (6.6) and (6.7)).
The finer the cement, the faster the hydration process. Also at elevated temperatures, the
hydration process develops faster. Concrete having a high degree of hydration exhibits
less creep if loaded. The temperature effect influences the age of the concrete at loading
to be used in eq. (6.5). The actual age is corrected for the hardening temperature by using
the so-called adjusted age of concrete:
9
t0 t0,T 1 0,5
2 t0,T
1,2
where:
α = -1 for cement class S
= 0 for cement class N
= 1 for cement class R
4000
n 13,65
273T ( ti )
t0,T is the temperature adjusted concrete age e
ti
i 1
T(Δti) is the temperature (in ºC) during the time period Δti
Δt i is the number of days where a temperature T prevails
Instead of performing a calculation, the designer can choose to read the expected creep
coefficient from EN 1992-1-1 fig. 3.1, see fig. 6.4. The curves can be used for a
temperature range from -40ºC to +40ºC and a RH between 40% and 100%. Two curves
are available; one for inside conditions (RH = 50%) and one for outside conditions (RH =
80%).
Fig. 6.4 Curves to determine the creep coefficient as a function of the concrete
strength class, the cement type, the notional size of the member and the age of
concrete at loading (EN 1992-1-1 fig. 3.1)
Fig. 6.5 Example to demonstrate the use of the curves (outside conditions; h0 =
500 mm; t0 = 30 d; concrete C30/37)
In case the concrete compressive stress at the time of loading t0 is more than 45% of the
cilinder compressive strength at that time (> 0,45 fck(t0)), non-linear creep has to be
accounted for. This might hold for precast prestressed concrete fully stressed at a
relatively low concrete strength. The creep coefficient then is (EN 1992-1-1 eq. (3.7)):
where:
6.2.2 Shrinkage
cs cd ca (6.8)
where:
εcs is total shrinkage
Drying shrinkage is the result of the drying of a member. Drying shrinkage can go on for
many years for structures having large dimensions.
cd (t ) ds (t , t0 ) kh cd,0 (6.9)
cd, kh cd,0
where
2Ac
kh is a coefficient that depends on the notional size h0 = of the cross-section,
u
see table 6.1 (EN 1992-1-1 table 3.3)
t ts
βds(t,t0) =
t ts 0, 04 3 h0
t is the age of concrete in days
t0 is the age of concrete at the beginning of drying shrinkage (normally after
demoulding or at the end of curing) in days
εcd,0 is the nominal unrestrained drying shrinkage (EN 1992-1-1 table 3.2), see
table 6.2
h0 [mm] kh
100 1,0
200 0,85
300 0,75
> 500 0,70
Table 6.2 Nominal unrestrained drying shrinkage εcd,0 [‰] for concrete with cement
CEM class N (EN 1992-1-1 table 3.2)
ca (t ) as (t ) ca ()
where
εca(∞) is the final autogenous shrinkage = 2,5 f ck 10 106 (EN 1992-1-1 eq.
(3.12))
βas(t) is the influence of the age of the concrete = 1 e 0,2 t
(EN 1992-1-1 eq.
(3.13))
t is the age of concrete in days
At a low RH, the structure dries faster, which speeds up the development of shrinkage.
On the other hand, the permeability of concrete having a higher strength is lower than the
permeability of a lower strength concrete, which slows down the drying process. Both
aspects are present in table 6.2.
Just as in the case of creep, the geometry of the cross-section, in terms of the notional
size h0, plays an important role: the larger the dimensions of the structure, the longer
shrinkage will take to fully develop and the smaller the shrinkage the designer has to
account for. This is clearly shown in table 6.1.
As shown, apart from time, the dimensions influence the development of shrinkage in
time. The larger the dimensions of a structure, the smaller the theoretical final value of
shrinkage to account for and the slower the process, see eq. (6.9). This is strictly speaking
not correct since the final value of shrinkage is independent of the dimensions of the
structure. However, since shrinkage might take decades to fully develop in a large
structure, it is regarded as appropriate to use this proposed method in design.
The final value of autogenous shrinkage strongly depends on the permeability of the
concrete and the amount of water available for the hydration process. Concrete having a
high strength has a low permeability on the one hand, but on the other hand it has a low
water-binder ratio and, as a result, exhibit a larger autogenous shrinkage. EN 1992-1-1
relates autogenous shrinkage primarily to the strength class. In practice, a low strength
concrete might also have a low permeability, for instance self-compacting concrete. The
autogenous shrinkage of such a type of concrete might be higher than the value from EN
1992-1-1. Long-term measurements (> 10 yrs) on the first large high strength concrete
bridge constructed in The Netherlands, the 2e Stichtse Brug, indicate that autogenous
shrinkage might reduce in time.
For many creep calculations it is not possible to use a single load that is constant with
time. In many cases, the stress gradually increases with the progress in construction. It is
also possible that in spite of a constant dead weight, the stress even decreases due to
creep of the concrete. Because reinforcement does not creep, a part of the force in the
concrete is transferred to the reinforcement.
For an arbitrary stress development, the creep deformation is described by relation (6.2).
For a discrete stress increment Δσi applied at time ti it holds:
i
cc (t ) (t , ti ) (6.10a)
Ec
(t, t )
1
cc (t ) i i (6.10b)
Ec i 0
Accordingly, the creep deformation Δεcc(t) at time t has to be determined separately for
each stress increment Δσi (see fig. 6.6).
t [time]
φ(t,ti)
φ(t,t0)
φ(t,t1)
t0 t1
t [time]
Fig. 6.6 Calculation of the creep caused by a number of successive load increments
For a continuously varying stress σ(τ), equation (6.10b) can be rewritten as:
t
1 d ( )
cc (t ) (t , ) d (6.11)
Ec d
0
The principle of superposition gives accurate results provided that the following four
conditions are met [6.4]:
The stresses during the service life have to be smaller than about 40% of the
compressive strength.
The total strain should not decrease due to a decrease of the load. Relaxation
however, i.e. the reduction of the stress under constant strain, is allowed.
The variations in relative humidity should not be too large.
No high stress increments should occur over the period after initial load application.
In this incremental calculation of the creep deformation, the full stress development as a
function of time should be accounted for. This usually results in very time-consuming
calculations, which can be performed only by a computer program. It must be kept in
mind that accurate calculations do not necessarily yield accurate results, because a
number of influencing factors (such as relative humidity and loading history) must be
estimated in advance. In order to avoid complicated calculations, approximation methods
have been developed.
6.3 Simplified methods for the calculation of the shrinkage and creep
behaviour of structures
Several approximating methods are available for the calculation of shrinkage and creep, a
number of which are:
The effective-modulus method (section 6.3.1).
The method of Dischinger (section 6.3.2).
The method of Trost (section 6.3.3).
The total strain of the concrete consists of the elastic deformation εce, the creep
deformation εcc(t), and the shrinkage deformation εcs(t):
c (t ) ce cc (t ) cs (t )
n n
i i
i 0
Ec
E
i 0 c
(t , ti ) cs (t )
n
i
c (t ) E
i 0 c
1 (t , ti ) cs (t )
(6.12)
For t it holds:
n
1 (, t )
1
c ( ) i i cs ( ) (6.13)
Ec i 0
n
Using φ∞ = φ(∞,t0) and i transforms equation (6.13) into:
i 0
1
c ( ) 1 cs ()
Ec
or:
c ( ) cs () (6.14)
Ec'
where:
Ec
Ec' (6.15)
1
Ec is also called the fictitious E-modulus of concrete.
This approach makes the calculation of the deformations much easier: the fictitious E-
modulus Ec is determined using eq. (6.15) and is used in the strain calculation as if it
were the normal E-modulus Ec.
A remarkable property of equation (6.14) is that the creep deformation is fully recovered
during unloading. This implies that it is assumed that the concrete behaves as an ideal
visco-elastic material and that the creep deformation is completely reversible.
Moreover, the creep factor used corresponds to the age of the concrete at first loading.
This implies that the continuing ageing of the concrete is not taken into account (ongoing
hydration), although the creep coefficient is smaller for loads applied at a later point of
time. Therefore, the creep is overestimated when using this calculation method.
Also the method developed by Dischinger uses a basic creep function, corresponding
with the time the first load is exerted on the structure, φ(∞,t0).
The influence of ageing of the concrete can nevertheless be accounted for. This is done
by adjusting the creep function for loads applied at a later point in time ti. The result is a
creep coefficient φ(ti,t0).
(t , ti ) (t , t0 ) (ti , t0 ) (6.16)
Figure 6.7 shows the strain development as assumed by Dischinger. The basic idea
behind it is as follows: the curve that presents the strains after applying a stress increase
Δσi at time ti is parallel to the curve obtained if this stress increment Δσi were applied in
the beginning.
However, this standard creep function does not result in a reliable estimation of the
influence of concrete ageing. Contrary to the effective E-modulus method that
overestimates creep at increasing concrete age, Dischinger's method underestimates
creep.
Compared with the effective E-modulus method, there are also some differences in case
of unloading. In Dischinger's method creep deformation is irreversible, as can be derived
from fig. 6.8.
actual behaviour
Fig. 6.7 Creep behaviour according to Dischinger in case of a stepwise increase of the
stress
actual behaviour
This method does not have the drawbacks of the two methods presented before.
Moreover, it results in a relatively simple calculation procedure. According to relation
(6.10b) it holds:
n
(t, t )
1
cc (t ) i i (6.17)
Ec i 0
For many structures, initially (t0) a relatively high stress σ0 is present, after which
relatively small stress increments Δσi(ti) follow (fig. 6.9).
i
0
t00 ti1
t
Adding the elastic deformation and shrinkage results in the total strain:
n
1
c (t ) 0 1 (t , t0 ) i 1 (t , ti ) cs (t )
(6.18)
Ec i 1
n
1
c (t ) 0 1 (t , t0 ) (t ) 0 i (t , ti ) cs (t )
(6.19)
Ec i 1
In order to obtain a simple equation, which can be applied in practice, Trost proposes to
change equation (6.19) into:
1
c (t ) 0 1 (t , t0 ) (t ) 0 1 (t , t0 ) cs (t )
Ec
1
c (t ) 0 1 (t , t0 ) (t ) 1 (t , t0 ) cs (t ) (6.20)
Ec
where Δσ(t) = σ(t) - σ0 and ρ is the so-called ageing coefficient. With the aid of equations
(6.19) and (6.20) it can be derived:
n
(t, t )
i 1
i i
(6.21)
(t ) 0 (t , t0 )
This is a considerable improvement, since it turned out that the value of ρ is between
rather narrow boundaries. Generally, a value ρ = 0,8 appears to be a good approximation
for normally loaded concrete structures with a creep factor 1,5 < φ(∞,t0) < 4,0.
This result shows that the total deformation due to elastic shortening and creep can be
calculated with:
Ec Ec 1 as fictitious modulus of elasticity for stresses that are present at t = t0.
Ec Ec 1 as fictitious modulus of elasticity for stresses that develop during
the creep process.
the stiff supports and the wall are filled with a low shrinkage mortar. It will be judged
whether the prestressing force applied decreases due to relaxation.
Expressions derived in section 6.3 can be used to determine the time-dependent decrease
of the axial compressive force. When using expression (6.20) from Trost, a relaxation
function can be derived.
wall jack
Since εc(t) - εc0 = 0, the following expression for the change of stress Δσ(t) is derived:
0 cs (t ) Ec
(t )
1
cs (t ) Ec / 0
(t ) 0 (t ) 0 1 (6.23)
1
The development of shrinkage and creep over time is described by the function given in
figure 6.11.
(t ) cs (t )
cs,
5y
Fig. 6.11 The assumed development of shrinkage and creep over time
After 90 d (about 50% of the end values of shrinkage and creep are reached):
φ(90 d) = 0,5 · 2,5 = 1,25
εcs(90 d) = 0,5·(-0,25 · 10-3) = -0,125 · 10-3
After substituting these values in eq. (6.23), the concrete stress is found to be:
1, 25 0,125 103 30 103 /(10)
c (90 d ) 10, 0 1 1,9 N/mm 2
1 0,8 1, 25
The development of the concrete stress as a function of time is presented in fig. 6.12.
-1,9 N/mm2
360
time [d]
The initial compressive stress of -10 N/mm2 is vanished after 360 days. It will turn into a
tensile stress after 360 d. This example demonstrates that prestressing a structure fixed at
two points is a very inefficient way of prestressing.
Time-dependent deformations also occur in steel subjected to the high stress levels that
are normally present in prestressing steel. Because the time-dependent deformations of
the concrete are small compared with the strains of the prestressing steel, a constant time-
independent deformation in the steel can be assumed. Relaxation will cause the stress in
the prestressing steel to decrease under this imposed deformation, which is almost kept
constant in time.
The relaxation behaviour of the prestressing steel is determined with tests, which last
1000 to 3000 hours (fig. 6.13).
log t [h]
Fig. 6.13 Relaxation of the prestressing steel as a function of the loading time
Class 2 having a somewhat lower stress loss and referring to heat-treated wires
and strands. Stress loss after 1000 h in a standard relaxation test is about 2,5% →
ρ1000 = 2,5%. The following expression applies (EN 1992-1-1 eq. (3.29)):
0,75 (1 )
pr 9,1 t
0, 66 1000 e 105
pi 1000
Class 3 which refers to bars. Stress loss after 1000 h in a standard relaxation test
is about 4,0% → ρ1000 = 4,0%. The following expression applies (EN 1992-1-1
eq. (3.30)):
0,75 (1 )
pr 8,0 t
1,98 1000 e 105
pi 1000
The final stress loss caused by relaxation can be calculated using t = 500000 h.
Example
When comparing the prestressing steel relaxation expressions from EN 1992-1-1 with the
ones used in the Dutch code NEN 6720 (VBC 1995), it might be concluded that in EN
1992-1-1, time-dependent stress losses caused by shrinkage and creep do not influence
the relaxation loss. In NEN 6720 relaxation loss is directly reduced since concrete
shrinkage and creep reduce the steel stress level, which in turn reduces the relaxation loss
since the latter depends on the stress level. However, EN 1992-1-1 accounts for this too,
but afterwards, when all time-dependent steel stress losses are added. Relaxation loss is
then reduced by 20% to 80% of the calculated value (EN 1992-1-1 eq. (5.46)).
A column is loaded by an axial force as shown in fig. 6.14. Immediately after loading, the
force distribution between steel and concrete can be determined using equation (3.3).
It then holds:
N N e s
N c0 ; N s0
1 e s 1 e s
N
c0 s0
Ec Ac 1 e s
N
The concrete exhibits creep and shrinkage deformation. According to Trost it holds:
c t c0 1 t , t0 c t 1 t , t0 cs
s t s0 s t
The concrete will reduce its initial load by shrinkage and creep and transfer it to the steel.
The changes in the forces over the course of time are indicated by ΔNc(t) and ΔNs(t).
From force equilibrium it follows:
N c t N s t 0
c t s t
Since co = so, the following should also hold for the time-dependent components of the
total concrete and steel strains:
c t c0 s t
N c t
c t c0 c0 t , t0
Ec Ac
1 t , t0 cs (6.24)
N s t
s t (6.25)
Es As
cs e s t , t0
N c t N s t N c t Ec Ac (6.26)
t , t0 1 e s 1 t , t0
The effects of shrinkage and creep on the development of the stresses are demonstrated
by an example. For the column as shown in fig. 6.13, the following data are assumed:
The initial load at the cross-section is N0 = -640 kN. Directly after applying this load it
holds (also see eq. (3.3)):
1 1
N c0 N0 640 503 kN
1 e s 1 6,83 0, 04
e s
N s0 N 0 137 kN
1 e s
c0 N c Ac 8, 0 N/mm 2
s0 e c0 54, 6 N/mm 2
After 5 years, the concrete stress is reduced to 49% of its initial value. In fig. 6.15, the
stress reduction as a function of time is presented for several reinforcement ratios. In all
cases, an initial concrete compressive stress of -8,0 N/mm2 is assumed.
ρs = 0%
ρs = 2%
ρs = 4%
ρs = 6%
ρs = 8%
t [years]
Fig. 6.15 Stress reduction in the concrete due to shrinkage and creep of a symmetrically
reinforced cross-section loaded in compression
When detailing structures, creep should be considered as well. Figure 6.16 shows an
example of a column clad with natural stone tiles. When creep is not taken into account
when cladding the column (for example by applying wide joints), the tiles may spall off
after some time.
Fig. 6.16 Spalling of a natural stone cladding due to creep of the column
reinforcing steel As. The centre of gravity of the joined steel cross-section is at a distance
eps from the centroidal axis of the total cross-section. The change in steel stress due to
shrinkage and creep of the concrete is calculated.
concrete stress
at t = 0
The initial strain of the concrete (c) at the centre of gravity of the steel (ps) is:
cps (0)
cps (0) (6.27)
Ec
where Ec is the modulus of elasticity of the concrete at the time of loading.
If the concrete can deform without any restraint from the steel, the following additional
deformation from creep (c) and shrinkage (s) (= c + s) is expected in the concrete at the
level of the prestressing and reinforcing steel:
cps (0)
cps,c+s t (t , 0) cs (t ) (6.28)
Ec
where εcs(t) is the absolute value of the shrinkage developed at time t.
In reality, the concrete and the steel are joined together, and the steel will react with a
force ΔF(t) to restore compatibility (fig. 6.17c). This force results in a strain change
Δεps(t) in the steel (Ap + As):
F t
ps t (6.29)
( As Ap ) Eps
where Eps is the modulus of elasticity of both prestressing and reinforcing steel (for the
sake of convenience, no distinction between the two is made).
From equilibrium it follows that the same force is exerted to the concrete in opposite
direction. The force develops gradually since creep and shrinkage develop in time. This
implies that the deformation of the concrete caused by the force ΔF(t) is also influenced
by creep. Since the force is not constant in time, but gradually increases, the coefficient ρ
(from Trost) is used.
The result is the following change in concrete strain at the centre of gravity of the steel:
F t F t eps2 1
cps
A Ic
c Ec
F t F t eps2 1
The result is: cps t (6.30)
A
c Ic Ec /(1 (t , 0))
cps (0)
cps,c+s t (t , 0) cs (t ) cps t ps t
Ec
cps (0) F t F t eps2 1 F t
(t , 0) cs (t ) (6.31)
Ec Ac Ic Ec /(1 (t , 0)) ( As Ap ) Eps
cps (0) (t , 0) cs (t )
F t
1 eps2 1 (t , 0) 1
Ac I c Ec ( As Ap ) Eps
The stress change in the steel caused by creep and shrinkage of the concrete is:
F t 1 cps (0) (t , 0) cs (t )
ps,c+s t
( As Ap ) ( As Ap ) 1 eps 1 (t , 0)
2
1
Ac I c
Ec ( As Ap ) Eps (6.32)
Eps cps (0) (t , 0) cs (t )
ps,c+s t
Ac eps2
( s p ) 1
1 (t , 0) e 1
I c
Equation (6.32) can be extended by adding the influence of the relaxation of the
prestressing steel. Relaxation of steel not only implies a reduction of the steel stress but
also of the prestressing force exerted on the concrete. As a result, the time-dependent
response of the concrete is also influenced by the relaxation of steel. The interaction
between creep and shrinkage of concrete and relaxation of steel is taken into account by
reducing relaxation with a factor 0,8 (EN 1992-1-1 cl. 5.10.6 (1)).
In this expression the cross-section factor f is used (see also section 4.3):
Ac eps2
f 1
Ic
When ρp = 0 and f = 1 (only reinforcing steel and a symmetrical cross-section) are
substituted in this general expression, the earlier derived relation (6.26) remains.
stress loss from prestressing steel relaxation has a positive value (see EN 1992-1-1
eqs. (3.28)-(3.30));
the concrete stress at the level of the prestressing steel caused by quasi-permanent
loads has a positive value in case of compression.
Example
A prestressed hollow-core slab is simply supported at the two ends. It has a span l =
10 m. The selfweight is 4 kN/m and the static load is 1 kN/m. Further, the following data
hold:
Ec 36 103 N/mm 2 (C45 / 55) po 1395 N/mm 2 (table 2.4) p,max 1488 N/mm 2
Fig. 6.18 Calculation of the stress redistribution due to shrinkage and creep in a
prestressed hollow-core slab (dimensions mm (top) and m (bottom))
p Ap Ac 0, 0039
The following part of the prestressing force Pmax is transferred to the concrete (see
equation (4.11)):
1 1
Nc Pmax 6511488 932 103 N
1 e p f 1 5, 42 0, 0039 1,84
932 103
p0 1432 N/mm 2 1395 N/mm 2
651
1
Pmax 908 103 N Pmax 943 103 N
1 5, 42 0, 0039 1,84
At the beginning of the shrinkage and creep process, the concrete stress at the level of the
prestressing steel is:
2
N N e M e
cp (0) c c ps G ps
Ac Ic Ic
908 103 908 103 852 62,5 106 85
6, 4 N/mm 2
165,8 10 3
1427 10 6
1427 10 6
pi 1395
pi p0 1395 N/mm 2 0, 75
f pk 1860
Remark
The previous calculations may create the impression that time-dependent deformations
can be calculated very accurately. This is not the case since many factors can only be
estimated roughly: the relative humidity and temperature for example are not exactly
known, while in practice also the magnitude of the permanent load and the quasi-
permanent part of the variable load are only known by approximation.
When the influence of the steel on the shrinkage and creep process is neglected, a simpler
expression results, which provides a more conservative (i.e. too large) estimation of the
prestress losses.
or:
In the example presented, the prestress loss now is -190 N/mm2 instead of -170 N/mm2.
The difference in this case is 10%. The accuracy depends on the shape of the cross-
section, the modulus of elasticity of the concrete and the amount of steel.
In section 6.3.3 it was already pointed out that a simple method is obtained if the
deformations caused by the elastic shortening of the concrete are calculated using:
Ec
Ec' (6.36)
1
Ec
Ec'' (6.37)
1
with = 0,8, for stresses which develop over the course of time.
As an example a beam on three supports is considered, see fig. 6.19. The beam is
subjected to an equivalent load caused by prestressing and by a constant uniformly
distributed load qconst. Only creep is considered. It is assumed that the structure is
uncracked in the SLS.
For the calculation it is assumed that the mid support is removed. A uniformly distributed
unity load on the structure, now only supported at A and C, results in a displacement 10
(fig. 6.19). The value k10 represents the spring flexibility of the beam AC:
5 (2l ) 4
k10
384 EI c
The deflection caused by the permanent load qg and the prestressing load qp (upward
prestressing load Pm / R) then is (qg - qp)·k10. Furthermore, it is assumed that a deflection
11 occurs as a result of an upward force RB (fig. 6.19c).
(2l )3
k11
48 EI c
Over the course of time the equivalent prestressing load is reduced due to time-dependent
prestress losses. This then also holds for the reaction force RB:
B (t ) (qg (t ) qp (t )) k10 RB (t ) k11 0
where:
qp (t ) qp0 qp
RB (t ) RB0 RB
Since k10 and k11 both depend on the time-dependent value of Ec, in analogy with eq.
(6.18) the following relation must hold:
in which the deformations caused by the loads present already from t = 0 are multiplied
with a factor (1+(t,0)), and the deformations caused by loads which develop over time
are multiplied with a factor (1+(t,0)).
A change in the structural system occurs when individual structural concrete members,
which are loaded already, are coupled. As a first example the coupling of two precast
beams to one continuous, statically indeterminate, beam is regarded, see fig. 6.20. Before
coupling, the individual beams were already loaded by their permanent load qg and by the
equivalent prestressing load qp. The angular rotation of the beams at the intermediate
support B, before coupling, at time t = t0, is:
(qg qp ) l 3
B (t0 ) (6.40)
24 Ec I c
qG + q p
Fig. 6.20 Coupling of two single statically determinate precast beams to one continuous
(statically indeterminate) beam
Directly after coupling at time t0, the moment at the intermediate support MB(0) = 0.
However, as time elapses, a bending moment MB(t) develops because the following
angular rotation due to creep at the beam ends at B is restrained.
If the loads (qg - qp) and the moment MB(t) could develop independent of each other, the
angular rotation B would become, after Trost:
(qg qp ) l 3
B (t )
24 Ec' I c
(qg qp ) l 3
B1 (t ) (1 ) (6.41)
24 Ec I c
However, there would also be an angular rotation caused by the moment MB(t) that
develops over time:
M B (t ) l
B2 (t ) (1 ) (6.42)
3 Ec I c
But, since the beams are connected from t = t0, from then on no angular rotation can
occur anymore. This means that:
B1 (t ) B2 (t ) B (t0 ) 0
(qg qp ) l 2
M B (t )
8 1
This equation leads to a remarkable observation. The part -(qg - qp)·l2/8 is exactly equal
to the moment introduced by (qg - qp) if the structural system is continuous from the early
beginning, so if the system were built as a “monolithic” system immediately. Therefore:
M B (t ) M B,mono
1
The development of the support moment MB over time is shown in table 6.3. The
calculation is based on the creep function shown in fig. 6.21 (the influence of prestress
losses, as discussed in section 6.7.1, has not been considered).
(t )
For loads applied after the coupling of the beams, the monolithic structural system
applies. According to section 6.7.1, no redistribution occurs because of this load.
For the general case of redistribution due to a change in the structural system, Trost
derived the following equation:
M (t ) M 0 ( M mono M 0 )
1
or:
M (t ) M 0 1 M mono (6.43)
1 1
where:
M(t) is the moment at time t caused by a load that was applied before the coupling;
M0 is the moment at time t = t0, before coupling;
Mmono is the moment that would have occurred if the system were monolithic from
the early beginning.
Figure 6.22 shows the case of a bridge, built according to the cantilevering erection
method. In this case the bridge is built from two sides. As soon as the cantilevering parts
have reached the mid of the span, they are coupled, which leads to a change of the
structural system.
Fig. 6.22 Redistribution of moments caused by creep of the concrete in a bridge built
by the cantilevering erection method
In the middle of the span Mspan(0) = 0 (just before coupling at time 0), so it follows:
M span (t ) M span,mono
1
At the support there is already a moment Msupport,0 before the coupling is realized, so here
the result is:
M support (t ) M support,0 1 M support,mono
1 1
There are also construction methods where the structural system changes a number of
times. An example of this is a segmental bridge, which is built in one direction, see fig.
6.23.
3
2
construction part 1
M1
M2
M3
Mmono
Mmono
M
M
n
(t , ti ) (t , ti )
M (t ) M i 1 M mono
i 0 1 (t , ti ) 1 (t , ti )
where:
Due to a sudden support settlement B (fig. 6.24) a redistribution of forces and moments
occurs. The supports A and C are more heavily loaded, whereas the support reaction at B
is reduced. Due to the settlement at B, moments occur which have to be superimposed on
the moments that already exist. The development in time of those moments caused by a
sudden support settlement, will be analysed.
RB011 B (6.44)
In the course of time no further settlement occurs, in spite of the change of the reaction
force RB in time:
RB (t ) RB0 RB
RB RB0
1
RB (t ) RB0 RB RB0 1
1
The effect of a sudden support settlement is significantly reduced in time: with = 2,5
and = 0,8 one finds:
It was assumed that the structure remains uncracked. If the structure would crack, its
bending stiffness would be reduced. As a result, RB() would be reduced even more.
Now the structure shown in fig. 6.25 is regarded, which undergoes a slowly increasing
support settlement.
It is assumed that the development of the support settlement in time is similar to that of
the creep function:
(t )
B (t ) ( )
B
RB 11 (1 ) ( )
B
or:
B ( )
RB . RB,fic
11 (1 ) (1 )
where RB,fic is the fictitious support reaction in B if B, would have occurred
immediately at t = t0. Because of the condition that RB(t) = RB the following relation is
obtained:
RB (t )
RB,fic (1 )
The results of this function, with the time-dependent development of the creep according
to fig. 6.21, is given in table 6.4.
The analysis of the redistribution of forces due to creep leads to a number of important
conclusions, which are summarised here:
1) The forces and moments which arise in a structure due to a sudden deformation
(support settlement) can be calculated by multiplying the corresponding elastic
distribution of forces and moments with a factor:
kφ1 1 with = 0,8 (see section 6.7.3)
1
2) The forces and moments that occur in a structure due to a gradually developing
deformation (slow support settlement, slow temperature change, shrinkage) can be
determined by multiplying the corresponding distribution of forces and moments by a
factor:
1
kφ2 (see section 6.7.4)
1
3) In the case of a changing structural system, the finally occurring forces and moments
can be calculated as the sum of the following contributions:
The linear elastic distribution of forces in any phase, immediately after
application of the load (selfweight, prestressing), multiplied with:
kφ1 1
1
The linear elastic distribution of forces and moments, which would occur in the
structure if the monolithic structural system would have been present from the
early beginning, multiplied with:
kφ3 (see section 6.7.2)
1
Here only the loads should be considered that are already present before the final
structural system is activated. The loads applied afterwards act on the monolithic
system and do not result in redistribution of forces.
In a statically indeterminate structure which is monolithic from the beginning, the forces
and moments only change due to prestress losses (section 6.7.1.)
6.8 Literature
6.1 Trost, H., Mainz, B., Wolff, H.J.: “Calculation of prestressed concrete structures
in the serviceability limit state under consideration of the time-dependent
behaviour of concrete”, Beton- und Stahlbetonbau, 1971, Nos. 9 & 10 (in
German)
6.2 Menn, C.: ”Long term processes in concrete structures”, Lecture Notes, ETH-
Zürich (in German)
6.3 König, G., Gerhardt, H.C.: ”Redistribution of internal forces and moments due to
creep and shrinkage of concrete in reinforced and prestressed concrete structures”,
Mitteilungen aus dem Institut für Massivbau der TU Darmstadt, Nr. 34, Ernst &
Sohn (in German).
After the structure is designed, it must be shown that SLS requirements concerning
maximum initial concrete compressive stress (EN 1992-1-1 cl. 5.10.2.2) and concrete
tensile stress or crack width (EN 1992-1-1 cl. 7.3) are met and that structural resistance
meets the ULS requirements. With regard to ULS, one of the resistances to check is the
bending moment resistance of the structure.
The minimum amount of reinforcement required to resist the cracking moment follows
from the cracking moment of a cross-section. Figure 7.1 shows a statically determinate
beam prestressed by a straight tendon having an eccentricity ep relative to the centroidal
axis (c.a.) and a working prestressing force Pm,.
c.a
where fcr is the concrete flexural tensile strength (r = rupture) (EN 1992-1-1 eq. (3.23)).
The bending moment (ME,r) at which the cross-section cracks, follows from:
P
M g+q Pm, ep f cr m, Wcb (7.2)
Ac
To prevent brittle failure to occur at the onset of first cracking, the bending moment
resistance MRd must be greater than the cracking moment from eq. (7.2). However, there
is a limit: MRd has not to be greater than 1,25 MEd (EN 1992-1-1 cl. 9.2.1.1 (1) & NL
National Annex) where MEd is the design bending moment.
Reinforced concrete
M E,r M cr 16 bt h 2 f ctm,fl
M Rd 0,9d f yd As,min
M Rd M E,r
f ctm,fl
As,min 0, 26 bt d
f yk (7.3)
As,min 0, 0013 bt d
Prestressed concrete
Note that eq. (7.3) does not hold for non-rectangular cross-sections, nor for prestressed
concrete. When following the same approach for prestressed concrete, the cracking
moment is calculated using the following expression:
P Pm,
M cr f cr m, Wcb f ctm,fl Wcb
Ac Ac
Note that the minimum amount of reinforcement has not to be based on a bending
moment greater than 1,25 MEd (EN 1992-1-1 cl. 9.2.1.1 (1) & NL National Annex).
Also note that in the case of prestressed concrete the cracking moment includes the axial
compressive stress introduced by the prestressing. Prestressing might be regarded as a
fictitious increase of the concrete flexural tensile strength.
The Dutch National Annex to EN 1992-1-1 takes into account the influence of an axial
compressive or tensile force on the cracking moment. Moreover, in this annex equation
(9.1N) from EN 1992-1-1 is deleted and presented in a more general way, making it
applicable to any cross-section. The result is a set of expressions that is similar to the
expression for Mcr that holds for prestressed concrete.
EN 1992-2 ‘Concrete bridges – Design and detailing rules’ follows a different approach
to prevent brittle failure to occur. Any of the following methods may be used (EN 1992-2
cl. 6.1 (109)):
Calculate (1) the applied bending moment from the frequent combination of actions
and (2) the reduced area of prestress that results in a tensile stress fctm at the extreme
tension fibre in the cross-section considered. Now calculate the ultimate flexural
capacity of the cross-section, adding reinforcing steel to resist the moment due to
the frequent load combination. (Note: Reduced partial material safety factors are
used).
Calculate the cracking bending moment of a cross-section, using an appropriate
concrete tensile strength at the extreme tension fibre, ignoring any effect of
prestressing. Provide a minimum reinforcing steel area such that the cracking
moment calculated can be resisted by the reinforcing steel only. Note that this
calculation method results in the use of reinforcing steel, whereas in some
pretensioned members, e.g. hollow core slabs, only prestressing steel is applied.
The code then allows the use of an alternative approach in which the pretensioned
tendons can be effective in As,min.
Agree on an appropriate inspection regime (Note: Not allowed in The Netherlands).
The rotational capacity of a member is its capacity to deform. Rotation is the product of
length and curvature. Therefore, the rotational capacity depends directly on the
compression zone height in a cross-section which is a measure for the curvature. When is
rotational capacity required or beneficial for the behaviour of a structure? Rotational
capacity is required in case plastic hinges are assumed in a calculation based on the
theory of plasticity or when bending moment redistribution is applied. In case of
statically indeterminate structures, sufficient rotational capacity enables the structure to
resist, up to a certain limit, differential settlements and imposed deformations, without
them having substantial impact on structural resistance. Figure 7.2 contains an example.
If the beam shown in fig. 7.2 is reinforced according to the theory of elasticity, yielding
of the steel at the span and support cross-sections occurs at the same moment. A
differential settlement (fig. 7.2b), however, introduces an imposed support bending
moment which makes that the yield moment is reached at the support first. The rotational
capacity of the plastic hinge at the support (fig. 7.2c) makes that the yield moment can
now also be developed in the span, before failure occurs at the support (caused by
reaching the ultimate strain of the concrete, e.g. εcu3 in EN 1992-1-1 fig. 3.4). Sufficient
rotational capacity makes that, in general, the designer does not have to take into account
the effect of imposed deformations on structural resistance (EN 1992-1-1 cl. 2.3.1.2 (2) &
cl. 2.3.1.3 (2)).
Fig. 7.2 Example of a differential settlement of a support and the development of the
full plastic moment of resistance
According to EN 1992-1-1 cl. 5.5 the bending moments from linear elastic analysis in a
structure primarily loaded in bending can be redistributed to a limited extent. The extent
to which redistribution is allowed, depends on the concrete strength class, the height of
the concrete compression zone in ULS after redistribution on the basis of the
reinforcement applied, and the type of reinforcing steel used. According to EN 1992-1-1
cl. 5.5 (eqs. (5.10a) & (5.10b)) the following holds:
0, 0014 xu
0, 44 1, 25 0, 6 for fck < 50 N/mm2
cu2 d
0, 0014 xu
0,54 1, 25 0, 6 for fck > 50 N/mm2
cu2 d
where:
δ is the ratio of the redistributed moment and the moment from an elastic analysis;
xu is the concrete compression zone height (= depth of neutral axis) after
redistribution;
d is the effective depth of the cross-section.
It is not allowed to apply redistribution in case the rotational capacity cannot be defined
with confidence, as is the case for columns (EN 1992-1-1 cl. 5.5 (5) & 5.5 (6)).
The expressions use εcu2 as ultimate strain of concrete. It has the same value as εcu3 from
EN 1992-1-1 fig. 3.4 (see EN 1992-1-1 table 3.1).
EN 1992-1-1 makes it possible to carry out a plastic analysis, provided that additional
requirements are met. One of these is a more strict reduction of the concrete compression
zone height compared to the one that holds in case of limited redistribution. Since
prestressed concrete structures are usually not checked using plastic analysis, this item is
not further addressed.
f x
u for fck < 50 N/mm2
500 f d
7f x
u for fck > 50 N/mm2
cu 10 7 f d
6
where
f pk
pm, Ap f yd As
f s
Ap As
Note that f is in N/mm2 and must not be mistaken for the drape of a tendon.
The reasoning behind the introduction of f is that the expressions from EN 1992-1-1 seem
to be applicable in case of reinforced concrete structures only. Therefore, the equations
were extended to account for possible prestressing steel too. Additionally, the equations
obtained for reinforced concrete only (Ap = 0) were tuned to give the same results as NEN
6720 (VBC 1995). The modifications can be relevant in design: in case of reinforced
concrete and no redistribution (δ = 1) EN 1992-1-1 results in xu / d < 0,448 for < C50/60;
the Dutch National Annex results in xu / d < 0,535 for < C50/60 and steel B500 (fyd =
435 N/mm2).
In case the relative compression zone height exceeds its limit value, the designer has
several options. For instance, it can be decided to increase the height of the cross-section.
The increase of internal lever arms will reduce the amount of steel required. It’s also an
option to apply reinforcement in the compression zone. The compression zone height is
reduced since the part of the compressive force now carried by reinforcing steel has not
to be carried by concrete. It is noted that this option is not preferred in The Netherlands.
The Dutch National Annex to EN 1992-1-1 provides another option, namely performing
the check taking into account a fictitious amount of (prestressing and reinforcing) steel.
This is the amount of steel required to resist the design bending moment in the cross-
section. The reasoning behind this is as follows: The amount of prestressing steel
required in a concrete structure often follows from SLS requirements (e.g. no tensile
stresses are allowed or crack widths must be limited), not from ULS requirements. A
ULS cross-sectional analysis then often demonstrates that the bending moment resistance
is (much) greater than the design bending moment (MRd >> MEd). From a ULS point of
view, the structure is provided with (much) more steel than required. This enhances
structural safety, but has a negative impact on the rotational capacity. To compensate for
the latter, the designer then is allowed to fictitiously remove steel from the cross-section
when calculating the compression zone height. This of course provided that the modified
(fictitious) cross-section can resist the design bending moment (MRd,fictitious > MEd).
The shape of the stress-strain relationship of concrete is described by the surface factor α
and the centre of gravity factor β:
Just as for reinforced concrete, the following conditions must be taken into account when
determining the bending moment resistance of a prestressed cross-section:
The concrete reaches the ultimate compressive strain εcu2 if a parabolic stress-strain
relationship is used; εcu3 for a bi-linear stress-strain relationship (fig. 7.3).
At least a part1 of the reinforcing steel in the cross-section yields. This implies that εs
> fyd / Es. The ultimate strain εud of the reinforcing steel is only of importance when
using the upward sloped part of the curve (fig. 7.4). When the horizontal branch is
used, the steel strain has not to be checked (EN 1992-1-1 cl. 3.2.7).
The strain in the prestressing steel is higher than the “computational value” of the
0,1% proof-stress fpd = fp0,1k / s. The strain limit εud (EN 1992-1-1 cl. 3.3.6) of the
prestressing steel only has to be accounted for when the sloped branch of the curve is
used; not if the horizontal part is used (fig. 7.5).
Note that the ultimate strain of concrete is usually indicated by per mille ( 0 00 ), whereas
the ultimate strain of both reinforcing and prestressing steel is often expressed in percent
(%)!
Figure 7.6 shows a beam loaded by its selfweight (qg), a variable load (qq) and
prestressing (qp). In the midspan cross-section, the following moment is present from the
design loads:
1
In case more than one layer of reinforcement is applied, it might be possible that only the layer having the
most eccentric position with respect to the neutral axis yields.
1
M Ed
8
g qg q qq p qp l 2 p Pm, epo
where g = 1,2 for permanent loads, g = 1,5 for variable loading and p = 1,0 for the
prestressing load.
The calculation of the bending moment resistance MRd is discussed step by step. It is
assumed that (MEd, NEd) > (Mcr, Ncr) where Mcr and Ncr are the moment and force for
which cracking occurs, respectively. Structural safety requires that MRd > MEd.
c.a.
Fig. 7.6 Beam subjected to permanent load, variable load and prestressing load
1) First, the height of the compressive zone is estimated. For that purpose the
compressive forces are determined. When the reinforcing steel yields (using the
horizontal branch from fig. 7.4 in the calculation), the total force in this steel is As fyd.
When the prestressing steel is in the plastic state (using the horizontal branch from
fig. 7.5), its total force is Ap fpd. Since equilibrium of horizontal forces is required, it
should hold (see fig. 7.7):
N c Pm, N p N s Ap p, Ap f pd p, As f yd
From this first approximation the height xu of the compressive zone is obtained. In
case of a rectangular concrete compressive zone cross-section it reads:
As f yd Ap f pd p, Pm,
xu (7.5)
b f cd
(in case of a
compression zone
having a constant width)
M Rd
c.a.
N p Ap f pd p,
2) Check whether the height xu of the compressive zone meets the requirement with
regard to the maximum height of the compressive zone according to the Dutch
National Annex to EN 1992-1-1 cl. 5.5:
f x
u for fck < 50 N/mm2
500 f d
where:
f pk
pm, Ap f yd As
f s
Ap As
and where δ is the ratio of the redistributed moment to the elastic bending moment
(according to the linear theory of elasticity).
3) With the obtained height xu of the compressive zone, the concrete and prestressing
steel strains can be determined (fig. 7.8).
c.a.
Fig. 7.8 Determination of the strains in concrete and prestressing steel for a certain
height of the compressive zone xu
cu cu s
xu ds
ds
s cu 1 (7.6)
xu
The increase of the strain in the prestressing steel Δεp can be determined in a similar
way:
cu cu p
xu dp
dp
p cu 1 (7.7)
xu
4) With these strains εs and Δεp, the stresses in the reinforcing and prestressing steel σsu
and σpu can be calculated. From these strains the force in the reinforcing steel and the
increase of the force in the prestressing steel are obtained:
N N s N p As su Ap pu p, (7.8)
It is noted that the stress in the reinforcing steel often will be σsu = fyd, but of course
this has to be checked.
Fig. 7.9 Graphical representation of the iteration process for the determination of the
height of the concrete compressive zone xu
Figure 7.9 shows a graphical representation of the iteration process. For the first
assumed compressive zone height xu, the magnitude of the concrete compressive
force Nc from eq. (7.4) and the steel force N + Pm, (with N from eq. (7.8)) are
presented in horizontal direction, using the solid vertical line from the strain diagram
as axis. If N + Pm, is larger than Nc, as is the case in the first iteration step shown in
fig. 7.9, the compression zone height is too small and the calculation has to be
repeated using a larger xu. So, with graphical support as shown in the figure, the
correct value for xu can be found for which axial equilibrium is satisfied.
6) With the correct height xu of the concrete compressive zone, the magnitude of the
bending moment resistance MRd can be determined. This follows from (also see fig.
7.10):
where:
y is the position of the resulting compressive force;
zc is the distance from the controidal axis to the concrete fiber that has the
highest compressive strain.
MRd relative to S
centroidal axis
For a rectangular cross-section y = β xu; for concrete strength class < C50/60 β = 7/18 =
0,39.
Example
As an example, the bending moment resistance of the cross-section shown in fig. 7.11 is
determined. The data relevant for this calculation are:
MRd
10,7 N/mm2
As fyd
Figure 7.11a shows the midspan cross-section of the beam. The equilibrium between the
internal and external loads is visualised in fig. 7.11d. At the left hand side of the element
the following loads act: the centrical prestressing force Pm, and the flexural bending
moment MRd.
In the case considered, MEd is:
M Ed g M g q M q p M p (7.11)
where:
Mg is the bending moment from permanent loads (selfweight and static loads);
Mq is the bending moment from variable loads;
Mp is the bending moment from prestressing (only caused by an upward tendon
pressure since in this case the tendon is anchored in the centroidal axis at both beam
ends).
Step 1
The stress in the prestressing steel at the moment of failure is estimated to be σpu = fpd =
1522 N/mm2. Further, it is assumed that the reinforcing steel yields. The following
equilibrium equation holds in case of a rectangular cross-section:
N c Pm, As f yd Ap f pd p,
Note that the working prestressing force is incorporated in the load (eq. (7.10)). The part
σp, of the tensile strength of the prestressing steel is used already. That is why the
component Ap (fpd – σp,) is in the right-hand side part of eq. (7.10).
The value found for xu is larger than the height of the flange (200 mm). This implies that
the force in the compressive zone is overestimated, since part of it has a width equal to
the web width (250 mm), not the flange width (800 mm). The actual compressive zone
height must be greater to resist the reinforcing and prestressing steel force assumed. If the
800 mm flange width would be available over the full compressive zone height, the
concrete compressive force is 3444 kN. The compressive zone height must be increased
to xu = 258 mm to carry the same force (3444 kN) when taking into account that part of
the compression zone width is reduced to web width. Increasing the compressive zone
height reduces the internal lever arms. In this case, however, its impact is small since the
part of the web loaded in compression has a minor contribution to the compressive force
(76 kN; about 2%).
Step 2
Now the height of the compressive zone is checked according to the Dutch National
Annex to EN 1992-1-1 cl. 5.5:
f x
u for fck < 50 N/mm2
500 f d
where:
f pk
pm, Ap f yd As
f s
Ap As
and where δ is the ratio of the redistributed moment to the elastic bending moment
(according to the linear theory of elasticity).
No redistribution is applied, so δ = 1,0.
f pk
pm, Ap f yd As
1860
1058 2000 435 905
1,1
f s 571 N/mm 2
Ap As 2000 905
xu f 500 500
1 0, 47 xu 0, 47 820 385 mm
d 500 f 500 f 500 571
Step 3
For the assumed height of 258 mm for the compressive zone, the strain in the reinforcing
steel is (also see fig. 7.11f):
ds 820
s cu 1 3,5 103 1 7, 6 103
xu 258
The centers of gravity of the reinforcing steel and the prestressing steel coincide. The
strain in the prestressing steel directly follows from:
1058
p p pw 0, 0076 13, 0 103
195 10 3
Step 4
The stress in the prestressing steel assumed in the analysis is fpd = 1522 N/mm2. This
stress is reached at a strain fpd / Ep = 1522 / 195 · 103 = 7,81 · 10-3 (the strain at the kink
in the stress-strain relationship from fig. 7.5). The actual strain is higher (13,0 · 10-3) and,
as a result, the actual stress will be higher than the 1522 N/mm2 assumed.
The strain at which εud is reached depends on the type of prestressing steel. For strands,
εuk = 35 · 10-3 can be assumed.
The prestressing steel stress at the actual 13,0 · 10-3 strain then is:
Summation yields:
N Ns N p 1396 kN
Step 5
The strain and the concrete compressive stress at the bottom of the compression flange
are:
258 200
c,flange 3,5 103 0,8 10
3
258
c,flange 0,8 10 3
c,flange f cd 23,3 10, 7 N/mm 2
c3 1, 75 10 3
The stress distribution over the concrete compressive zone is presented in fig. 7.11g.
The concrete compressive force is:
N c 800 12 258 23,3 200 12 258 800 12 23,3 10, 7 12 250 258 200 10, 7
2404, 6 965, 6 77, 6 103 N 3448 kN
Now it will be checked if the requirement from horizontal force equilibrium is satisfied.
In this case it should hold:
N Pm, N c
Substitution of these values using Pm, = 2000 · 1058 = 2116 · 103 N results in:
It can be concluded that the height of the compressive zone is somewhat underestimated.
Therefore, the calculation will be repeated using a higher value of xu.
ds 820
s cu 1 3,5 103 1 7,3 103
x
u 265
1058
p p pw 7,3 103 12, 7 103
195 10 3
The strain and the concrete compressive stress at the bottom of the compression flange
are:
265 200
c,flange 3,5 103 0,86 10
3
265
c,flange 11, 4 N/mm 2
The bending moment resistance of the cross-section is calculated relative to the centroidal
axis. It is first assumed that the cross-section has a width equal to the flange width over
its full height. Then, the actual web width is taken into account by extracting the part of
the cross-section that’s actually not present:
x h
M Rd bflange xu f cd zct xu 0,5 xu hflange bflange bweb c,flange zct hflange u flange
3
As f yd ds zct Ap p d p zct
0, 75 800 265 23,3 (378 0,39 265)
265 200
0,5 (265 200) (800 250) 11, 4 378 200
3
905 435 (820 378) 2000 (1557 1058) (820 378) 1603 106 Nmm
(7.12)
1
M Ed
8
g qg q qq p qp l 2 (7.13)
The selfweight of 0,36 · 25 = 9,0 kN/m and a static load of 1,5 kN/m, result in qg =
10,5 kN/m.
For a working prestressing force Pm, = 2116 kN and a drape f = 0,62 – 0,18 = 0,44 m
(fig. 7.11b), it is found that:
qp 8 Pm, f / l 2 8 2116 0, 44 252 11,9 kN/m
With g = 1,2; q = 1,5 and p = 1,0, eq. (7.12) and (7.13) demonstrate that a live load qq =
13,2 kN/m can be resisted by the structure with sufficient safety against failure:
1
1603 1, 2 10,5 1,5 qq 1, 0 11,9 252 (7.13)
8
0,5 Pm, e0
Pm, e0 Mp2
c.a.
In this case the moment introduced by the equivalent prestressing load (by tendon
curvature) qp is:
8 P f
M p1 18 qp l 2 18 m,2 l 2 Pm, f
l
However, there is also a second component due to prestressing. This is caused by the
eccentric introduction of the prestressing force at both ends of the structure, which also
causes a moment at the intermediate support. This moment is:
M p2 12 Pm, eo
Furthermore, in the centre of gravity of the cross-section an axial compressive force Pm,
acts.
The bending moments Mp1 and Mp2 and the axial force Pm, do not change when the
structure is loaded to failure. For all loads originating from prestressing, the load factor is
p = 1,0, so:
M p p ( M p1 M p2 ) Pm, ( f 12 eo )
and
N p p Pm, Pm,
Because the internal (resistance) moment and the external (design) moment at the
intermediate support must be in equilibrium, the next condition should be satisfied:
where:
zp distance from the concrete compressive force Nc to the centre of gravity of the
prestressing steel
zs distance from the concrete compressive force Nc to the centre of gravity of the
reinforcing steel
ec distance from the concrete compressive force Nc to the line of action of the axial
prestressing force Pm, introduced at the ends of the structure.
The left hand side of eq. (7.14) represents the moment resistance MRd. The right hand side
is the design moment MEd. If, at the control, it turns out that MRd < MEd, then reinforcing
steel As should be added until MRd ≥ MEd. It is of course also an option to increase the
prestressing.
8. Shear
8.1 Introduction
In this chapter the shear serviceability and ultimate limit state analysis of prestressed
elements is discussed. As an introduction a recapitulation is given on the shear analysis of
non-prestressed structural elements containing longitudinal reinforcement only (section
8.2.1) and also provided with shear reinforcement (section 8.2.2). In section 8.3 the shear-
stress analysis of prestressed beams is discussed. SLS is discussed in section 8.3 and ULS in
section 8.4. In the last section (8.5) a special item, namely vertical prestressing of the web, is
discussed.
To better understand the influence of prestressing on the shear behaviour, a summary is given
of the shear behaviour of an “ordinary” reinforced concrete structure. Figure 8.1 shows a
beam subjected to two symmetrically positioned point loads. Between the two external loads
the beam is loaded in pure bending.
Fig. 8.1 Crack formation due to shear and bending in a reinforced beam without shear
reinforcement
At increasing load, one of the flexural cracks (caused by bending) will develop into a shear
crack that propagates over a large distance. As a consequence, internal equilibrium is lost and
failure occurs. This case, where a critical shear crack develops out of a flexural crack, is
denoted as flexural-shear failure. The area of concrete in compression above the advancing
inclined crack is that much reduced, that it can no longer support the compressive force
caused by flexure. Since the behaviour of this type of failure is quite complicated, the
analysis is based on empirically derived expressions. A good prediction of the 5% lower limit
of the shear capacity of beams without shear reinforcement is obtained by the following
expression (8.1):
1
d 3 1
VRk,c 0,15 3 100 l f ck 3 bw d (8.1)
a
where:
Expression (8.1) has a coefficient of variation of 12,5% (for the mean value VRm the coefficient
0,15 has to be replaced by 0,18 and the characteristic value of the cylinder compressive strength
has to be replaced by its mean value fcm). A design value is obtained by dividing VRm from eq.
(8.1) by c = 1,5.
EN 1992-1-1 cl. 6.2.2 (1) presents an expression that is similar to eq. (8.1):
1
VRd,c CRd,c k 100 l f ck 3 k1 cp bw d
(8.2)
VRd,c vmin k1 cp bw d
where:
3 1
vmin 0, 035k 2 f ck2
When an element is subjected to an axial compressive force, the shear capacity of that
element increases since the cracks are partially closed and crack growth is reduced. An axial
tensile force has the opposite effect. The Dutch Annex of EN 1992-1-1 uses the
recommended values: CRd,c = 0,12 and k1 = 0,15.
If the design shear force is higher than the shear resistance of a member that contains no
shear reinforcement, shear reinforcement has to be provided to increase the resistance. As a
result of this provision, the behaviour significantly changes: the formation of an inclined
crack does not anymore result in failure and a new load transfer mechanism that can be
described by a truss model develops, see fig. 8.2. The load transfer of a truss model is
discussed first.
It is assumed that the truss has to carry a design shear force VEd. At first, the required amount
of shear reinforcement is calculated. It is assumed that closed stirrups (two legs active per
strirrup; total cross-sectional area Asw for each stirrup) are spaced at a distance s. In the truss
model a number of stirrups are represented by one tensile tie.
A tensile tie is the equivalent for the stirrups in the beam over a distance z (cot + cot), see
fig. 8.2, where z is the distance from the tensile reinforcement to the resulting concrete
compressive force. The tensile bar has an equivalent steel cross-sectional area of Aequi:
Asw
Aequi z (cot cot ) (8.3)
s
where θ is the angle of the compression struts and α is the inclination of the tensile ties.
On the basis of equilibrium it follows that, in case of a full truss model, the tensile force NT
in a tensile tie in ULS is:
VRd,s
NT (8.4)
sin
where VRd,s is the shear force that is carried by the shear reinforcement, see fig. 8.2.
In case the stress in the shear reinforcement reaches the design yield stress fywd, it follows
from eqs. (8.3) and (8.4) that
Asw V
z (cot cot ) f ywd Rd,s
s sin
internal
lever arm
Fig. 8.2 Truss model with compression struts at an angle and inclined tensile ties at an
angle
The shear force that can be resisted follows from the applied amount of shear reinforcement
per unit length of the beam (EN 1992-1-1 eq. (6.13)):
Asw
VRd,s z f ywd (cot cot ) sin (8.5)
s
In case of vertical stirrups ( = 90º) and compression struts at an angle = 45º, eq. (8.5)
becomes:
Asw
VRd,s z f ywd (8.6)
s
Tests showed that designing shear reinforcement using eq. (8.5) is conservative. This is
caused by the following aspects:
the pure truss is a simple representation of the actual behaviour: the connections between
the truss bars are not perfect hinges;
redundancy from dowel action of the longitudinal reinforcement;
From series of tests it turned out that the extra resistance (often denoted as “concrete”
capacity) is equal to the bearing resistance of the member not provided with shear
reinforcement (eq. (8.2)). It was, therefore, often assumed that the shear resistance is the sum
of the concrete and steel resistances from eqs. (8.2) and (8.6). This approach was followed in
the Dutch code NEN 6720.
Additional research demonstrated that this approach can be modified. Results from shear
resistance experiments demonstrated that, at failure, the first initial bending/shear-cracks are
crossed by a second crack pattern that has a different angle to the beam axis than the first
pattern. The difference between both crack pattern angles depends on the amount of shear
reinforcement applied: the smaller the amount of shear reinforcement, the larger the
difference. The failure mode seems to adapt itself to the shear reinforcement applied: the
smaller the amount of shear reinforcement, the smaller the angle to the beam axis of the
second crack pattern and the larger the number of shear reinforcement bars crossed and
activated by an inclined shear crack. This is denoted as rotation of the compression
struts/diagonals. When applying this reduced crack angle in eq. (8.5), the shear resistance is
accurately predicted. Figure 8.3 shows an example in which the resistance of the truss not
taking into account an additional “concrete” part is shown. This line is denoted as Mörsch’s
line.
Fig. 8.3 Shear force test: stress in shear reinforcement as a function of the shear force
from a test and for truss models with different compression strut angle
The angle can be reduced to 18,4º. EN 1992-1-1 assumes a lower limit value = 21,8º
(cot 21,8º = 2,5).
In case vertical shear reinforcement is applied ( = 90º), eq. (8.5) reduces to EN 1992-1-1 eq.
(6.8):
Asw
VRd,s z f ywd cot (8.7)
s
The width of a strut in the plane of a beam is indicated in fig. 8.4. The width is equal to the
distance bD between the struts, which is:
Fig. 8.4 Calculation of the dimensions of the compression struts from the geometry of the
truss model
From force equilibrium it follows that the force in the compression strut is:
VEd
ND (8.9)
sin
where VEd is the design value of the shear force. When using eqs. (8.8) and (8.9), the concrete
compressive stress σcD in the strut follows from:
The resistance of the concrete compressive strut can be calculated if the allowed concrete
compressive stress is known. This stress is called σcd. The shear force resistance of the
concrete struts is denoted as VRd,max. From eq. (8.10) it follows that (EN 1992-1-1 eq. (6.14)):
cot cot
VRd,max bw z (cot cot ) sin 2 cd bw z cd
1 cot 2
The compressive struts cannot be loaded up to the uni-axial concrete compressive strength
because the stirrups that cross the concrete diagonal struts in the web are loaded in tension.
The tensile forces perpendicular to the struts’ direction are transferred by bond. Their impact
is that strut strength is reduced, see fig. 8.5a. Figure 8.5b shows a failure envelope of
concrete, loaded in two directions. In the struts a combination of a tensile and a compressive
stress occurs. The dashed line from fig. 8.5b shows that the stress combination results in a
reduction of the ultimate compressive stress relative to the cylinder compressive strength fck.
Furthermore, from experiments it turns out that the maximum stress does not proportionally
increase with the concrete strength class. The following expression is derived (EN 1992-1-1
cl. 6.2.3 (3)):
cd cw 1 f cd (8.11)
where:
cw is a coefficient taking into account the state of stress in the compression strut;
1 is a strength reduction factor for concrete cracked in shear.
If the design stress of the shear reinforcement is below 80% of the characteristic yield stress
fyk (EN 1992-1-1 eq. (6.10aN) & (6.10bN)):
f 'c
fc
compression
Fig. 8.5 Reduction of the concrete compressive strength due to transverse tension from
bond forces introduced by the shear reinforcement
where σcp is the mean compressive stress (compression is positive) in the concrete. Figure 8.6
shows cw as a function of σcp.
Fig. 8.6 Coefficient for the concrete strut compressive strength as a function of the axial
compressive stress
The designer is free to choose an angle between 21,8º and 45,0º. A small angle has a
positive effect on the shear resistance of the shear reinforcement (eq. (8.7)): The smaller the
angle, the more reinforcement is crossed by the crack. As a result, the shear resistance
increases. However, the smaller the angle, the higher the compressive stress in the inclined
concrete struts (eq. (8.10) and, as a result, the lower the shear resistance. Designers often
assume = 21,8º and increase this value only if the concrete strut compressive stress is
governing.
As a result of the shear resistance mechanism with its inclined crack, the force in the
longitudinal reinforcement is larger than follows from a cross-sectional equilibrium analysis.
This is illustrated in fig. 8.7. Note that z is the distance from the tensile reinforcement to the
concrete compression zone, which is often assumed to be almost equal to the distance to the
resulting concrete compressive force (see fig. 8.2, 8.4 and EN 1992-1-1 fig. 6.5).
al
I
VEdI M Ed
Aswfywd
z
Fig. 8.7 Equilibrium analysis to explain the bending moment shift rule
It is assumed that the “concrete” part VRd,c of the shear force capacity can be totally attributed
to the shear capacity of the uncracked compression zone. In case all stirrups yield, the
resulting force VRd,s is at a distance ½ z cot from section I-I. The resultant R of the forces
VRd,s and VRd,c is through point S. The distance al from point S to section I-I follows from:
V V VRd,c
al z cot 12 Rd,s Rd,c 12 z cot 1 (8.13)
VRd VRd VRd
From the bending moment equilibrium around point S, the force NsI in the steel is found.
Here one finds:
I
M Ed VI a
N sI Ed l (8.14)
z z
So, the steel force NsI in section I-I is obtained from the bending moment in section II-II, by
dividing it by the internal lever arm.
In practice this is solved by shifting the moment line over a distance (bending moment line
shift rule):
V
al 1 2 z cot 1 Rd,c (8.15)
VRd
Fig. 8.8 Shift (B) of original bending moment curve (A) according to EN 1992-1-1 fig.
9.2
Shifting the bending moment line over al = ½ z (cot - cot ) for members with shear
reinforcement. Alternative: shift over al = d.
Instead of shifting: add an additional tensile force to the longitudinal reinforcement force
(EN 1992-1-1 cl. 6.2.3 (7)).
M y Pm Pm ep y
x (8.16)
Ic Ac Ic
c.a.
x x2
I xy
2
(8.18)
2 4
x x2
II xy
2
(8.19)
2 4
xy
tan 2 I (8.20)
1
2x
compression tension
Fig. 8.10 Determination of the principal stresses in the serviceability limit state.
If the principal tensile stress appears to be too high, the stress can be reduced by increasing
the thickness of the web or by prestressing in vertical direction as shown in fig. 8.11.
vertical prestressing
Fig. 8.11 Influence on the principal stresses of a vertical prestress in the web
With Mohr’s circle, it can be shown that the principal tensile stress can be compensated for
completely by applying prestressing in vertical direction (fig. 8.11, σI < 0). For this case, the
principal stresses are:
y
2
x y x
I xy
2
(8.21)
2 4
y
2
x y x
II xy
2
(8.22)
2 4
If cracks in SLS are allowed, their width should be well controlled. The stress in the stirrups
at SLS can be determined using the truss analogy (eq. (8.7) with s instead of fywd). With the
tensile member model, which will be discussed in Chapter 9, the crack width can be
calculated.
Example
Figure 8.12 shows the beam from section 4.7. Stresses in the cross-section which is 2 m from
a beam end are analysed. Beam length l = 21,5 m.
At midspan, the centroidal axis of the four anchors is 125 mm from the bottom fibre. THe
tendon profile is a parabola. The anchors are at neutral axis level.
The cross-section is now modelled using three rectangles, see fig. 8.12.
qq = 12 kN/m
qg = 2,5 kN/m
qselfweight
zct
zcb
Fig. 8.12 Example of the calculation of the principal stresses in SLS (dimensions mm,
stresses N/mm2).
Loads in SLS:
Concrete selfweight plus additional static load: qG = 8,2 + 2,5 = 10,7 kN/m
75% of the variable load (“representative load”) qQ,rep = ψ qQ = 0,75·12,0 = 9,0 kN/m
Working prestressing force P∞ = 2187 kN qP = -2187/136,6 = -16,0 kN/m
2187 10 3
c 3
6,67 N/mm 2
328 10
The following figures contain the bending stress and shear stress over beam height. The
principal stress that follow from these two stresses are also included, as well as the angle θ
between the longitudinal axis of the beam and the principal stress direction. The shear stress
is relatively small compared to the shear stress; the minimum principal stress (compression)
is almost equal to the bending stress and the angle θ is almost zero.
800
700
600
z [mm]
500
400
300
200
100
0
-5.6771
-7.6 -7.4 -7.2 -7 -6.8 -6.6 -6.4 -6.2 -6 -5.8 -5.6
2
x [N/mm ]
In section 8.2.1 it is demonstrated that the shear capacity of an element without shear
reinforcement depends on several influencing factors (eqs. (8.1) and (8.2)). The shear
capacity also depends on the development of (bending) tensile cracks since they may connect
and develop into shear cracks.
Pm Pm ep
cb
Ac Wcb
W
M 0 cb Wcb Pm cb ep Pm 1
6 h ep
Ac
M0
Vn (8.24)
a
Fig. 8.13 Influence of the prestressing force on the shear capacity of a structural element
without shear reinforcement
When loading is continued from the situation shown in fig. 8.13b, the same situation is
obtained as when starting to load a non-prestressed beam. This implies that the shear capacity
of the prestressed beam is increased by Vn = M0 / a relative to a reinforced beam.
M Md
a
V Vd
where M and V are the moment and shear force at the position of the load, respectively. Thus,
equation (8.24) can be rewritten as:
M0
Vn Vd (8.25)
Md
For general use, EN 1992-1-1 uses eq. (8.24) as a starting point. Assuming that the beam has
a rectangular cross-section with d = 0,85 h and ep = 0,35 h (as often used in experiments), the
following results are obtained.
Pm Pm 0,35h 0,5h Pm
cb 3
3,1
bh 12 bh
1
bh
M cb 16 bh 2 0,52 h Pm
This bending moment must be introduced by the support reaction Vn at a distance a from the
cross-section considered:
M 0,52 h Pm Vn a
h Pm d Pm P
Vn 0,52 0,52 0, 61 m
a 0,85 a a/d
Many tests have been carried out for a ratio a / d between 2,5 and 4,0. The following results
apply:
a / d = 2,5 : Vn = 0,24 Pm
a / d = 4,0 : Vn = 0,15 Pm
Research indicated that 0,15 Pm is a conservative lower bound value. This value is used in
EN 1992-1-1; see the coefficient k1 in eq. (8.2). This increase of the shear resistance can only
be applied in case the member has not to be provided with shear reinforcement.
A second aspect is that the prestressing results in a part of the pre-compressed tensile zone
remaining uncracked, even in the ultimate limit state (fig. 8.14). In these parts, flexural
tensile stresses are compensated for by the prestressing (area a in fig. 8.14). Flexural cracks
will occur at bottom side parts of the beam where the flexural tensile stress exceeds the
tensile strength of the concrete (area b in fig. 8.14).
due to prestressing
tensile strength concrete
due to bending
Fig. 8.14 Crack pattern in a prestressed beam, and the stresses in the bottom part
In area a, where the bottom part remains uncracked, even in the ultimate limit state, shear
cracks may still develop. These cracks don’t start at the bottom fibre but origin in the web at
the position where the principal tensile stress in the concrete reaches the concrete tensile
strength. From eq. (8.18) it follows that this occurs in case:
x x2
I xy
2
f ctd
2 4
where:
VEd Sc
xy
bw I c
It can be found that an inclined crack occurs in case (EN 1992-1-1 eq. (6.4)):
bw I
VRd,c f ctd2 x f ctd (8.26)
S
Note that eq. (8.26) contains no αl since it is assumed that the full prestressing force is
transferred at the beam end. The influence of the transmission length is discussed later.
When no shear reinforcement is applied in area a, the development of an inclined crack under
the load from eq. (8.26), will generally lead to failure. This type of failure is called tensile
splitting shear failure. EN 1992-1-1 cl. 6.2.2 (2) states that it is only allowed to use eq. (8.26)
at locations where the tensile bending stress in the outer fibre of the cross-section is less than
fctk,0,05 / γc.
40
185
265
40
40 185 45 185 45 185 45 185 45 185 40
1195
The slab is prestressed with 14 Y1860S7 strands of 9,3 mm (3/8''; 52 mm2). To avoid
splitting of the concrete, the two outer strands are stressed to half their capacity, which
implies that, with regard to prestressing, effectively 13 strands are present. The total effective
steel area is Ap = 13 · 52 = 676 mm2. The following calculations are based on 13 effective
strands, not taking into account that 14 strands are present in the cross-section.
Material properties:
concrete strength class C45/55
f ctd 1,8 N/mm 2 ,
I c 1,55 109 mm 4 ,
Ac 1, 78 105 mm 2 ,
ecb 132 mm ; ect 133 mm
Wcb 11, 7 106 mm3 (concrete bottom fibre),
Sc 7, 66 106 mm3 ,
bw 260 mm (governing cross-section at the level of maximum hollow core width),
d d p 225 mm ,
p, 1100 N/mm 2 .
The working prestressing force Pm, 676 1100 744 103 N . The bending moment
introduced by prestressing is constant over the slab length:
concrete at the location considered. The steel stress is 0,404 · 1100 = 445 N/mm2. The shear
capacity of the concrete (tensile splitting shear failure) is (eq. (8.26)):
Note that the concrete is assumed to crack at the design value of its tensile strength, fctd =
fctk,0,05/γc (EN 1992-1-1 cl. 6.2.2 (2)).
Figure 8.16a shows this 132 kN starting point of the tensile splitting shear resistance curve.
The curve shows that this failure mechanism first shows a gradual increase of the resistance.
This comes from the gradual introduction of the prestressing force in the slab. The maximum
resistance is reached at lpt2 = 500 mm from the end of the slab. At that position, the full
prestressing force is in the slab which implies that σcp = 4,18 N/mm2 (compression) and
VRd,c,max = 173 kN.
It is now assumed that the prestressing steel at midspan reaches a stress pu = fpd =
1522 N/mm2 in ULS (at the kink in the stress-strain diagram, which is conservative). With Ap
= 676 mm2, p, = 1100 N/mm2 and an assumed internal lever arm z 200 mm, the bending
moment resistance is (see fig. 8.17):
MRd = 676 · (1522 – 1100) · 200 + 676 · 1100 · (132 – 40) = (57,1 + 68,4) · 106
MRd = 135,5 · 106 Nmm
If the slab is simply supported with a span of 10,0 m and subjected to a uniformly distributed
load, the design value of the load that can be resisted in bending is:
1 2 8 (135,5 68, 4) 106
M Ed M Rd qd l M p, 135,5 10 qd
6
16,3 N/mm 16,3 kN/m
8 10, 02
At the cross-section where the prestressing force is fully transferred to the concrete, the
concrete stress in the bottom fibre is:
Pm M p P Pm ep
cb m 4, 2 5,9 10,1 N/mm 2
Ac Wcb Ac Wcb
W
M o cb Wcb Pm cb ep Pm 1
6 h ep
Ac
VRd,c,max
governing width
Since the concrete is assumed to crack when the design value of its tensile strength fctd =
fctk,0,05/γc is reached (EN 1992-1-1 cl. 6.2.2 (2)), the design value of the flexural cracking
bending moment is:
P
M cr Wcb m f ctd 11, 7 106 (4, 2 1,80) 70 106 Nmm 70 kNm
Ac
When the bending moment resistance MRd = 135,5 kNm (at a design load qd = 16,3 kN/m and
Mp, = -68,4 kNm) is reached, the length of area a, assuming a parabolic distribution of the
bending moment, follows from:
1
2 x (10 x) 16,3 68, 4 70 kNm
81,5 81,52 4 8,15 138, 4
x 5, 00 2,83 2,17 m or 7,83 m
2 8,15
This result indicates that the slab is uncracked up to a distance of 2,17 m from the supports.
The remaining part (10,0 – 2,17 – 2,17 = 5,66 m in length) is cracked in ULS and, as a result,
will demonstrate flexural-shear failure, whereas the two uncracked parts exhibit tensile
splitting shear failure.
The shear capacity VRd,c of the cracked area b is now calculated, using EN 1992-1-1 cl.
6.2.2 (1). From equation (8.2) (EN 1992-1-1 eq. (6.2a) & (6.2b)) it follows that:
where
200 200
k 1 2, 0 k 1 1,94
d 265 40
A 676
l sl 0, 012
bw d 260 (265 40)
VRd,c vmin k1 cp bw d
VRd,c 0, 035 1,943/ 2 451/ 2 0,15 4,18 260 225 73,8 103 N
The shear resistance VRd,c = 88,2 kN from flexural-shear failure is shown in fig. 8.17,
together with the shear resistance that follows from tensile splitting shear failure. The figure
also shows the linear shear force curve that holds for uniform loading.
Shear failure occurs if the design shear force curve crosses one of the shear resistance curves.
The design shear force is zero at midspan. An increase of the uniform load on the slab makes
that the straight VEd-line rotates around the point where VEd = 0 at midspan. The governing
load and shear force are found when the rotating line first crosses the resistance lines VRd,c. It
then follows that tensile splitting shear failure occurs first. Failure is at 0,2 m from the end
face of the slab. It then holds:
1 132
qd l 0, 2 132 kN qd 27,5 kN/m
2 5, 0 0, 2
27,5
qd 22,9 kN/m 2
1, 2
At a selfweight of:
Ac 25 0,178 25
qg 3, 7 kN/m 2
1, 2 1, 2
it can be written:
qq 12,3 kN/m 2
Mcr = 70 kNm
88,2
Bending moment failure occurs at a design value of the load qd = 16,3 kN/m (MEd =
135,5 kNm). Shear failure, namely tensile splitting shear failure, occurs at qd = 27,5 kN/m.
These results demonstrate that the slab fails in bending, not in shear.
Fig. 8.18 shows a typical tensile splitting shear fracture in a laboratory test. Note that the line
load is close to the line support to have a relatively high shear force and a small bending
moment.
Fig. 8.18 Tensile splitting shear fracture in a beam that is prestressed by pre-tensioned steel
In the case of structures prestressed with straight tendons, the effect of prestressing is as
follows:
In area b, cracked in bending, the calculation of the required shear reinforcement is almost
the same as for reinforced concrete. The only difference is that in eq. (8.2) the axial
compressive stress σcp results in an increase of shear resistance.
In area a, uncracked in bending, the uncracked lower flange has a positive effect on shear
resistance. This can be explained by cutting a section along a shear crack (fig. 8.19).
a b
In area a both the uncracked compression zone and the uncracked tensile zone contribute to
shear resistance. Because the cracks do not proceed to the outer fibres of the beam, they
hardly open. Therefore, the crack width is small and the shear reinforcement is only lightly
stressed. Figure 8.20 shows the results of a shear test on a prestressed beam, carried out by
Leonhardt, Koch and Rostásy [8.2].
Fig. 8.20 Stresses in the shear reinforcement of a prestressed beam in the areas uncracked
in bending (a) and cracked in bending (b) [8.2]
Because of the constant shear force in the areas between the load and the supports, the beam
was provided with a constant shear reinforcement. Measuring the strains in the stirrups
demonstrated that in area a the stresses were only about 1/3 of those in area b.
If the shear force is higher than the force that causes tensile splitting shear failure, which
implies that shear reinforcement has to be applied, the contribution of the "concrete" (which
is the component between Mörsch's line and the stirrup stress measured, see fig. 8.3) is
somewhat higher than in case of shear bending failure. EN 1992-1-1 could take this positive
effect into account by allowing for a larger rotation of the compressive diagonal concrete
struts. This is not accounted for in the code: it is prescribed that the calculation of the
required amount of shear reinforcement should be carried out following the same procedure
as for area b.
With regard to the resistance of the concrete compressive struts other considerations apply.
An axial compressive force has basically two effects. On one hand it postpones cracking and
crack propagation in the web of a beam. On the other hand, the axial load generates inclined
compressive stresses in the struts as well, so that a part of their capacity is consumed. An
evaluation of tests showed, that a small normal force improves the bearing capacity of the
struts, but a large force reduces it, see fig. 8.6 [8.9].
Figure 8.21a again shows the total load exerted on a beam prestressed using draped tendons.
The beam is analysed at ULS. Apart from the load components qg (permanent load;
selfweight and static loads) with load factor g = 1,2 and qq (variable load) with a load factor
q = 1,5, a prestressing load acts on the structure, consisting of:
loads
“passive” steel
Ap pu p,
Since the prestressing load does not change when the structure is loaded to failure, the load
factor p = 1,0.
The stress in the prestressing steel is p,∞ which implies that the part pu - p,∞ is "not used"
yet. This part can be regarded as being "reinforcing steel" (passive prestressing steel) present
in the beam and will from now on be denoted as the ”rest-system”. This rest-system can be
provided with additional reinforcing steel to have the required resistance in bending and
shear. The rest-system is dealt with as if it was a normal reinforced concrete beam. It should
be checked whether it has sufficient resistance.
In the beam shown in fig. 8.22, in section I, the passive part of the prestressing steel and the
active force Pm,∞ introduced at the ends, provide a bending resistance:
If this were not sufficient, reinforcing steel could be added, so that the bending moment
resistance is increased to:
where zpI and zsI are the internal lever arms of the prestressing steel and the reinforcing steel
respectively, and where it is assumed that the reinforcement reaches the design yield stress
fyd. If this might not be the case, the more general variable fs needs to be used.
al
MII MI
MRd II
Asfyd
MRd,I
Asfyd
In cross-section II, which is in the area subjected to shear, the shift rule should be applied. To
be able to design the reinforcement in cross-section II, at first the bending moment curve is
shifted over a distance al, according to eq. (8.13).
The bending moment resistance around the concrete compressive force in section II is:
Also here the bending moment resistance can be increased to the required level by providing
reinforcing steel at the bottom of the beam, so that:
This shows that it is always favourable to have tensile reinforcement at the bottom of the
beam. If the beam is prestressed with more than one prestressing tendon, it is advised to have
at least one tendon along the bottom side of the beam, up to the beam end. It is advised to, if
possible, add reinforcing steel since this is not only favourable with regard to bending
moment resistance, but also for shear resistance and crack width control.
Fig. 8.23 Determination of shear resistance ( is the inclination of the shear crack)
Subsequently, shear resistance is regarded. Also here, like in the case of reinforced concrete,
an inclination of the struts between 21,8º and 45º can be chosen. The shear force to be carried
by the section considered, see fig. 8.23, is:
VEd qd 12 l x
where qd g qg q qq qp
In case the shear resistance is assumed to be the design value of the shear force, vertical
equilibrium results in the following requirement:
where:
Asw
VRd,s z f ywd cot cot sin (EN 1992-1-1 eq. (6.13))
s
z is the internal lever arm between the force in the compression zone and the centre of
gravity of the longitudinal steel.
d is the effective depth of the cross-section calculated from the centre of gravity of the
longitudinal steel
α is the angle of the shear reinforcement relative to the horizontal axis.
Also here it is demonstrated that providing a tensile tie at the bottom of the beam is
favourable, since both z and d increase, so the shear resistance increases, even if the same
amount of shear reinforcement is applied.
For statically indeterminate structures, the approach is the same. Figure 8.24 shows the loads
for a statically indeterminate beam. Also here the design load is composed of a number of
contributions:
qd g qg q qq qp
In the area close to the intermediate support, the equivalent prestressing load has a downward
direction, so that the load is:
qd g qg q qq qp
In the shear force diagram this effect is directly recognised (a kink close to the intermediate
support).
bending moment
shear force
axial force
Fig. 8.24 Design for shear with the equivalent prestressing load method
Special attention has to be given to the control of the strength of the concrete compression
struts. The ducts for the prestressing tendons create a discontinuity that should be regarded
when checking the stress in the struts (fig. 8.25).
For non-grouted ducts (as well as for grouted plastic ducts and unbonded tendons), the net
width of the compression struts, available for resisting the inclined compressive forces, is
(EN 1992-1-1 eq. (6.17)):
b bw 1, 2 Ø (8.28)
where Ø is the diameter of the prestressing ducts. The reduction is 1,2 Ø, which is more
than Ø, since it can happen that not the compressive strength of the concrete, but the
splitting tensile strength of the concrete (because of transverse tension) is governing (fig.
8.26) (EN 1992-1-1 cl. 6.2.3 (6)).
duct
b bw 12 Ø (8.29)
In case the widths of metal grouted ducts Ø < bw / 8, it can be assumed that bw,nom = bw.
It is noted that the influence of the ducts also has to be accounted for when checking tensile
splitting shear failure.
Figure 8.27 shows the crushing of the concrete compression struts due to strength loss in the
web caused by a duct. The figure shows the left span of a two span statically indeterminate
beam. The tendon profile consists of straight parts and curved parts (at midspan A and
intermediate support B). The beam is loaded by point loads, one at each span. The crack
pattern at failure is shown. Concrete crushing in the beam close to the intermediate support is
indicated. For an extended numerical analysis of the strength of the concrete struts, reference
is made to [8.8].
tendon profile
stirrups
Fig. 8.27 Shear test on a beam with inclined prestressing tendon [8.6]:
a. Side view (dimensions mm).
b. Cross-sections (dimensions mm).
c. Failure of concrete struts.
In EN 1992-1-1 the control of the strength of the compression struts is in the check of VRd,max,
see eq. (8.12) (EN 1992-1-1 eq. (6.9)). The impact of the ducts is accounted for by
introducing bw,nom according to eq. (8.28) or (8.29).
As a result of the equivalent prestressing load, a substantial part of the load can be
“balanced”. Figue 8.28 shows the inclined tendons of a box girder bridge close to an
intermediate support.
Fig. 8.28 Inclined prestressing tendons near an intermediate support of a box girder bridge
The last two advantages mentioned are a result of the much higher yield stress (or 0,1%
strain limit) of the prestressing steel than the yield stress of reinforcing steel, so that the
cross-sectional area of the shear reinforcement can be much smaller. Applying vertical
prestressing is only economic in the case of high webs because of the high costs of the
anchorages.
An example of the application of vertical prestressing is the Rhine Bridge near to Bensdorf
(Germany, 1965), with a span of 208 m (fig. 8.29). The depth of the concrete cross-section
near to the supports is 10,45 m. In the direction of midspan the depth of the cross-section is
reduced such that the nominal shear stress is approximately constant over a large distance.
Therefore, also the width of the web could remain constant at 300 mm over almost the entire
bridge. Only close to the supports the web width had to be increased to 370 mm. The webs
are prestressed with bars Ø 32 mm, steel type Y1030H, inclined at an angle of 45º [8.5].
longitudinal section
hinge
Fig. 8.29 Example of a structure with prestressed shear reinforcement: the Rhine bridge
near to Bensdorf in Germany (dimensions m) [8.5]
With regard to the shear resistance of girders with vertical prestressing, the following two
questions can be raised:
1) From eq. (8.23) it turns out that, due to vertical prestressing, the inclination of the shear
cracks increases, so that less shear reinforcement is intersected by a crack. Should this be
considered in design?
2) The difference between the design strength (ULS) and the working stress (SLS) of the
prestressing can be substantial (order of magnitude 500-600 N/mm2). Fully utilising this
reserve capacity requires a considerable strain in the prestressing steel. A large strain
implies a large deformation of the truss. Does the corresponding rotation of the struts
introduce such high secondary compressive stresses in the concrete, that strut resistance
is reduced?
To answer these questions, Kupfer and Ruhnau [8.5] tested a beam that was at one side
prestressed with inclined and at the other side with vertical prestressing tendons, see fig.
8.30. In both parts of the beam additional shear reinforcement was applied. The
reinforcement in the webs was designed such that the prestressed shear reinforcement carries
70% of the shear force; the passive reinforcement 30%. Both parts of the beam are designed
to have the same shear force resistance.
Fig. 8.30 Test on a beam with two types of vertical prestressing (dimensions mm) [8.5]
In fig. 8.31 the stresses measured in the prestressing steel and the reinforcing steel are
presented.
prestressing bar n° 2 4 6 8 10
1 3 5 7 9 11
stirrup n° 1 3 5 7 9 11
2 4 6 8 10 12
[N/mm2] 1500
load step 10
1400 V = 822 kN
load step 9
1300
V = 740 kN
1200 load step 8
V = 658 kN
1100
1000
load step 1
(prestress)
900
prestressing bar n° 1 3 5 6 8 10
2 4 7 9
[N/mm2] 500
load step 10
400 V = 822 kN
0
1 3 5 7 9 11
stirrup n° 2 4 6 8 10 12
Fig. 8.31 Stresses in the prestressing and reinforcing steel measured at different load levels
[8.5]
The stress in the inclined prestressing tendons is higher than in the vertical tendons. This can
be explained by the truss analogy.
It was concluded that the tendon stress reserve of 580 N/mm2 between SLS and ULS can be
fully utilised. On the basis of this investigation, it was proposed to put a limit of 600 N/mm2
to the prestressing steel stress reserve that can be activated. The design failure strength of the
prestressing steel then is:
8.6 Literature
8.3 Walraven, J.C., Mercx, W.P., “The bearing capacity of prestressed hollow core slabs”
Heron, Vol. 28, 1983, Nr. 3.
8.4 Walraven, J.C., “Reinforced concrete” Lecture Notes TU Delft, nov. 1991 (in Dutch).
8.5 Kupfer, H., Ruhnau, J., “Vorgespannte Schubbewerung”, Deutscher Ausschuss für
Stahlbeton, Heft 359, Berlin 1985.
8.6 Trinh, J., “Effort tranchant en beton armé et beton précontraint: influence de la
précontrainte”, Annales de l’Institut Technique du Batiments et des Travaux Publics”,
No. 360, April 1978, pp. 140-144.
8.7 Anderson, A.R., Anderson, R.G. "An assurance criterion for flexural bond in
pretensioned hollow core units", ACI-Journal, Vol. 73, Aug. 1976, pp. 457-464.
8.9 Keller, K., N.V. Tue & M. Zink, “Influence of prestressing forces on the shear
capacity- part 2: Beams with shear reinforcement", Lacer no. 7, 2002.
Fig. 9.1 Crack in a reinforced concrete beam with crack width w > 0,4 mm (not a
calculation error but damage during transport) [9.1]
In general, crack width limitation does not play a dominant role in the design of a
concrete structure. Only in the case of a low reinforcement ratio, or when due to
aggressive environmental conditions strict limits apply (e.g. w < 0,1-0,2 mm), or when
particular liquid tightness requirements must be met, the crack width criterion can be
governing.
For prestressed concrete structures other arguments apply: prestressing steel is more
prone to corrosion than reinforcing steel. The first prestressed concrete structures were
therefore designed such that in serviceability limit state conditions no cracks would
Moreover, the loads to which the structure is exposed during its life time must be exactly
known if one aims at preventing all possible cracking. However, it is often impossible to
accurately predict all loading situations that might occur during the service life,
especially if temperature effects or differential settlements are concerned.
Figure 9.2 shows the results of bending tests on two prestressed concrete beams, both
prestressed with one bar Ø 26 mm. The first one is, in the lower flange, additionally
reinforced with 2 reinforcing bars Ø 6 mm, whereas the second one has 2 reinforcing bars
Ø 16 mm. Figure. 9.2 shows the crack patterns for those two beams, tested at the
Technical University of Aachen in Germany. The first beam demonstrated one large
crack (width w = 0,85 mm), whereas the second beam showed a large number of cracks
having small widths (w = 0,3 mm).
Fig. 9.2 Crack pattern in a prestressed beam having different ratios of reinforcing to
prestressing steel (dimensions mm) [9.2]
The awareness that effective crack width control can be obtained by using a combination
of reinforcing and prestressing steel is an important basic consideration for the
application of partially prestressed concrete. When using this mixed type of
reinforcement, an optimum solution is possible for any practical case. This offers the
designer various advantages, see references [9.3] and [9.4].
The requirements with regard to the allowable crack width wmax are based on the durability
of the structure and aesthetics. Requirements from EN 1992-1-1 table 7.1N are given in
table 9.1.
Table 9.1 Recommended values of wmax [mm] (EN 1992-1-1 table 7.1N)
It is noted that the Dutch National Annex to EN 1992-1-1 prescribes other crack widths
limits and load combinations. The following table presents some data.
According to the Dutch National Annex to EN 1992-1-1, the crack width limit values
from the table can be multiplied by a factor
capplied
kx 2, 0
cnom
if the applied concrete cover on the reinforcement is larger than the nominal concrete
cover that follows from durability requirements.
Crack formation has been studied since decades. Test series have been carried out on,
among others, axially loaded reinforced concrete tensile bars, see fig. 9.3. By varying the
concrete strength class, the reinforcement ratio, the bar diameter and/or the number of
bars, a good impression was obtained of the basics of the cracking mechanisms.
In modern methods for crack width control in structures, the structural behaviour is
mostly modelled by defining an axially loaded, centrically reinforced concrete tensile bar
as shown in fig. 9.3. Therefore, this basic case is dealt with in detail first. At first the case
of a concrete tensile bar containing reinforcing steel only is discussed. Subsequently the
case of combined reinforcing and prestressing steel is dealt with.
1 2 3 4 5 6 7 8 9 10 11 12
ρs =s 0,0113
= 1.13% ρs =
s0,0201
= 2.01% ρs =0,0314
s = 3.14% ρs =0,0452
s = 4.52%
Fig. 9.3 Reinforced concrete tensile members subjected to axial tension (Stevin
laboratory, TU Delft, 1976)
Fig. 9.4 Simplified bond stress-slip relationship for short term static loading
For ε = εs = εc:
N tot Ec Ac 1 e c (9.1)
where:
Es
e = is the ratio between the Young's moduli of steel and concrete;
Ecm
As
= is the reinforcement ratio.
Ac
When the concrete strain (c) increases, at a certain moment the tensile stress in the
concrete will reach the tensile strength of the concrete. Since the tensile strength of the
concrete over the length of the member is subjected to scatter, the first crack will appear at
a location where the tensile strength is the lowest, see figure 9.5a.
At the location where the concrete is cracked, the concrete tensile stress ct = 0. In a crack
only the steel carries the tensile force. As a result of the bond stresses between the steel and
the concrete, acting at both sides of the crack, the concrete is active again in carrying the
tensile force (fig. 9.5b). At a certain distance lt from the crack (the transfer length), the
concrete carries its original part of the tensile force N. Outside the transfer lengths, the
strains of concrete and steel are again equal, so that the undisturbed situation (as before
cracking) is present.
The distance required to again introduce a part of the cracking force into the concrete
depends on, among others, the bond strength bm. For the calculation of this distance, the
basics as shown in fig. 9.5 are used. From the assumption that the bond stress is constant
(fig. 9.5e), it follows that the course of the steel stress and the concrete stress along the
transfer length lt are linear, see figs. 9.5c and d.
When further increasing the strain, the force increases again. However, the force cannot
be larger than Ncr,2, because then a new crack appears. The stage in which again and
again new cracks are formed at increasing imposed deformation is denoted as the “crack
formation stage”. In the crack formation stage the stress in the reinforcing steel s in a
crack reaches a maximum just before, at another location, a new crack arises. At that very
moment the stress is s = sr , see fig. 9.5c.
σse = e fctm
σc = fctm
N
N cr,1
Nr,1
Ncr,2
Nr,2
t 2.t 2. t 2. t t Ncr Nr
N N
N0
disturbed area
If, as a simplification, the calculations are based on the mean concrete tensile strength
fctm, the steel stress directly after the formation of a new crack is:
N cr f ctm
s sr 1 e (9.2)
As
In the undisturbed areas, see fig. 9.6, the steel stress is directly proportional to the
concrete stress:
se = e f ctm (9.3)
At a further increase of the imposed strain, the force increases again. However, the force
cannot exceed Ncr,2, since then a new crack appears.
The concrete tensile stress ct in a crack is zero, whereas at the end of the transfer length
the concrete stress is fctm (fig. 9.5d). This implies that the force transmitted over the
transfer length is:
N Ac f ctm (9.4)
This force is transmitted by bond stresses along the perimeters of the bars over a distance
lt. It then holds:
N = bm lt m Ø (9.5)
where m is the number of reinforcing bars and Ø is their diameter. Combining eqs. (9.4)
and (9.5) (and using ρ = As / Ac and As = m 1/4 π Ø2) gives an expression for the transfer
length lt:
1 f ctm Ø
lt = (9.6)
4 bm ρ
The maximum crack width is equal to the difference between the elongation of the steel
and the elongation of the concrete over the length 2lt, so:
where sm and cm are the mean steel strain and concrete strain, respectively, along the
transfer length lt. The course of the stresses at both sides of a crack is shown in fig. 9.7.
The strains can be calculated from the stresses.
sr
se
steel stress
t t
fctm
concrete stress
Fig. 9.7 Course of steel and concrete stresses at both sides of a crack
f ctm
εcm 2 1 α f (9.10)
e ctm
Ec 2 Es
Substituting sm and cm (from eqs. (9.9) and (9.10)) and eq. (9.6) in eq. (9.7) yields:
1 f ctm Ø 1
wmax = σ sr (9.11)
4 bm ρ Es
When the strain is further increased, more and more cracks occur. The cracking process
continues until the tensile bar consists of “disturbed regions” only. When a certain
number of cracks are formed, the disturbed regions start to overlap each other. The
smallest spacing between two cracks is found, where at the end of a disturbed region
(which is at a distance lt from an already existing crack), a new crack has occurred. The
largest spacing between two cracks is found, where a new crack has occurred at a
distance just smaller than 2lt from an already existing crack. The length of the tensile
member part in between the two cracks then is just too short for the bond stresses to make
that the concrete reaches its tensile strength again. As a result of this, the final crack
spacing varies between lt and 2lt. When, finally, the reinforced member consists of
disturbed regions only, the crack formation stage is finished. Although during further
increase of the strain the external force increases, no new cracks are formed. The stage
after the crack formation stage is denoted as the “stabilised cracking stage”. In this stage,
no new cracks occur and existing cracks widen,.
At a further increase of the strain, and as a result also an increase of the force N (fig. 9.6),
the steel stress in the crack s exceeds sr (eq. (9.2)). Because the force transmitted from
the steel to the concrete does not increase (the bond stress is constant), the concrete strain
between the cracks does not increase. As a result, the increase of the crack width follows
from the additional elongation of the steel only.
Figure 9.8 shows a crack where at both sides the maximum crack spacing is 2 ℓt. In the
crack formation stage the maximum stress in the steel was sr (eq. (9.2)). After
completion of the crack formation stage, the steel stress further increases because of the
increasing external tensile force N. The increase of the steel stress is s = s - sr. All
corresponding elongation of the steel over the distance 2 ℓt results in an increase of the
crack width:
( s sr ) 2 t
w (9.12)
Es
Fig. 9.8 Determination of the maximum crack width in the stabilised cracking stage
The total crack width in the stabilised cracking stage is obtained by adding w from eq.
(9.12) to wmax from eq. (9.11). In combination with ℓt from eq. (9.6), the following
expression is obtained:
1 f ctm Ø 1
wmax = σs 0,5 σsr (9.13)
2 bm ρ Es
Eqs. (9.11) and (9.13) are continuous since if in eq. (9.13) s is replaced by sr, this
expression reduces to eq. (9.11) (crack formation stage). Equation (9.13) is the general
expression for the calculation of the maximum crack width in both the crack formation
and the stabilised cracking stage. To calculate the crack width, it is only necessary to
determine the transition point (strain) between the crack formation stage and the
stabilised cracking stage. This is discussed in section 9.5.
The effect of shrinkage in the crack formation stage differs from that in the stabilised
cracking stage. When in the crack formation stage shrinkage occurs while,
simultaneously, the external imposed strain is kept constant, the external force tends to
increase. Since in the crack formation stage the external force cannot exceed the cracking
load Ncr, this implies that the existing crack widths will not increase. The result is that
additional cracks will develop.
In the stabilised cracking stage the shrinkage influences the crack width. In this stage no
new cracks can be formed. The shortening of the concrete then can only result in
widening of the already existing cracks. The influence of the shrinkage on the crack
width is explained on the basis of the behaviour of a reinforced concrete element having a
length 2 ℓt and restrained at both ends, see fig. 9.9(a). As a result of previous loading, the
element exhibits one crack in the centre and it is assumed that the stabilised cracking
stage is reached. On behalf of symmetry, only one half of the element is regarded, see fig.
9.9(b).
lt
Fig. 9.9 Influence of shrinkage on the crack width (stabilised cracking stage)
The concrete tends to shrink, but this is counteracted by the steel. If the concrete would
be able to shrink freely (assuming that there is no bond between the steel and concrete),
the concrete strain would be cs, see fig. 9.9(c).
Note: Shrinkage implies a shortening of the concrete, so εcs < 0. In the following
expressions it will be assumed that the absolute value of the shrinkage is used, so εcs > 0.
To restore compatibility between steel and concrete, in a first step, a compressive force N
is applied to the steel. The length of the steel is made equal to the length of the concrete,
see fig. 9.9(d). The steel stress increases with Δs = cs Es whereas the stress in the
concrete c remains constant. This situation is regarded as the initial situation. In the next
step the steel and the concrete are bonded and the same force N is applied to the steel, but
now in the opposite direction. Now concrete and steel can act together to carry the tensile
force N. However, slip occurs between steel and concrete. Compared with the initial
situation, the final result is an increase of the steel stress s = cs Es and the
corresponding increase of the crack width follows directly from eq. (9.13).
The total crack width (load + shrinkage) in the stabilised cracking stage is:
1 f ctm Ø 1
wmax = σs 0,5 σsr εcs Es (9.14)
2 bm ρ Es
Now the effect of a long term constant load and/or a varying load will be discussed.
It was stated before that assuming a bond stress τbm = 2,0 fctm between steel and concrete
gives good results for ribbed bars. Under a long term or dynamic load, the bond stress
decreases. Tests have shown that a value τbm,∞ = 1,6 fctm is realistic.
In the crack formation stage, the reduced bond stress results in an increase of the transfer
length of 25%, and, as a result, a similar increase of the crack width. This follows directly
from eq. (9.11), when instead of τbm = 2,0 fctm the value τbm = 1,6 fctm is used.
For the stabilised cracking stage, the situation is different. In most cases, the load has
been applied over a short period of time. This implies that the value bm = 2,0 fctm holds
for the transfer length, and, as a result, also for the crack spacing. The influence of the
concrete in between the cracks depends on the bond stress developed. As a result, this
influence decreases when the bond stress decreases. It can be assumed that this reduction
is about 40%. This can be taken into account by replacing the coefficient 0,5 from eq.
(9.14) by 0,3.
When taking these effects into account, the following more general expression for the
crack width is obtained:
1 f ctm Ø 1
wmax = σs σsr εcs Es (9.15)
2 bm ρ Es
where
Table 9.2 Values for τbm, and from eq. (9.15) for various conditions. The values
for between brackets are the recalibrated values as applied in EN 1992-
1-1 by means of the coefficient kt (EN 1992-1-1 eq. (7.9))
An expression for the maximum crack width wmax that agrees with the EN 1992-1-1
expression is obtained by substituting ℓt (eq. (9.6)) and sr (eq. (9.2)) in eq. (9.13). The
result is:
f ctm
σs - ( 1 + αe ρ)
ρ
wmax = 2 lt
Es
where
α = 0,5 in both the crack formation stage and the stabilised cracking stage in case of
short term loading.
f ct,eff
s kt (1 e p,eff )
p,eff
wmax sr,max
Es
where:
fct,eff is the mean value of the concrete tensile strength at the time when the crack may
first be expected to occur;
fct,eff = fctm or lower (fctm(t), if cracking is expected earlier than at a concrete age of 28
days);
kt is the influence of load duration: 0,6 (short-term loading) or 0,4 (long-term
loading);
ρp.eff is the reinforcement ratio of the effective tension area (see EN 1992-1-1 fig. 7.1),
which is a “hidden tensile member”.
When taking into account the influence of shrinkage, the equation becomes:
f ct,eff
s kt (1 e p,eff ) cs Es
p,eff
wmax sr,max
Es
EN 1992-1-1 uses the following expression for the maximum crack spacing (EN 1992-1-
1 eq. (7.11)):
Ø
sr,max k3 c k1k2 k4
p,eff
Since for k3 and k4 usually the recommended values 3,4 and 0,425 are used, sr,max reads:
Ø
sr,max 3, 4 c 0 , 425 k1k2
p,eff
where:
Note that the EN 1992-1-1 expression for maximum crack spacing is similar to eq. (9.6)
if it is assumed that the mean bond stress is directly proportional to the concrete tensile
strength. EN 1992-1-1 uses a minimum value for the crack spacing (k3 c) since this
demonstrated to give better results, especially for members having a high reinforcement
ratio.
Note:
The EN 1992-1-1 wmax expression (EN 1992-1-1 eq. (7.9)) has ρp,eff, the
reinforcement ratio of the “hidden” tensile member (see section 9.6) to calculate
tension stiffening. This implies that the steel stress σs follows from a cross-sectional
analysis, whereas σsr, the steel stress directly after cracking, follows from the
cracking force of the hidden tensile member. This approach might give unexpected
results, especially for a member loaded in bending to a level that is not much higher
than the cracking moment Mcr. Therefore, in this textbook, the steel stress directly
after cracking follows from a cross-sectional analysis, using the cracking force
(bending moment) as input.
9.5 The transition point between the crack formation stage and the
stabilised cracking stage
Figure 9.10 shows the schematised behaviour of the tensile member.
When the cracking load Ncr (r = rupture) is reached, crack formation starts (2). At
increasing deformation the load N can not exceed Ncr.
After the completion of crack formation, the force N increases. The dashed line shows the
N - Δℓ / ℓ relation of the steel reinforcement only. The line representing the behaviour of
the reinforcement with the surrounding, cracked concrete (3) is assumed to be parallel to
the line of the steel only (4). To calculate the position of line (3), it is assumed that the
mean crack spacing is 1,5 ℓt. The resulting representative zone in between two cracks is
shown in fig. 9.11.
1 = uncracked stage
2 = crack formation stage
3 = stabilised cracking stage
Ncr 4 = steel bar(s) only (unbonded)
5 = yielding of reinforcement
s.As s.As
sx s
sm
0,75. t 0,75. t
Fig. 9.11 Variation of stresses along a representative part of the member having a
length equal to the mean crack spacing and located in between two cracks
where:
1 f ctm Ø
bm 2, 0 f ctm and lt = (eq. (9.6)).
4 bm ρ
Equation (9.16) then becomes:
0, 75 f ctm
σ sx =σ s - (9.17)
ρ
The mean steel stress:
0,375 f ctm
σ sm =σ s - (9.18)
ρ
The mean steel strain:
σ s 0,375 f ctm f
sm = - s - 0, 4 ctm (9.19)
Es Es ρ Es ρ
With the aid of fig. 9.10 it is now possible to determine the strain for which the cracking
pattern can be regarded to be complete (the end of the crack formation stage).
N cr Ac f ctm 1 e (9.20)
f
N Es As s 0, 4 ctm (9.21)
Es ρ
When substituting N = Ncr in eq. (9.21) and by using eq. (9.20), the intersection point of
branches (2) and (3) is found:
f ctm 0, 6 αe ρ 0, 6 f ctm
s (9.22)
Es ρ Es ρ
If the imposed strain of the reinforced concrete tensile member is smaller than the value
resulting from eq. (9.22), the member is in the crack formation stage. For a higher value
of the strain the stabilised cracking stage is reached. In most practical applications where
imposed deformations apply, for instance in case of a temperature drop or concrete
shrinkage at fixed boundary conditions, the imposed strain is mostly smaller than the
value given by eq. (9.22). The crack pattern then is not stabilised and the structure is in
the crack formation stage (2).
If, on the contrary, the member is subjected to a tensile load N > Ncr, fig. 9.10 shows that
the member is in the stabilised cracking stage (3).
Note:
In order to simplify the calculations, fig. 9.10 is a schematised representation of the
actual behaviour. A horizontal plateau (2) will in reality not occur, since the cracking
force gradually increases: the first crack is formed at the weakest spot and each following
crack occurs at a location where the tensile strength of the concrete is slightly higher. The
most realistic description might therefore be to use the lower bound 5% characteristic
concrete tensile strength for the first crack and to end with the mean tensile strength for
the last crack. However, in the crack formation phase, the maximum crack width is found
at the highest cracking force. This implies that the designer should focus on the
maximum cracking force. By using a constant cracking force based on the mean concrete
tensile strength fctm, this is incorporated in the model.
When the expression for the crack width was derived, it was implicitly assumed that the
transfer length lt is large compared with the dimensions of the cross-section. At the
location where the tensile strength of the concrete is reached, the tensile stresses are then
almost uniformly distributed over the concrete cross-section. As a result, the cross-
section will fully crack through once the concrete tensile strength is reached.
If one dimension of the cross-section is much larger than the other, the behaviour is
different. At the position where the tensile strength of the concrete is reached, cracking
starts. The distribution of the tensile stresses that spread into the concrete is not uniform,
see fig. 9.12a. The crack now does not proceed over the full width of the tensile member.
Only if the force, introduced by the reinforcement, is more or less uniformly spread over
the width of the element, the full cross-section cracks. Close to the reinforcing steel, the
behaviour is identical to that of the reinforced tensile member discussed in the previous
sections. In the wide member only a few cracks reach the outer surface of the concrete,
fig. 9.12b.
Because several internal cracks join, continuous cracks are formed. A continuous crack
often has a disproportionally large width since, at the outside of the concrete, the
deformation is concentrated in a small number of cracks. If the reinforcement is
concentrated at the outside of the cross-section, see fig. 9.12c, the outer surface
demonstrates many cracks having small widths, whereas wide internal cracks occur.
These findings demonstrate that there is a so-called "effective concrete area" around the
reinforcement. The width of cracks that occur in this area is controlled by the
reinforcement, whereas the crack width outside this area is uncontrolled. The relations
derived in the previous sections apply to the effective concrete area only.
c
a b
fc
concentrated crack
b d
effective area
c b
Fig. 9.12 Cracking behaviour of elements with concentrated reinforcement and a high
ratio of element width to transfer length (b / lt)
In elements loaded in bending, similar phenomena occur. In the case of deep beams, the
main tensile reinforcement limits the crack widths over an area close to the
Fig. 9.13 Crack pattern in deep beam with concentrated longitudinal reinforcement at
the bottom and hardly any longitudinal web reinforcement, loaded in pure
bending (left: side view; right: cross-section)
It is found from tests that the effective area around the reinforcement (the “hidden”
reinforced concrete tensile member) in beams, walls and slabs can be defined as shown in
fig. 9.14 [9.6] (EN 1992-1-1 cl. 7.3.2 (3) & fig. 7.1). The height of the effective area is:
hc,eff 2,5 (h d )
bending: hc,eff (h x) / 3
tension : hc,eff h / 2
The reinforcement ratio used in the crack width expression must be based on the effective
concrete area. In case of reinforced concrete (EN 1992-1-1 eq. (7.10)):
As
ρp,eff (9.23)
Ac,eff
It is noted that the reinforcement ratio from expression (9.23) has a general setup by
using the subscript p. This makes it suited for combinations of reinforcing (s) and
prestressing steel (p).
Fig. 9.14 Effective concrete area [9.6] (EN 1992-1-1 fig. 7.1)
1 f ctm Ø 1
wmax = σs σsr εcs Es (9.24)
2 bm ρ Es
where
Table 9.3 Values for τbm, and from eq. (9.24) for various conditions.
To distinguish between the crack formation stage and the stabilised cracking stage, the
following principles generally apply:
0, 6 f ctm
(9.25)
Es p,eff
External loads: Generally, the stabilised cracking stage applies, provided that the
tensile force is higher than the cracking force:
This chapter deals with the question how to calculate the largest crack width that occurs
in a structure in the serviceability limit state (SLS). However, requirements might also
focus on the mean crack width. In the crack formation stage, scatter in the crack width
only comes from the scatter in the concrete tensile strength. The relation between the
maximum and the mean crack width then is (approximated):
In the stabilised cracking stage, the scatter is larger because now also the stochastic
nature of the crack spacing plays a role. The relation now is (approximated):
In case of imposed deformations, which occur, for example, due to a decrease of the
temperature or restrained shrinkage, the structure will generally be in the crack formation
stage.
For a number of cases eq. (9.24) was used for the crack formation stage, see fig. 9.15.
The figure shows diagrams for concrete strength classes C20/25 and C35/45, both for
short term and long term or dynamic loading. An important question is which concrete
tensile strength has to be used in the calculations. In practice, there are various
influencing factors that reduce the tensile strength of the concrete, e.g. temperature
gradients from solar radiation, moisture gradients, or micro cracking from temperature
gradients in the hardening phase of the concrete. In practical situations, the tensile
strength will, therefore, be somewhat smaller than in laboratory conditions. To take all
these factors in account, a general reduction factor of 0,75 for the concrete tensile
strength is introduced. Furthermore, it turns out that the actual concrete strength in a
structure can be considerably higher than the characteristic cube compressive strength of
the mixture. The 28-day characteristic strength is the basis for most structural resistance
calculations (e.g. bending, shear, torsion, etc). With regard to crack width, however, an
increase of the strength results in an increase of the cracking force and, as a result, in an
increase of crack width. To be on the safe side, it is suggested to increase the
characteristic strength with 10 N/mm2. This implies that for a concrete C20/25, a
characteristic strength of 20 + 10 = 30 N/mm2 is used in the following calculations.
When implementing the two aforementioned corrections (namely a 0,75 tensile strength
reduction factor and a 10 N/mm2 compressive strength increase), the mean concrete
tensile strength in the actual structure is (using the EN 1992-1-1 table 3.1 expression):
f ctm 0, 75 0,30 20 10
2/3
2, 2 N/mm 2
f ctm 0, 75 0,30 35 10
2/3
2,9 N/mm 2
For the various cases, the data listed in table 9.4 apply. The long term bond stress is 80%
of the short term bond stress (table 9.2). With regard to concrete creep, it is assumed that
the creep coefficient is 2,0. Creep is taken into account by reducing the Young's modulus
of the concrete with a factor 1 + 2,0 = 3,0.
Table 9.4 Basic values used to construct the design curves from fig. 9.15
concrete load Ec fctm bm αe = Es / Ec
duration [GPa] [N/mm2] [N/mm ] 2
[-]
In the crack formation stage, the maximum steel stress in a crack is (eq. (9.2)):
f ctm
s sr 1 e
In fig. 9.15a-d, the relations between wmax, Ø and ρ are presented for the four cases
considered. The allowable bar diameter Ø can be read for a given reinforcement ratio and
an allowed maximum crack width, provided that the tensile member is in the crack
formation stage.
ρ [%]
[%] 3.0 ρ [%]
[%] 3.0
Wmax =0.1 mm
Wmax =0.1 mm
0.2
2.0 2.0
0.2 0.3
0.3
0.4
1.0 0.4 1.0
a b
0.0 0.0
0 10 20 30 40 0 10 20 30 40
[mm] [mm]
Ø [mm] Ø [mm]
ρ [%] ρ [%]
[%] 3.0 [%] 3.0
Wmax =0.1 mm Wmax =0.1 mm
0.2
0.0 0.0
0 10 20 30 40 0 10 20 30 40
[mm] Ø
[mm]
[mm]
Ø [mm]
Fig. 9.15 Required reinforcement ratio for a given bar diameter and maximum crack
width for reinforced concrete tensile members in the crack formation stage:
a. C20/25, short term loading
b. C20/25, long term or dynamic loading
c. C35/45, short term loading
d. C35/45, long term or dynamic loading
Apart from the allowable crack width there is another important criterion: the yield stress
of the steel should not be exceeded at first cracking. On the basis of eq. (9.2), it can be
derived that the minimum reinforcement ratio to prevent yielding is:
1
min (9.30)
s / f ctm e
For s the design yield stress of the steel fyd = 435 N/mm2 (reinforcing steel B500) is
used. With regard to the tensile strength of the concrete, it should be taken into account
that it might be higher than the mean strength as derived from the characteristic 28-day
cylinder compressive strength (EN 1992-1-1 table 3.1). Previously, fck was increased with
10 N/mm2. To be at the safe side, the reduction with a factor 0,75 is not applied in this
case.
f ctm 0,30 20 10
2/3
2,9 N/mm 2
When using αe = 7 (short term loading), eq. (9.30) results in ρmin = 0,70%.
An overview of the minimum reinforcement ratio for a range of concrete strength classes
is given in table 9.5. Note that EN 1992-1-1 table 3.1 uses a different expression for
strength classes > C50/60. When using the parameters discussed before, this expression
is:
2/3
53 8 10
f ctm 2,12 ln 1 4, 4 N/mm 2 for strength class C53/65.
10
Example 1a
A long slab without expansion joints has a thickness h = 150 mm and is on a stiff
bedding, see fig. 9.16. Due to the assumed high friction between the slab and the subsoil,
deformation of the slab from shrinkage of the concrete and a temperature drop, cannot
occur. In the calculation it is assumed that the shrinkage cs = 0,25 · 10-3 and that the
mean temperature will drop 25 ºC below the initial temperature during construction. The
concrete strength class is C20/25 and the reinforcing steel class is B500. The crack width
allowed is 0,20 mm. Calculate the required amount of reinforcement and the bar
diameter.
Solution
When the concrete shrinkage and the temperature drop are not restrained, the total
deformation of the slab is:
Because shortening cannot occur, tensile stresses will develop. Since it is a long term
deformation, creep should be taken into account. In case of a creep coefficient of 2, the
fictitious Young's modulus of the concrete is about 30 / (1 + 2) = 10 GPa. In case of an
uncracked slab, the fully restrained deformation would cause a concrete tensile stress:
The concrete will crack because its tensile strength is only about 2,2 N/mm2 (table 9.4).
It is now assumed that the cracked structure is in the crack formation stage (which often
holds in case of an imposed deformation). This assumption has to be validated
afterwards. From fig. 9.15b (long term loading) it follows that for wmax < 0,20 mm and Ø
= 12 mm, the reinforcement ratio should be at least ρ = 1,1%. If the reinforcement is
provided in two layers (at bottom and top of the slab), the maximum bar spacing is
135 mm. The required reinforcement ratio is of course higher than the minimum
reinforcement ratio (ρmin = 0,70%, table 9.5) since minimum reinforcement is applied to
prevent the steel from yielding whereas the reinforcement ratio from fig. 9.15 is required
to control crack widths. At a reinforcement ratio ρ = 1,1%, the transition point from the
crack formation stage to the stabilised cracking stage is at a strain:
0, 6 f ctm 0, 6 2, 2
0, 60 103
Es 200 103 0, 011
This shows that the assumption that the structure is in the crack formation stage, was
correct.
Example 1b
Now, the same situation is regarded, but for a slab having a thickness of 400 mm (fig.
9.17). A reinforcement ratio of 1,1% would also in this case be sufficient to meet the
crack width requirement. However, it will be discussed whether a lower reinforcement
ratio might be applied if the reinforcement is concentrated in the outer zones (top and
bottom) of the slab since concentrating reinforcement in the outer zones has proven to be
more effective than uniformly distributing it over the full cross-section (fig. 9.12c).
Solution
The height of the effective tensile area at the bottom and the top of the slab (fig. 9.14):
It is now assumed that the slab is in the crack formation stage. This implies that the
tensile force in the slab is constant: N = Ncr. The steel stress directly after cracking and
steel stress at the SLS both follow from the cracking force off the full cross-sectional area
of the slab:
f ctm
s sr 1 e
effective tensile
member; 78 mm
effective tensile
member; 78 mm
Fig. 9.17 Thick reinforced concrete member subjected to an axial force from an
imposed deformation (dimensions mm)
After substituting the previous expressions into eq. (9.24) and assuming = 0,3, = 1
and bm = 2 fctm, the result is:
1 1 Ø 1 f ctm f
wmax = 1 + e ρ 0,3 ctm 1 + e ρ εcs Es
2 2 ρp,eff Es ρ ρ
(9.31)
1 1 Ø 1 f ctm
wmax = 0, 7 1 + e ρ εcs Es
2 2 ρp,eff Es ρ
In this expression, a clear distinction is made between ρ (based on the total height h of the
cross-section) and ρp,eff (related to the effective tensile area and used to calculate the
crack spacing in the effective tension zones or “hidden” tensile members). The ratio
between ρp,eff and ρ is:
ρp,eff 400
= 2, 6
ρ 2 78
To meet the requirement wmax = 0,20 mm, using ρp,eff = 2,6 ρ, fctm = 2,2 N/mm2, αe = 21,
Ø = 12 mm and cs = 0,2510-3, it is found from eq. (9.31) that
1 1 12 1 2, 2
wmax = 3
0, 7 1 + 21ρ 0, 25 103 200 103
2 2 ρp,eff 200 10 ρ
1,5 1 1,54
0, 2= 5 1 + 21ρ 50
2, 6 ρ 10 ρ
When applying reinforcing bars Ø 12 mm, the maximum bar spacing is 70 mm.
It is now checked whether the slab is actually in the crack formation stage as assumed.
The transition point is at a strain:
0, 6 f ctm 0, 6 2, 2
0,83 103 0,50 103 -> crack formation stage
Es 200 10 0, 0080
3
Example 2
A slab, spanning in one direction, is subjected to a variable load qq of 4 kN/m2 (fig. 9.18).
The concrete strength class is C20/25. The reinforcement (at the tension side) consists of
bars Ø12 – 175 mm, which is As = 645 mm2/m. The concrete cover c = 15 mm. The
longitudinal reinforcement ratio:
As 645
sl 0, 25 102
bd 12
1000 275 15
2
It can be assumed that the maximum load acts only for a short period of time at the
structure. The maximum crack width in that situation should be not larger than 0,4 mm.
4 kN/m22
qqq q==4kN/mm
275
6000
15
12-175
Fig. 9.18 Slab spanning in one direction with qq = 4 kN/m2 (dimensions mm)
Solution
According to table 9.4, the cracking stress of concrete C20/25 is 2,2 N/mm2. The flexural
tensile strength is (EN 1992-1-1 eq. (3.23)):
When applying this expression to the slab (h = 0,275 m), the cracking stress is
2,9 N/mm2.
The maximum bending moment in SLS is:
compression
Fig. 9.19 Internal equilibrium in a cross-section just before (left) and after cracking
(right)
x
e sl ( e sl ) 2 2 e sl
d
Using αe = 7 for short term loading and ρsl = 0,25 · 10-2 (based on d = 275 – 15 =
260 mm) the result is:
x
0,17
d
This implies that x = 43 mm and the internal lever arm after cracking is:
z = d – x / 3 = 254 – 43 / 3 = 240 mm
The maximum steel stress s in a crack follows from:
For a structural element loaded in bending, the height of the effective tension area around
the steel reinforcement is the smallest value of 2,5 (c + Ø/2) and (h - x) / 3 (see EN 1992-
1-1 cl. 7.3.2 (3) and section 9.6).
hidden
77.3
77 hidden
tie tensile member
12
Fig. 9.20 Hidden reinforced tensile member in the slab (dimensions mm)
Note:
It is noted that the EN 1992-1-1 section on cracking (7.3) does not use the flexural
tensile strength; crack formation and the behaviour of the hidden tensile member
are related to the uni-axial tensile stress.
The steel stress sr in a crack directly after cracking follows from:
M cr 36, 6 106
s 236 N/mm 2
As z 645 240
Note:
The steel stress in the tensile member in a crack in the crack formation stage
according to EN 1992-1-1:
f ctm
sr
p,eff
1 0,2,0084
e p,eff
2
1 7 0, 0084 277 N / mm 2
Because the maximum moment in SLS is larger than the cracking moment, the structural
slab is in the stabilised cracking stage. The load considered has a short term character. It
then follows from table 9.3 that = 0,5, = 0 and bm = 2 fctm.
1 f ctm Ø 1 1 1 12 1
wmax = σs 0,5 σsr 316 0,5 236 0,35 mm
2 bm ρp,eff Es 2 2 0, 0084 200 103
One of the main aims of the full prestressing of a structure is to keep it uncracked under
any possible SLS load combination. So, basically, a crack width control for fully
Fortunately, practice has learned that limited cracking in prestressed concrete does not
endanger its durability. On the basis of this important conclusion, a new way of
prestressing was developed, the so-called partial prestressing, which allows controlled
crack formation. In Chapter 12 the technology of partial prestressing will be dealt with in
more detail. In this section it will, in advance of the more general aspects of partially
prestressed concrete, be investigated how the crack width can be calculated and
controlled in the case of cross-sections provided with a combination of reinforcing and
prestressing steel.
The bond properties of prestressing steel are not as good as those of reinforcing steel. In
practice, therefore, reinforcing steel is often applied additionally to prestressing steel.
Figure 9.21 shows the development of the stresses in reinforcing steel, prestressing steel
and concrete. Since the bond properties of prestressing steel are less good than those of
reinforcing steel bars having the same diameter, prestressing steel has a larger transfer
length than reinforcing steel. However, the same crack width must be found for both
types of steel. This implies that the increase of the stress in the crack is smaller for
prestressing steel than for reinforcing steel.
The cross-sectional areas of the reinforcing and prestressing steel are denoted as As and
Ap, their diameters as Øs and Øp, and their bond stresses as bms and bmp. Also a
distinction is made between their transfer lengths, using lst for reinforcing steel and lpt for
prestressing steel.
The increase of the tensile force in the steel compared when the stress in the concrete is
zero, is:
N = As s + Ap p (9.32)
Since the crack width is the same for reinforcing steel and prestressing steel, ws = wp.
Assuming Es = Ep and neglecting the concrete deformation over the transfer length, it is
found that:
As a result:
s p
2lst 2lpt (9.33)
2 2
or:
s Øs
lst (9.35a)
4 bms
p Øp
lpt (9.35b)
4 bmp
where:
p is the increase of the steel stress in the prestressing steel in a crack;
s is the steel stress in the reinforcing steel in a crack.
s s Øs p p Øp
Es 4 bms Ep 4 bmp
bmp Øs
p s s 1 (9.36)
bms Øp
where 1 is the so-called bond factor, also taking into account different diameters (EN
1992-1-1 cl. 6.8.2 (2) & eq. (6.64); cl. 7.3.2 (3) & eq. (7.5)):
bmp Øs Ø
1 s (9.37)
bms Øp Øp
N = As.s + Ap.p
Ap
4.As
sr p
pr
s p f
e ctm
ne.fctm
steel stress
st
pt
concrete stress
Fig. 9.21 Development of steel and concrete stresses in the area adjacent to a crack
As can be seen from eq. (9.37), the bond factor contains both the ratio between the bond
strengths (ξ) and the ratio between the bar and tendon diameters. In EN 1992-1-1 table
6.2 the bond strength ratio is given (see table 9.6).
Because of the high sensitivity of prestressing steel to corrosion, the allowable crack
widths are generally 0,1 mm smaller than for reinforcing steel that is under the same
environmental conditions (exposure classes).
By virtue of the use of the bond factor, combinations of prestressing and reinforcing steel
can easily be transformed to the case of only reinforcing steel. The same principles as
previously discussed in detail, can be used. The only difference is that the reinforcement
ratio is defined in another way.
As 1 Ap
p,eff s 1 p (9.38)
Ac,eff
For the calculation of the crack width, the stress in the reinforcing steel is used as a basis,
so eq. (9.24) can still be used:
1 f ctm Ø 1
wmax = σs σsr εcs Es (9.39)
2 bm ρp,eff Es
The only difference is that the (combined) reinforcement ratio ρp,eff follows from eq.
(9.38). Also with regard to the calculation of s and sr, the influence of the prestressing
steel can easily be taken into account.
The use of eq. (9.39) will be illustrated with an example in section 9.10. For detailed
information about crack width control for combinations of reinforcing steel and
prestressing steel, reference is made to [9.7].
The beam shown in fig. 9.22 is designed for a maximum variable load qq = 15 kN/m. The
selfweight of the structure is 8,5 kN/m and there is an additional permanent load of
0,9 kN/m. The layout of the post-tensioned prestressing tendons along the beam is
parabolic. At the beam ends, the centre of gravity of the prestressing steel coincides with
the centre of gravity of the cross-section of the beam. The concrete strength class is
C35/45.
Ac = 0,34 m2
Ic = 0,031 m4
d = 0,850 m
zb = 0,635 m (distance from centre of gravity to concrete bottom fiber)
zt = 0,365 m (distance from centre of gravity to concrete top fiber)
Wb = 0,049 m3
qq = 15 kN/m
qvar=15 kN/m
800
1200 kN
483
200
20 m
c=25
1000
10 Ac = 0.34 m2
635
12 57 zu = 0.635 m
= 0.031 m4
300
. Wu = 0.049 m3
3x75
97
qdl = 9 kN/m
.
225 d = 0.850 m
za = 0.365 m
103
dimensionsin mm
225
The bending reinforcement is concentrated at the bottom of the beam, see fig. 9.22. It
contains mixed reinforcement, and has a height of approximately 300 mm (fig. 9.22, left).
This part of the beam contains 9 reinforcing bars Ø 12 mm (As = 1018 mm2) and two
Dywidag prestressing tendons 4/15,7 Y1860C, both having a cross-sectional area of
4 · 150 = 600 mm2 (Ap = 1200 mm2) and an external duct diameter of 57 mm.
For the calculation of the crack width, a tensile strength fctm = 3,2 N/mm2 (EN 1992-1-1
table 3.1) is used. The flexural tensile strength fctm,fl follows from the axial tensile
strength fctm using the relation (EN 1992-1-1 eq. (3.23)):
The prestressing force is 1200 kN, which implies a working stress pm∞ = 1000 N/mm2.
The allowable short term crack width is 0,2 mm. Verify whether this criterion is met.
Solution
For a prestressing force of 1200 kN and a parabolic layout of the prestressing tendons, the
equivalent prestressing load-balancing force is:
The mean concrete compressive stress cpm in the cross-section caused by the axial
prestressing force is:
This demonstrates that, under the maximum moment at the SLS, the beam is cracked in
bending. To calculate the crack width, the maximum steel stress s must be known.
Therefore, first the internal equilibrium is regarded. To obtain equilibrium the following
requirements apply:
H 0 and M 0
The corresponding forces and strains for H = 0 are shown in fig. 9.23.
[N/mm2]
2723,3
1.75 3.50
co
' [‰]
ε [‰]
c
In this case (fig. 9.22), the centers of gravity of reinforcing steel and prestressing steel
almost coincide, so that approximately ds = dp = d = 850 mm. The bond factor is 0,5
(table 9.6).
When assuming that the concrete is still in the elastic stage (c 1,75‰, figs. 9.23 &
9.24), horizontal equilibrium requires:
Pm Ap p As s N c1 N c2 0
or
dx d x 1 c
Pm Ap1 c Ep As c Es f cd bf x
x x 2 1, 75 103
(9.40a)
1 c0
f cd bf bw x hf 0
2 1, 75 103
where:
x 200
c0 c (9.40b)
x
It is noted that the concrete stress-strain relationship from fig. 9.24 refers to ULS design.
It is strictly speaking not applicable in a SLS check. Since concrete is usually in the linear
elastic stage in SLS, the only relevant concrete property in SLS design is the Young's
modulus. This modulus can be read from EN 1992-1-1 table 3.1. However, this table
presents a modulus related to short term loading only. Long term loading can be
incorporated by taking into account the creep coefficient. As demonstrated in Chapter 6,
creep is often difficult to quantify precisely. This is partly caused by the difficulties
encountered when predicting the loading history (short and long term load components
and their duration). Therefore, in practice often estimated values are used. This is also
done in this case study, where the Young's modulus of concrete is derived from the ULS
stress-strain diagram. As mentioned before, this is strictly speaking not correct and a
precise calculation should reveal the magnitude of the creep coefficient.
35
f cd 1,5
Ec 13,3 103 N/mm 2
c3 1, 75 103
EN 1992-1-1 table 3.1 shows Ec = 34 · 103 N/mm2 for C35/45. This implies that
implicitly a creep coefficient of 34 / 13,3 - 1 = 1,6 is included in the analysis.
To have equilibrium of moments, the following condition should be met when using the
centroidal axis as reference:
or
1 c 1 c0
M max f b x e1
3 cd f
f cd bf bw x hf e2
2 1, 75 10 2 1, 75 103
(9.41a)
dx d x
Ap1 c Ep d zt As c Es d zt
x x
where
e1 zt 13 x (9.41b)
e2 zt hf 13 x hf (9.41c)
The bond factor of the tendons, consisting of strands, is ξ = 0,5 (EN 1992-1-1 table 6.2;
also see table 9.6). The ratio of the diameters of the prestressing and reinforcing steel also
must be included to find the actual bond factor ξ1. EN 1992-1-1 cl. 6.8.2 prescribes:
Øp 1, 6 600 39 mm
The bond factor including the influence of the diameter (eq. (9.37)):
bmp Øs Ø 12
1 s 0,5 0,39
bms Øp Øp 39
In the eqs. (9.40a-b) and (9.41a-c), two unknown variables exist, namely c and x.
Solving the equations results in:
e1 = 248 mm
e2 = 115 mm
x 350
The stress in the steel is s Es = 286 N/mm2, which also follows from the previously
presented steel force (Ns / As).
The same calculation procedure is used to calculate the steel stress directly after cracking.
The applied bending moment is Mcr = 328 kNm.
Results:
e1 = 168 mm
e2 = 35 mm
x 590
The height of the hidden tensile member is 2,5 (h - d) but not greater than (h - x ) / 3.
h = 1000 mm
d = 850 mm
x = 350 mm at the SLS
1 f ctm Ø 1
wmax = σs σsr εcs Es
2 bm ρ p,eff Es
where = 0,5, bm = 2 fctm, s = 286 N/mm2, sr = 52 N/mm2 and Ø = 12 mm. No
shrinkage will be included (short term loading).
1 1 12 1
wmax = 286 0,5 52 0 0,13 mm
2 2 0, 030 200 103
In this calculation, x and c can be estimated to quickly verify whether the crack width
criterion might be critical. To present a procedure, the stress distribution is presented for
the maximum bending moment in SLS assuming that the concrete would not crack. The
stresses at the top and the bottom of the beam are (fig. 9.25):
2
11,0
-10.8N/mm
N/mm2
Nt
458
466 mm
9.1 N/mm2 2
9,6 N/mm
After cracking, the tensile force has to be taken over by the steel. Due to cracking, the
internal lever arm increases. Therefore, the force that has to be resisted by the steel, is
smaller than Nt. A value of 0,9 Nt can be regarded as a conservative estimation.
If s = 305 N/mm2 is substituted in eq. (9.39), it becomes clear that the crack width
criterion is met. On the basis of this result it can be decided immediately that a more
accurate calculation is not necessary.
Apart from the control of the crack width at the level of the main reinforcement at the
bottom of the beam, it should be avoided that large cracks in the web occur, see fig. 9.13.
Figure 9.26 gives a design aid that shows up to which height crack distributing
longitudinal web reinforcement is required [9.8].
In the case considered, h = 1000 mm and h – x = 1000 – 350 = 650 mm. From fig. 9.26 it
follows that for this value, in combination with the requirement wmax < 0,2 mm, hw ≥
200 mm. This condition was already met, see fig. 9.22, so that no further crack
distributing reinforcement has to be provided.
1000
hw [mm]
1 2
800 0. 0. 0.4
w= w= w=
x-c
600
h-x-c
400
hw
200
0
0 400 800 1200 1600 2000
h-x [mm]
Fig. 9.26 Height of the web that should be provided with crack distributing
reinforcement, in order to avoid uncontrolled cracking in the web [9.8]
Fig. 9.27 Effective tension area of a longitudinal bar in the web of the beam
For more information on crack width control in high beams, reference is made to [9.9].
EN 1992-1-1 cl. 7.3.3 (3) states that a crack width calculation for the web can be carried
out by assuming that it behaves as a member loaded in pure tension at a steel stress that is
half the steel stress in the main (longitudinal) reinforcement.
9.11 Literature
9.1 Kay. E.A., Fookes, P.G., Pollack, D.J.: “Deterioration related to chloride ingress”,
Concrete, Nov. 1981, pp. 22-28.
9.2 Falkner, H.: “Risse im Beton – Theorie und Praxis”, Vorträge Lindauer
Bauseminar 1985, Veröffentlichung Bauakademie Bieberach, Band 38.
9.3 Bruggeling, A.S.G.: "Structural concrete: science into praxis“, Heron, Vol. 32,
No. 2, 1987.
9.8. CEB Manual on "Cracking and deformations", Bulletin d’Information No. 143,
1981, pp. 2.1-2.80.
9.9 Braam. C.R., “Control of crack width in deep reinforced concrete beams”, Heron,
Vol. 35, 1990, No. 4.
9.11 Falkner, H., "Zur Frage der Rissbildung durch Eigen- und Zwangsspannungen
infolge Temperatur in Stahlbetonbauteilen", Deutscher Ausschuss für Stahlbeton,
Heft 208, 1969.
10.1 Introduction
Professional detailing of prestressed concrete structures not only results in more easy
construction, but also has a positive influence on the quality of the structure. By skillful
detailing, not only the amount of damage will reduce, but also the repair costs during the life
time of the structure will be minimized.
On the basis of literature [10.1] – [10.4], it turns out that a large part of the damage can be
traced back to lack of understanding of the flow of forces in the structure and the ignorance
of a number of simple basic rules for design and construction.
It is possible to avoid part of the errors that lead to damage, namely by representing the flow
of forces using strut and tie models and by using the equivalent prestressing load method
(load balancing method; prestressing is preloading). Such representations can be the basis for
detailing. This principle is used in this chapter.
In the case of prestressing with post-tensioned steel, the tensile forces in the prestressing
tendons are introduced in the member by prestressing anchors. The concentrated forces
gradually spread in the structure. The length required to have a uniform load distribution, is
called the disturbance length of de St.-Venant. This disturbance length s is independent of the
magnitude of the force, but depends on the geometry of the member and the position(s) of the
prestressing anchor(s). According to de St.-Venant, the length s is equal to the largest width
across which the load has to be spread.
Figure 10.1 gives an example of a centrically introduced prestressing force, where the stress
trajectories are represented by a strut and tie model. From the truss model shown in fig. 10.1,
it turns out that the splitting tensile force Nspl is required to have force equilibrium in
transverse direction. The magnitude of this force Nspl follows from equilibrium:
h
N spl 14 Fp 1 1 (10.1)
h
It is noted that the force is not denoted as P, but as F, since the theory is not exclusively
applicable to prestressing forces, but holds for the introduction of forces in general.
Generally, the anchor force must be distributed in both the height and the width direction (the
splitting tensile force for the width direction is found by replacing h and h1 in eq. (10.1) by b
and b1). Splitting reinforcement is then required in two principal directions. In order to carry
the splitting tensile force Nspl, stirrups or mesh reinforcement can be used. In a circular cross-
section one can also choose a spiral reinforcement (This reinforcement should not be
confused with the spiral reinforcement that is used as confining reinforcement in an
anchorage system to strengthen the concrete to resist the high compressive stresses under the
anchor plate).
sh = Fp/Ac
Fp/2
h/2 b1
Fp/2
Nspl
h1
Fp/2
Fp/2 h/2
h/2 h/2
b
a b
Fig. 10.1 a. Truss model to determine the splitting tensile force Nsp
b. Splitting tensile reinforcement
The cross-sectional area Aspl follows from the relation Aspl = Nspl / s, where s is the
maximum steel stress allowed. It makes sense to choose a stress level s below the design
yield stress fyd. If s is equal to the design yield stress fyd, the equilibrium requirement from
ULS is met, but the crack width in SLS might be too large. EN 1992-1-1 cl. 7.3.3 presents (in
Tables 7.2N and 7.3N) a good indication for the maximum stress in the steel allowed to avoid
crack width problems (s = 200 – 400 N/mm2, depending on maximum crack width allowed).
The splitting tensile reinforcement Aspl should be distributed. It is preferred to use much bars
with a small diameter instead of a smaller number of bars with a larger diameter, see fig.
10.1b. The splitting tensile reinforcement should confine the cross-section and be anchored
such that it is also effective at the outsides of the cross-section.
In practice, the prestressing force is often introduced by more than one tendon (fig. 10.2a).
By using more tendons, the splitting tensile force is reduced. Figure 10.2a shows a truss
model to determine the splitting tensile forces and their location. The splitting tensile
reinforcement As1 follows from the splitting tensile forces in the same way as shown before.
It makes sense to apply also some reinforcement at the free end of the beam (As2). This
reinforcement is required to resist unforeseen loads like impact loads or forces from
differential shrinkage. Furthermore, in the stage of prestressing, splitting forces along the free
end may occur. Their magnitude and location depends on the sequence of prestressing. The
stress distributions at the concrete surface following from linear elastic analyses indicate that
the surface tensile force between two point loads (Fp each) is about 0,1 Fp (Dutch code NEN
6720)
Fp/2
Fp/2
Nspl Fp
Fp/2 Fp/2
Fp
Fp
a b
Also when the prestressing force is introduced eccentrically, the position of the splitting
tensile force and its magnitude can be determined with a strut and tie model, see fig. 10.3. In
this case, the disturbance length is about equal to the beam height h because the force has to
be spread over this distance. The stress distribution at the edge of the disturbed area can
easily be determined, because here principally plane sections remain plane (Bernoulli). By
orientating the compression bars (struts) in the model such that they follow the compression
trajectories, the most efficient equilibrium system is obtained. From the forces in the tensile
ties T and the allowed steel stress s, the required cross-sectional area of the splitting
reinforcement is obtained.
Fp
h1
D -
T D x
h
T
T
0
+
sh
Fig. 10.3 Strut and tie model for eccentric load introduction
h1
D1
h1 -
T D2
x
D3
T
h1
Fig. 10.4 Flow of forces for a number of eccentrically introduced prestressing forces.
The TT-beam is provided with a transverse end beam. Due to prestressing, tensile forces
occur in this transverse beam too. Therefore, in transverse direction always sufficient
reinforcement should be provided. In large bridge beams it might even be necessary to apply
prestressing in transverse direction.
P
P
Fig. 10.5 Cracking in a transverse beam at the end of a TT-girder, caused by 3D-
spreading of the prestressing forces P
In prestressed TT-beams without transverse end beam, the tensile force caused by the
introduction of the prestressing forces can only occur in the deck. The introduction of the
forces occurs in almost the same way as in case of a rectangular cross-section, see fig. 10.6.
FpP FpP
A similar type of problem can occur in an I-shaped cross-section in which only the flanges
are prestressed. When the spatial spreading of the forces is not accounted for, cracks with
large widths occur at the end of the beam, see fig. 10.7.
cracking
Parts of the structure outside the influencing zone of the prestressing force partially restrain
the deformations caused by prestressing in other parts of the structure, see fig. 10.8. Those
parts have to stay connected to the part that is stressed by prestressing (compatibility
requirement). In this respect, attention has to be given to the support areas as well. Most
girders are supported as close as possible to their ends. A part of the supporting forces then
might act outside the influencing area of the reinforcement. It should then be assured that the
support reaction is introduced without causing damage to the structure, see fig. 10.9.
Fig. 10.8 Reinforcement partly outside the influencing area of the prestressing force
Introducing the prestressing force Pm0 from the steel into the concrete by bond, requires a
transmission length lpt (EN 1992-1-1 cl. 8.10.2.2 & eq. (8.16)). Further, a dispersion length
ldisp can be distinguished (EN 1992-1-1 fig. 8.16), which is required to develop a linear
distribution (over the cross-section) of the concrete stress caused by the prestressing force
(fig. 10.10).
h
a
c c
lpto
e
ldisp
prestressing
prestressing force
force Fp Pm0
1
/3 lpt
~1/3
a
o
a bond stress a
Fig. 10.10 Transmission length lpt and dispersion length ldisp in a beam prestressed with
pre-tensioned tendons
The centre of gravity of the bond stresses is at about 1/3 lpt from the end of the beam.
For the dispersion length ldisp it can be written:
where lpt is the transmission length and s is the interference length of de Saint-Venant. As
boundary condition it should hold:
EN 1992-1-1 cl. 8.10.2.2 provides the following expression for the transmission length lpt:
pm0
lpt 1 2 Ø (10.4)
f bpt
where:
α1 depends on the way of releasing (gradual or sudden);
α2 depends on the type of tendon;
where:
ηp1 depends on the type of tendon and the bond situation at release;
η1 depends on the bond conditions;
fctd(t) is the design value of the concrete tensile strength at time of release.
As an example, the transmission length of a 9,3 mm (3/8") strand (7 wires, see table 2.3) in a
C50/60 concrete is calculated. It is assumed to be released when the concrete has reached a
strength class C28/35. The strand is at the bottom side of the element. The Y1860S7 strand is
assumed to develop its maximum allowed initial stress σp0 (EN 1992-1-1 cl. 5.10.3):
σp0 = min.(0,75 fpk; 0,85 fp0,1k) = min.(0,75 · 1860 ; 0,85 · 1674) = min.(1395; 1423) =
1395 N/mm2.
1,94
f bpt p1 1 f ctd (t ) 3, 2 1, 0 4,14 N/mm 2
1,5
pm0 1395
lpt 1 2 Ø 1, 0 0,19 9,3 595 mm
f bpt 4,14
A short transmission length is favourable for the anchorage capacity but unfavorable with
respect to splitting action. Short transmission lengths result in a more concentrated transfer of
forces and, as a result, in more concentrated splitting forces which in turn lead to higher
splitting stresses. To have sufficient safety with respect to splitting, it is now assumed that
the splitting forces are introduced over half the transmission length from eq. (10.4).
Now the method will be discussed that enables the check and design of splitting action in a
structure prestressed with pre-tensioned steel.
Three different types of tensile stresses can be distinguished over the dispersion length,
namely (see fig. 10.11):
Spalling stresses.
Bursting stresses.
Splitting stresses.
Fig. 10.11 Different types of tensile stresses in concrete prestressed with pre-tensioned
steel (CEB bulletin no. 181, 1987)
Spalling stresses
These stresses are generated by the eccentric introduction of the prestressing force, see fig.
10.3. Figure 10.12 shows typical cracks caused by the spalling stresses. The tensile spalling
stresses are close to the loaded end surface and are most affected by tendon eccentricity.
They occur away from a single tendon or between tendons in case of multiple tendons.
A graphical method has been developed by Den Uijl1. This method is equivalent to the theory
of elasticity and uses the diagram shown in fig. 10.13, where k is the core radius, wspl is the
width of the cross-section at the level considered and ep0 is tendon eccentricity. The spalling
stress σspl is determined as follows: Starting from (ep0 - k) / h and lpt / ep0 the value for
σspl · wspl · ep0 / Fpi can be read from the diagram. After substitution of the known values for
wspl, ep0 and Fpi, the spalling stress σspl is obtained. In a calculation according to EN 1992-1-1
the force Fpi is Pm0. The maximum spalling stresses are considerably reduced as the
transmission length increases.
yy
spl.wspl.epo/Fpi
0.12
lpto//eeop0==00 1 sl
2
σspl spalling
splitting spl
stress
stress
0.10 3
4 0
x
0.08 5
6 wspl
y
0.06
8
0.04 10
x
h
15
k
epo
0.02 20
epo-k
50 Fpi
0 lo
0 0.04 0.08 0.12 0.16 0.20 0.24
(epo.k)/h lpt
(ep0 – k ) / h
Fig. 10.13 Diagram according to den Uijl to determine the spalling stress spl
It is also an option to use the equivalent prism analogy. The height of this prism (see fig.
10.14) follows from the condition that no shear stresses are transferred along the line B-B.
Therefore, the resulting axial force on the prism has to be zero. The length of the prism
follows from:
2
0, 6 lpt
lspl Cspl h where Cspl 1 (10.6)
h
1
Uijl J.A. den, ”Verbundverhalten von Spanndraht – Litzen im Zusammenhang mit Rissbildung im
Eintragungsbereich”, Betonwerk + Fertigteiltechnik, Volume 1, 1985, p. 28-36.
The spalling force can be determined by using the assumption that the horizontal internal
lever arm zspl between the resulting tensile and compressive forces (see fig. 10.14) is 0,5 lspl,
from which it follows that:
2M
N spl
lspl
M M
B B B
B
h yy Nspl
x
Nspl
B B
spl
Zspl 0,5. spl
Fig. 10.14 Analogue equivalent prism to determine the transverse reinforcement required
when prestressing with pre-tensioned steel
The bending moment M follows from the elastic stress distribution above the line B-B, at the
end of the equivalent prism (at a distance lspl from the end face of the beam). The spalling
stresses σspl can be calculated from the spalling force Nspl by assuming that Nspl is linearly
distributed over a quarter of the prism length lspl:
N spl 16 M
spl
1
2 14 lspl b 2
lspl b
where b is the width of the cross-section at the level where the spalling stresses occur. In case
of expected cracking, the total spalling tensile force Nspl has to be carried by transverse
reinforcement.
Bursting stresses
The tendon force is gradually transferred to the concrete by bond forces. The bursting
stresses are related to the longitudinal component of these forces. The bursting stresses are
perpendicular to the beam axis. They are around the tendon and at a distance from the beam
end. Their occurrence can be explained as follows: In case of a concrete specimen subjected
to a concentrated load, tensile stresses perpendicular to the specimen’s longitudinal axis
occur, see fig. 10.1. The transverse tensile force is in equilibrium with the transverse
compressive force that occurs close to the point where the load is introduced. This transverse
stress distribution is well known from, e.g., a tensile splitting test on a cylinder or cube. All
individual longitudinal bond force components present along the transfer length, cause
individual (transverse) bursting stress distributions. Their combined action is a bursting stress
distribution that results in compressive stresses at the beginning of the transfer length and
tensile stresses at a distance from the beam end (see fig. 10.11), in the anchorage zone. A
bursting crack will, therefore, occur at a distance from the beam end.
Next to the bonding stresses between the concrete and steel, the wedge action exerts a
pressure on the concrete (σr in fig. 10.15). As a result this internal pressure generates
tangential tensile stresses σt. The stresses develop in transverse direction.
The splitting stresses reach their maximum value where the bond stresses are highest. This is
close to the beam end, see fig. 10.11.
The splitting stresses are sufficiently accounted for when the reinforcement required for
bursting and spalling confines the tendons. If no confining reinforcement is present, the
concrete cover should meet certain requirements.
When the concrete cover on the strands and the mutual distance between the strands are large
enough, splitting and bursting are prevented to occur. On the basis of a linear-elastic stress
analysis, general rules can be derived. EN 1992-1-1 cl. 4.4.1.2 states that with respect to
transmitting bond forces safely, the cover on the prestressing wires or strands has to be at
least:
1,5Ø for strands and plain wires;
2,5Ø for indented wires.
EN 1992-1-1 cl. 8.10.1.2 & fig. 8.14 provide information on the arrangement of pre-
tensioned tendons. Minimum horizontal spacing is max. (2 Ø; 20 mm; max. aggregate size +
5 mm); minimum vertical spacing is max. (2 Ø; 20 mm).
r t
a tm
r
r
t
a = bond stress
r = radial compression stress
t = tangential tension stress (in circumference direction)
After cutting a wire or strand, it slips into the concrete. During this process of slipping, the
strands/wires expand in lateral direction due to the loss of lateral contraction (the so-called
effect of Hoyer; wedge action). The slipping stops as soon as a sufficiently large frictional
force has been built up along the tendon. This equilibrium state is disturbed when the steel
stress increases because of external loading, e.g. when cracking occurs. This is now
discussed.
The stress development in the prestressing steel, present over the transmission length lpt, is
then part of the limit criterion for anchorage failure. Tests show that it is sufficiently accurate
to assume a linear steel stress along the transmission length. The steel stress increases from
zero to σp0 over the distance lpt.
A B
o
lpt
pl
σp0
A
B
p
Fig. 10.16 Stress increase in the steel near a flexural crack but outside the transmission
zone
When next to the transmission zone lpt a flexural crack occurs, another mechanism is
introduced. At an increase of the stress level in the steel, bond stresses are generated at both
sides of the crack. The stress peak in the steel at the crack decreases over a certain distance as
shown in fig. 10.16. The stress peak Δσp in the flexural crack cause a stress increase in point
A at the end of the transmission length. In that case, the limit equilibrium state over lpt is
disturbed. The wire or strand then slips and anchorage failure occurs. When the stress peak
has just damped out in point A, the top of the peak (point B) is on the envelope for anchorage
failure, which is indicated by a dashed line in fig. 10.16.
EN 1992-1-1 cl. 8.10.2.3 provides information to determine the envelope for anchorage
failure (see also fig. 10.17).
The length lpt of the transmission zone follows from eq. (10.4). However, an additional factor
1,2 must be introduced to be on the safe side. Therefore, the transmission length to be used in
this analysis is lpt2 = 1,2 lpt. Further it holds for the total transmission length in case of a steel
stress increase to σpd (EN 1992-1-1 eq. (8.21)):
2Ø pd pm
lbpd lpt 2
f bpd
where fbpd follows from the actual concrete strength class in ULS.
Fig. 10.17 Ultimate limit state with respect to anchorage failure (EN 1992-1-1 fig. 8.17)
A = tendon stress; B = distance from end
In the example from the previous section with the 9,3 mm strand it was found that:
It is now assumed that the full design value of the tensile strength of the prestressing steel is
activated (which is a conservative approach):
where:
For C50/60: fctk,0,05 = 2,9 N/mm2, fctd = 2,9 / 1,5 = 1,9 N/mm2.
The initial stress σp0 was assumed to be 1395 N/mm2. If it is assumed that time-dependent
losses are 150 N/mm2, σpm∞ = 1395 - 150 = 1245 N/mm2.
The check with respect to anchorage failure can now be done with the aid of fig. 10.18. At
the design load, the distance x from the support to the first flexural crack is calculated. EN
1992-1-1 (cl. 8.10.2.3 (1)) uses the characteristic tensile strength of the concrete (fctk,0,05) to
calculate the cracking moment. (Note that the cracking moment also has a component from
the axial prestressing force.) The steel stress in this first crack follows from:
M x Vx d cot N p,x
N p,x ; p,x
z z Ap
where Np,x follows from the equilibrium of moments about S, as shown in fig. 10.19.
Because the equilibrium of moments is set up about a point S in a shifted cross-section (shift
is d cotθ with respect to position x) it is often denoted as the parallel axis theorem. The
parameter z is the distance between the working lines of Np and Nc and is equal to the sum of
the eccentricities ep and ec (see fig. 10.19).
The result of this analysis is a prestressing steel stress σpx (fig. 10.18), calculated using the
bending moment at the cross-section x + d cotθ from the support. The prestressing steel stress
can develop provided there is sufficient anchorage capacity to transfer the prestressing force
to the concrete. A distance x is available to transfer the force.
x d.cotg
px
σp0 pl
f
fpkpu/ 1,1
If σp,x is inside the envelope from fig. 10.18, sufficient anchoring capacity is guaranteed. The
prestressing steel force caused by the bending moment then is smaller than the force that can
be generated by bond; the strand is not pulled out. The positions of σpx and the envelope in
fig. 10.18 demonstrate that this is the case: There is sufficient anchorage length to anchor a
force that is somewhat higher than the force associated with σpx (fig. 10.18).
x d cot d cot
N
N cc
Vx S
crack Mx ec
VVx x+dcot
d cot x
z
centroidal axis ep
N p,x
prestressing steel
section of beam free-body diagram
support prestressing
continuous prestressing
Fig. 10.20 Box girder with prestressing tendons only in top and bottom flange
In such a case, the prestressing tendons must be anchored at different locations along the
girder (fig. 10.20). For the detailing of such intermediate anchors, four effects have to be
considered. Ignoring them might cause to damage (fig. 10.21):
1. The cover can spall-off from the prestressing tendons because of the transverse pressure
from the curved tendon (action 1 in fig. 10.21).
2. Splitting cracks can occur due to the spreading of compressive stress trajectories (action 2
in fig. 10.21).
3. The corbel can crack-off (action 3 in fig. 10.21).
4. The corbel and its adjacent concrete area deform as a result of the introduction of the
prestressing force. From compatibility requirements, the area behind the corbel is forced
to follow these deformations. As a result, tension occurs behind the corbel. Ignoring this
effect can result in large cracks behind the corbel (action 4 in fig. 10.21, and the cracking
from fig. 10.22).
1
3
2
The reinforcement should therefore be designed such that all those types of damage are
prevented. Figure 10.23 shows the various types of reinforcement required to avoid
problems. The functions of the reinforcement are:
As1 Tensile tie as main reinforcement in the corbel.
As2 Splitting reinforcement in relation to the spreading of compressive stress
trajectories (compare fig. 10.2).
As3 Reinforcement for an unforeseen pressure by curvature of the prestressing
tendons (it might be that the construction drawings indicate a straight tendon
profile, whereas due to ignorance at the construction site, the tendons have a
curved profile).
As4 Reinforcement for the (planned) force because of tendon deviation.
As5, As6 Suspension reinforcement for limiting the crack width behind the corbel (control
of compatibility cracking).
As6
As5
As4
As3
As2
As1
Fig. 10.24 Tendon curvature associated with the assembly of the reinforcement
Figure 10.24 shows a beam with an I-shaped cross-section. In the middle of the beam, the
tendons are positioned as low as possible in the cross-section to have a maximum internal
lever arm. At the end of the beam, the position of the anchorages is determined by
geometrical conditions. A complication arises because the tendons must be introduced in the
lower flange, and therefore must pass the stirrups that provide part of the shear force
resistance. Since the allowable curvature of prestressing tendons is relatively small (a high
radius), the shape of the stirrups has to be adapted over a considerable length. It is
questionable whether optimizing the position of the tendons is worth this effort. When in the
middle of the beam the tendons are positioned above each other instead of besides each other,
the internal lever arm is somewhat reduced, but construction is much easier.
It often occurs that one is not aware of the effect that the curvature of prestressing tendons
may have. Curved prestressing tendons do not only introduce compressive stresses but may
also introduce unintended tensile stresses. Figure 10.25 shows such a case. The figure shows
the bottom part of a box girder bridge. At the supports, the prestressing tendons have a
position high in the cross-section, whereas in the span, they are positioned in the bottom
flange to have a maximum internal lever arm. However, this implies that the tendons should
run from the top of the girder, through the webs, to the bottom flange. Then, they not only
have an upward curvature (to introduce the load balancing forces into the structure), but also
have a curvature in transverse direction (fig. 10.25a). This introduces forces in transverse
direction which can initiate longitudinal cracking in the bottom flange.
A similar effect occurs if the webs of a box girder are inclined (fig. 10.25b). Also in this case
the tendons have a (projected) curvature in the horizontal plane, which results in transverse
tension in the bottom flange [10.4].
a cracks b
NT
Fig. 10.25 Possible cracking due to the transverse loading effect from curved prestressing
tendons
Special attention has to be given to the effect of tendon curvature in curved bridges. Figure
10.26 shows the top view of such a structure [10.1]. In this case, the prestressing tendons are,
as is usual in statically indeterminate structures, positioned high at the supports and low in
the spans. In this case altogether 12 prestressing tendons are applied. They are continuous
over the full length of the bridge. In each of the three webs of the 2-cell box girder, 4 bundled
prestressing tendons are applied. They are stressed from both ends of the bridge. During the
prestressing operation of the last tendons four tendons broke out sideways over almost a full
span of the bridge (fig. 10.27). The cause of the damage was a combination of a relatively
high tendon curvature (small radius) and a small spacing between the high capacity tendons.
The damage would not have occurred, if the concrete cover would have been larger, and the
tendons would have been more uniformly distributed.
Fig. 10.27 Compressive forces on the concrete cover caused by a curved profile of the
prestressing tendons
When a relatively flexible tendon duct is supported at a too high spacing, or when the
tendons have been pushed down during construction (e.g. by walking over the duct), an
unintended curvature might result. During prestressing, the tendon is straightened and forces
act on the wall of the duct. This can result in spalling of the concrete cover (fig. 10.27).
Figure 10.28 shows damage that occurred during the construction of a lightweight concrete
bridge because of unintended curvatures.
Fig. 10.28 Push-off of concrete cover caused by unintended tendon curvatures in the deck
of a lightweight concrete bridge
Another problem concerns the hydration heat from hardening concrete. As a result of the
chemical process of hardening, the temperature of the concrete can considerably rise. When
the concrete starts to cool down, it implicitly shortens. New concrete is often cast against
older concrete that has hardened already. The shortening process of the new concrete is then
restrained by the older concrete. This results in tensile stresses and might cause cracking (fig.
10.29a) if the tensile stresses develop faster than the concrete tensile strength (fig. 10.29b).
By insulating the hardening concrete (fig. 10.29c), the cooling process is slowed down, so
that the tensile stresses develop slower (fig. 10.29d). The formation of temperature induced
cracks can then be prevented.
The tensile strength of the concrete in a construction joint is mostly lower than that of the
adjacent concrete. Therefore, it is necessary to apply a well distributed reinforcement (ρs =
0,7 – 1,0%, bar spacing ≤ 100 mm) through the joint. In the youngest concrete additional
reinforcement is required because of stresses induced by restrained shrinkage.
a b
stress
fcct
c
T = 20° T = 70° c
cracks
c insulation
isolation
d
stress
fcct
c
If a construction joint is used for coupling prestressing tendons, another unfavorable effect
can occur. Long bridges are usually built in stages and couplers are mostly applied in joints
close the points of zero moment (fig. 10.30a). The deformations occurring in the following
stages are schematically shown in fig. 10.30b-e.
At first, the part that has hardened already is prestressed, see fig. 10.30c.
Subsequently, the prestressing tendons are coupled and a new section is cast, see fig. 10.30d.
After hardening of the concrete, also this new part is prestressed. The prestressing force,
introduced at the end of the next section, concentrates on the coupler in the construction joint
(fig. 10.30e). This, however, will be accompanied by deformations, which may result in
cracking at the ends of the construction joints.
The probability of cracking is higher, the more the couplers are concentrated at a small
number of positions. To prevent cracking, the couplers should be spread well over the cross-
section, so that their mutual distance is small. In the area between the couplers, reinforcement
for crack width control should be applied, having a length equal to the distance to the nearest
coupler, see fig. 10.31 [10.4].
k k
couple anchor
k surface reinforcement
k
k
k k k
k
web
3ap
primary crack crack width limitation
bottom plate reinforcement
6ap
prestressing cable
ap
ap a0
Fig. 10.31 Crack width control in the area close to a coupling joint [10.4]
Schlaich and Schäfer [10.7] suggest two solutions, see fig. 10.32b and c. In the drawings the
dashed lines represent compression struts whereas the solid lines represent tensile ties.
One solution, according to the principle shown in fig. 10.32c, has been worked out in fig.
10.33a.
a b
c
T1 T1
Fig. 10.32 Indirect support at an intermediate pier of a statically indeterminate box girder
bridge [10.7]
Figure 10.33a shows how the inclined compressive force D in the web of the box girder is
carried by a vertical tensile tie T1 and transferred to the top of the girder. Tensile tie T1 can be
constructed from straight prestressing tendons. As a next step, the force, which is now at the
top of the transverse internal diaphragm, has to be transmitted to the bearing. The
representation as a strut and tie model shows that this requires a second tensile tie T2. To this
aim, curved prestressing tendons can be applied, see fig. 10.33b and c.
T2
b
a
D D
T2
c
T1
T1
It should also be planned carefully in which sequence the prestressing tendons are stressed.
The beam shown in fig. 10.35 is used as an example. If all tendons present in one web are
stressed, whereas the tendons in the other web stay passive (unstressed), the deformation at
the stressed side can give rise to tensile stresses which might lead to cracking. Also when at
both sides the lowest tendons are stressed first, the high eccentricity of this (temporary)
prestressing force can be the cause of cracking at the upper side of the structure. In such a
case it is advised to first stress the most centrically positioned tendons.
All beams are prestressed using 4 tendons. The maximum support reaction force at SLS is
1550 kN. The effect of dynamic loading should be taken into account. In order to avoid
damage caused by fatigue loading, it was required that the beams, even at maximum load,
should be uncracked. This requirement was governing for the design of the beam ends.
Dapped-end beams were used. They are mostly designed using strut and tie models. In
general, two solutions are possible, see fig. 10.38.
a b
T2
T1
Fig. 10.38 Two solutions for the beam end using strut and tie models
Often, these two basic solutions are combined. In the case considered, however, a number of
restrictions applied. The four prestressing tendons should be anchored at the beam ends. It
would have been a poor solution to anchor all four tendons either at the bottom or at the top
of the cross-section: the high capacity tendons require heavy anchors, which require a large
area. Furthermore, an eccentric introduction of the prestressing force would result in a creep
gradient, possibly resulting in cracking. Therefore, it was the most logical solution to anchor
two tendons at the top and two at the bottom of the cross-section. The vertical ties (T1 or T2)
can only be constructed from active prestressing steel because it was required that the
structure is uncracked at SLS. Because of the stress concentrations in the corner, the
prestressing steel should be provided as close as possible to this corner. For the control of the
tensile stresses in the corner, inclined prestressing tendons would give the best results. This is
an argument to apply model b from fig. 10.38. This would simultaneously solve another
problem, namely the sound detailing of the lower node of the truss model a from fig. 10.38.
This problem is illustrated in fig. 10.39.
T1
wrong
Therefore, only the solution shown in fig. 10.38b was acceptable. The corner was smoothly
curved to reduce stress concentrations. The principle of the solution chosen is shown in fig.
10.40.
The stirrups mark A have a double function. First, they act as splitting reinforcement for the
two lower prestressing tendons. Second, they resist the force in the compressive diagonal
(strut) D. The required amount of stirrups mark B follows from the shear force design. The
stirrups mark C and the hairpins mark A confine the corbel part of the beam. The prestressed
vertical tendons are designed such that the prestressing force is higher than the tensile force
in SLS from the strut and tie analysis. Thanks to this prestressing force, no cracking in SLS
was observed, even not in the corner. Figure 10.41 shows an overview of all the
reinforcement applied.
Fig. 10.41 Overview of the reinforcement in the end region of the beam
The safety with regard to failure was determined as follows. A crack was assumed to start at
the corner and to run at an angle (fig. 10.42). It was assumed that all reinforcing bars and
prestressing tendons that intersect the crack reach their yield strength. The depth of the
compression zone was calculated assuming that the concrete in the compression zone reaches
the strain limit. When using the condition of equilibrium of bending moments (M = 0), the
Ap.fpu
RVd,min
u,min
.
Ap fpu
Ap.fpu
RVdd
Because of the large number of identical beams to be used in this project, a test was carried
out. One of the full-scale beams was loaded to failure, see fig. 10.43. The test results
demonstrated that the design calculation was correct.
F6 F3 F1
a=305
50
10
Fig. 10.43 Full-scale test on a beam of the Metro viaduct (crack pattern and crack widths
in ULS; dimensions cm, crack widths in mm)
10.8 Literature
10.1 Podolny, W.: “The cause of cracking in post-tensioned concrete bridge girders and
retrofit procedures”, Journal of the PCI. March-April 1985, pp. 83-139.
10.2 König, G., Maurer, R., Zichner, T.: „Spannbeton: Bewährung im Brückenbau“,
Springer Verlag, 1986.
10.3 Brakel, J., Doorn, L. van, Pol, F.J.M. van de: „Schadegevallen bij brugconstructies in
Nederland“, Stuvo rapport 101, December 1985.
10.4 Leonhardt, F.: “Rißschäden an Betonbrücken – Ursachen und Abhilfe”, Beton- und
Stahlbetonbau, 2/1979, pp. 36-44.
10.6. Uijl, J.A. den: „Verbundverhalten von Spanndraht – Litzen im Zusammenhang mit
Rißbildung im Eintragungsbereich“ Betonwerk + Fertigteiltechnik, Heft 1, 1985, pp.
28-36.
10.7 Schlaich, J., Schäfer, K., Jennewein, M.: „Toward a consistent design of structural
concrete“, Journal of the Prestressed Concrete Institute, May-June 1987, pp. 74-147.
10.8 Blokland, P.: Loading tests on a full-size suspended beam and a model of this beam
for a Metro viaduct at Rotterdam”, CUR-Report 40, 1969.
11.1 Introduction
Serviceability limit state cracks in precast concrete are not harmful from the point of view
of durability provided that their width is small. This finding is of great importance for the
design in structural concrete. It implies that it is possible to combine passive and active
(prestressed) reinforcement. For each application the most suited combination can be
chosen.
At the beginning of the application of prestressed concrete, in most countries only fully
prestressed concrete was allowed. The requirement no tension in the concrete could
hardly be accomplished as it required prestressing in three directions. The result was that
full prestressing was replaced by prescribing that no tensile stresses in the main load
bearing direction were allowed. In this direction, however, principal tensile stresses can
be caused by shear and torsion as well. This alternative definition of full prestressing was
therefore not very consequent: cracks caused by shear and torsion are more dangerous
than well distributed, fine, bending cracks [11.6].
This resulted in the so-called “limited prestressing” concept, where small tensile stresses
were allowed. This method of prestressing has resulted in economically sound structures,
which meet high requirements with regard to durability.
Limited prestressing, however, does not exclude all inconsistencies. There is a general
misunderstanding that the definition applies only to the main load bearing direction, and
that it is therefore not necessary to apply transverse prestressing. This in fact implies that
the structure is prestressed in the main load bearing direction and is reinforced in the
transverse direction. Implicitly, cracking in transverse direction is accepted. The result is
that, as an example, the relatively slender box girder from fig. 11.1 is not prestressed in
vertical and transverse direction. As a result, however, longitudinal cracks can occur in
the webs of the box girder, and especially in the dynamically loaded bridge deck. Under
certain conditions, the cracks in the bridge deck above the webs can penetrate to the level
of the longitudinal prestressing steel. The cracks from loading in the transverse direction
are longitudinal cracks that run parallel to the longitudinal prestressing. These cracks can
coincide over a considerable length [11.6].
Fig. 11.1 (a) Box girder without transverse prestressing, (b) bending moment lines for
webs and top flange
Such cracks are more dangerous from the point of view of durability than the well-
distributed fine bending cracks, which one tries to avoid by prestressing. Furthermore, for
such structures it is not economic to apply prestressing to avoid cracking caused by
transverse moments. Because of the governing traffic load, the transverse prestressing
tendons should be applied almost centrically, because otherwise in the unloaded situation,
tensile stresses might occur in the pre-tensioned compression zone. Furthermore, because
of the small depth of the flange, there is hardly any possibility to use a draped tendon
profile. Full and limited prestressing are therefore not suited, since they would require an
uneconomically large amount of prestressing. The best solution is to apply a well-
distributed reinforcement, combined with a small number of prestressing tendons which
assist in controlling crack widths and deformations. The efficiency of the prestressing
then mainly results from introducing an axial compressive force. This force strongly
reduces the steel stress increase at the onset of cracking and, as a result, limits the
transmission length (see Chapter 9). Moreover, the compressive force might limit the
height of the effective tension area. This smart solution is denoted as “partial
prestressing”, because in SLS fine cracks are accepted to occur. These cracks will only
open at unusually high traffic loads, and will otherwise be (almost) closed.
There are many situations where partial prestressing is the best option, for example slabs
in office buildings. Flat slabs having a span larger than about 6 m can hardly be
constructed in reinforced concrete in an economic way. Full or limited prestressing has
the disadvantage that the governing design load hardly or even never at all occurs. As a
result, under quasi-permanent loading conditions (2 qq, where 2 = 0,3 for offices and
housing; see EN 1990 table A1.1 & NL National Annex), at the bottom of the slabs a
higher compressive stress occurs than at the top (fig. 11.2). Since creep deformation
depends on the magnitude of the compressive stresses, the slab has an increasing upward
deflection (camber) which can result in damage to the partitioning walls on the slabs.
c c
If one only would have the choice between reinforced concrete on one hand and full or
limited prestressing on the other, one would always have deal with the negative effects of
deflections developing over time: in case of reinforced concrete an increasing downward
deflection, for prestressed concrete an increasing upward deflection.
However, most damage can be avoided by choosing the right combination between
reinforced and prestressed concrete (the right degree of prestressing, see section 11.2).
With regard to corrosion, there are no arguments that hamper the application of partially
prestressed concrete in flat slabs. In Switzerland, this way of prestressing has developed
to the standard one. In the Netherlands, the application of partially prestressed concrete
was allowed after the introduction of NEN 6720 in 1990.
In fig. 11.3, the stresses caused by the maximum load q = qg + qq in SLS are shown for
full, limited and partial prestressing.
In the last case, the stress s in the reinforcing steel, or the increase of the stress in the
prestressing steel p, should be limited to ensure that durability requirements are met.
Depending on the bar spacing, EN 1992-1-1 table 7.3N allows steel stresses starting at
160 N/mm2. Crack widths can be limited to 0,4, 0,3 or 0,2 mm, depending on the
exposure class(es).
It is noted that the Dutch National Annex to EN 1992-1-1 reduces the maximum crack
widths by 0,1 mm in case prestressing steel is applied.
Pm
Fig. 11.3 Stress distribution as a result of the total load qg + qq for full, limited and
partial prestressing
M dec
K (11.1)
Mg Mq
where
Mdec is the so-called decompression moment. This is the moment for which the
stress in the outer fiber of the tension zone is 0.
The degree of prestressing indicates the part of the total load the structure can resist
without having tensile stresses at the tension side. K = 0 refers to a structure without
prestressing, whereas K > 1 refers to fully prestressed concrete. The most important
advantage of prestressed structures compared to structures without prestressing, is their
more favorable behaviour in SLS with regard to cracking and deflection. In ULS, the
behaviour of a partially prestressed structure is basically the same as for a reinforced
concrete structure. Therefore, an effective definition of the degree of prestressing should
be linked to the effect of prestressing in SLS. The definition of the degree of prestressing
given in eq. (11.1) satisfies this requirement.
Ap f pd
(11.2)
Ap f pd As f yd
where:
Ap is the cross-sectional area of the prestressing steel;
As is the cross-sectional area of the reinforcing steel;
fpd is the design value of the tensile strength of the prestressing steel;
fyd is the design value of the tensile strength of the reinforcing steel.
The ratio indicates which part of the resistance can be attributed to the prestressing
steel. This definition is relatively simple, but it does not provide any insight into the
effect of prestressing in SLS.
The diagrams from fig. 11.4 show the calculated safety against failure, the total cross-
sectional area of the steel A = Ap + As , the stress in the reinforcing steel s and the
increase of the stress in the prestressing steel p. The diagrams can be used to draw
some important conclusions:
For high degrees of prestressing, the safety factor (i.e. the ratio between the bending
moment from the characteristic loads and the actual bending moment resistance) is
higher than strictly required. (In Switzerland, where this comparison was made, the
margin was = 1,75).
The required cross-sectional area of the reinforcement reaches, in this specific case,
its minimum value for K = 0,6.
At high degrees of prestressing (K > 0,7) the stress in the reinforcing steel s and the
increase of the stress in the prestressing steel p are relatively low. For intermediate
degrees of prestressing (0,4 < K < 0,7), the stresses are still considerably lower than in
traditional reinforced structures.
1000 mm
220
270
300
Md.l.+v.l.
M g+M = 124
=q124 kNm kNm
Ap As
2
failure
collapsesafety
safety
1
0
mm2 300
minimal section reinforcing steel
200 Ap + applied reinforcing
A As cross-sectional
section area of steel
100 Ap,As, Ap+As
As
Ap
0
N/mm2 300
s stress increase in prestressing
200 steel and stress in reinforcing steel
stresses in reinforcing steel s
sp
stress increase in prestressing
100 p
steel p
0
0 0.2 0.4 0.6 0.8 1.0
grade of prestressing k
degree of prestressing K
Often, it will be most economic to choose the values for As and Ap such that As
corresponds with the minimum reinforcement and determine Ap such that the safety ratio
is not unnecessarily high.
Next, results of an experiment carried out at ETH Lausanne [11.7] are discussed. The
slabs tested are statically indeterminate, see fig. 11.5. The figure presents the side view of
one span and the cross-sections at midspan and at the support. In the calculation a linear-
elastic bending moment distribution was assumed. The combination of reinforcing steel
and prestressing steel in the section subjected to the highest moment was chosen such that
the moment resistance is the same in all cases researched. In the research, the definition
of the degree of prestressing according to eq. (11.2) was used, with varying between 0
(slab B4) and 0,76 (slab B2).
Fig. 11.5 Series of tests on prestressed slabs at ETH Lausanne (dimensions cm) [11.7]
In fig. 11.6, the maximum deflection is shown as a function of the load applied. The
influence of the degree of prestressing is clear: already at a relatively low degree of
prestressing (mean axial concrete compressive stress md = 0,86 N/mm2; slab B3; =
0,29), the deflection in SLS is 40% smaller than for the reinforced reference element. If
the prestressing is twice as high, the reduction is about 65%. A similar reduction is found
for the maximum crack widths at the intermediate support (fig. 11.7) and in the span. The
tests confirm that even a relatively low prestressing level can substantially improve the
behaviour in SLS. The vertical dashed lines in fig. 11.6 and 11.7 refer to the SLS / ULS
load ratio of 1 / 1,75 used in Switzerland at that time.
max [mm]
25
20
15
B4
10
B3
B2
wmax [mm]
1.0
0.8
0.6 wmax
B4
0.4
B3 B2
0.2
The first example is a statically indeterminate box girder bridge, having a main span of
41,6 m and two end spans of 32 m each. The bridge is prestressed by 12/12.9 VSL
tendons at an effective prestressing force (1200 - 1350 kN). To have a fully prestressed
structure, 18 tendons would be required; 9 in each web. Figure 11.8 shows the positions
of these tendons in the cross-section at a mid support and at the end span. It is difficult to
position all the tendons in the cross-section. When aiming at the largest possible internal
lever arm, a substantial number of tendons must be in the flanges. The result is that not
only vertical but also horizontal curvatures of the prestressing tendons are required.
Horizontal curvatures should be avoided as much as possible, see fig. 10.25. If, however,
a configuration with all tendons in the webs is chosen, 20 tendons are required because of
the reduced internal lever arm. Furthermore, the high concentration of tendons in the web
could result in problems during casting (sieve effect from the ducts). If the design of the
bridge is based on partially prestressed concrete, only 12 cables are required. The degree
of prestressing then is K = 0,68. To obtain the required resistance, additional reinforcing
steel is only required at midspan. The passive reinforcement is effective in reducing the
increase of the steel stress at cracking, and, as a consequence, the crack width.
The transition from full to partial prestressing offers a number of advantages, such as a
more practical layout of the prestressing tendons and a ULS resistance that is not
unnecessarily high. Moreover, cracks only occur at midspan at full SLS loading and close
again at unloading.
Fig. 11.8 Location of the prestressing tendons in the cross-section of a box girder at the
mid support (1) and at midspan (2), for K = 1 (black and gray tendons) and K
= 0,68 (black tendons only) [11.3]
The second example is a submerged tunnel. The minimum load caused by a 22 m water
column on the roof is 220 kN/m2.
Figure 11.9 shows a cross-section of the tunnel. Such tunnels are mostly constructed in
segments having a length of 100-200 m. The segments are built in a temporary dock that
is inundated after completion of the segments. The segments are provided with temporary
end walls to make them float. They are transported to their destination, where they are
submerged to their final position. This causes large differences in the load on the tunnel
roof in the various stages (construction, transport, submerging, final use). Full or limited
prestressing is not an option since there is no load in the construction and the transport
stage: the tendons would give a high upward load, whereas the vertical load is not yet
applied. Constructing in reinforced concrete is not an option since it would require very
high amounts of steel bars of large diameters that would have to be placed in several
layers. In such conditions, it is very difficult to meet the crack width criteria.
The best solution is to apply partial prestressing. In this specific case, the best solution is
obtained for K = 0,73 at the intermediate support (ρs = 0,27%, ρp = 0,27%).
The smallest concrete compressive stress at the mid support from prestressing and the
force from transverse water pressure is:
for t = 0 (just after prestressing): c = - 3,3 N/mm2
for t = (including losses): c = - 4,1 N/mm2
B.B.R.V. - UR - 500
stirrup
Ø 16
Øk 25-200 Øk 16-200
The first example is a frame of the Hans Martin Schleyer hall in Stuttgart, see fig. 11.11.
The main bearing structure of this hall consists of a number of prestressed concrete
frames. At its end, a frame supports a truss of the roof structure, see fig. 11.11. The
picture shows the position of the prestressing steel in the frame and gives the degree of
prestressing in some important cross-sections. It was obvious to prestress the
cantilevering frame with the high load at its end, especially because its top part is
permanently exposed to outside weather conditions. During design, it turned out that full
or limited prestressing was not very efficient. Applying partial prestressing was
stimulated by the possibility to prestress all the tendons at once. By reducing the number
of tendons and by adding reinforcing steel, all prestressing anchorages could be placed in
the end cross-section of a frame. The required degree of prestressing followed from the
requirement that the frame should be fully prestressed under permanent loads. The ratio
between the permanent loads and the total load was about 2/3. This resulted in the degrees
of prestressing shown in fig. 11.11.
Fig. 11.11 Hans Martin Schleyer Hall: partially prestressed frames with position of
tendons and degrees of prestressing
The second example concerns a viaduct in the highway A7 Ulm – Würzburg [11.5]. The
structure was designed in partially prestressed concrete. The viaduct is shown in fig.
11.12: a skew beam grid bridge having a span of 22,5 m and to be designed for a traffic
class 60 (heaviest vehicle = 600 kN). The main aim of the study was to determine the
most economic degree of prestressing.
In the transverse direction, the optimum degree of prestressing was K = 0,53. This
implied that full prestressing is available up to a load equal to the permanent loads plus
39% of the traffic load. In the longitudinal direction, the most economic degree of
prestressing was K = 0,60.
Using the method of partial prestressing, structures can be designed with excellent
durability (for instance, by not allowing cracks to occur at an intermediate support or by
limiting crack widths) and minimum creep deformation. Furthermore, economic solutions
are possible, providing a resistance that does not (unnecessarily) exceed the specified
value.
When determining the building costs, all aspects, such as the costs of the materials
(concrete, reinforcing and prestressing steel), the required temporary structures
(formwork, scaffolding), transportation, labour et cetera, have to be taken into account.
Therefore, Kupfer and Schulz [11.4] determined the most economic degree of
prestressing of the structures shown in fig. 11.13. The costs are presented as a function of
the degree of prestressing (fig. 11.14). For all types of structures regarded, the optimum is
found for degrees of prestressing in the range K = 0,3 – 0,7. Compared with structures
from reinforced concrete, savings are about 10%.
30600
10 10 10
16 37.5 20
165
115
20
17.5
3x8
16 16
40 31000 40
15
60
54
5 15
165
115
35
3x18 18 bmF = 2.35m
T - girder continous gider
continuous girder
30600
4
22
5
I-girder slab floor
Fig. 11.13 Structures for which the total costs are calculated as a function of the degree
of prestressing K [11.4]
-profile
100
-profile
continuous
continous
90 girder
girder
slab floor
Fig. 11.14 Relation between total costs and the degree of prestressing for the types of
structures from fig. 11.13
An important task for a structural designer is to calculate the required reinforcing and
prestressing steel such that all SLS and ULS requirements are met. The first step,
however, is to estimate the dimensions of the structure. One should realise that
“designing a structure” is in general not a linear straightforward process, but requires a
number of iterations. In this section, only the major steps are presented.
8 Pm, f
qp .
l2
The working prestressing force is estimated at first, for instance 0,8 Pmo. On the basis of
the condition from step 1, the cross-sectional area Ap of the prestressing steel is
calculated.
M Ed g M g q M q p M p
where:
The last component is still open. In each cross-section, the amount of reinforcing steel
applied must ensure that the requirements with regard to the resistance (ULS) are met
(also see sections 7.2 and 7.3).
It should be noted that the flanges of box girders are almost in pure tension, see fig.
11.16.
On the basis of the Swiss Code SIA 162 (1983) a number of design aids have been
derived. They enable a fast control of the crack width (fig. 11.15).
350
[N/mm2]
permisseble stresses
300
pure bending
250 ae/t = 1.0
200
s en p
150
pure tension
100 ae/t = 0 ae/t = 0.5
50
0
0 50 100 150 200 250 300
bar distance s [mm]
Fig. 11.15 Allowed maximum stress in the reinforcement as a function of the bar
spacing s, for which wm 0,15 mm [11.2]
On the horizontal axis of fig. 11.15, the bar spacing s is shown (bonded prestressing
tendons are included in this spacing) and on the vertical axis the maximum steel stress in
the reinforcing steel. The diagram has been derived for a mean crack width of 0,15 mm.
The upper curve is valid for a high stress gradient over the depth of the cross-section
(such as in massive slabs). The lower curve applies when the stress gradient is 0, which
implies pure tension. The stress gradient is defined by the ratio ae / t (fig. 11.16), where ae
is the height of the effective tensile member (compare to fig. 9.14) and t is the depth of
the concrete tensile zone, assuming that the member is uncracked. The ratio ae / t should
be less than 1 (top part of fig. 11.16).
f
heff ae=t
s s
t
ae
maximum ae = 0.2 m
permanent loads only and for permanents loads plus variable loading. The moment lines
are calculated assuming uncracked cross-sections. The cross-sections A, B and C will be
designed.
The design by Bachmann is based on the Swiss code SIA 162 from 1975, using an
overall load factor of 1,75 for both permanent and variable loading. No material factor is
used. Nowadays, codes do not use one overall load factor anymore, but use partial factors
for loads ("effect") and strength ("resistance"). Moreover, different load factors are used
for permanent and variable loading. Nevertheless, Bachmann's approach will be followed
since the example is meant to illustrate the design process.
This working example deals with the main items of partial prestressing. Therefore, only
preliminary design calculations are presented. More detailed analyses must be carried out
to verify whether all requirements are met. Since these analyses are based on the same
principles as the preliminary design, they will not be presented.
Fig. 11.17 View and cross-sections of the bridge (dimensions m; c = centroidal axis; kt =
top of kern area; kb = bottom of kern area) [11.2]
Loads:
volumetric weight concrete: 25 kN/m3
static loads: 2,4 kN/m2
variable loading (traffic): 7,0 kN/m2
Material properties:
concrete C35/45; Ec = 37 · 103 N/mm2
prestressing steel Y1770 (fpk = 1770 N/mm2; fp0,1k = 1570 N/mm2 )
reinforcing steel B500
[kNm]
permanent loads + variable load
permanent loads
The decompression bending moment is chosen such that the governing cross-sections are
uncracked at permanent loading only. The decompression moments MD follow from the
bending moment lines in fig. 11.18:
To estimate the required amount of prestressing steel, the prestressing is first applied as
equivalent loads to the structure. The profile of the prestressing is estimated, see fig.
11.19.
In sections A and C, the tendons are about 0,25 m from the bottom fibre; in section B
about 0,15 m from the top fibre.
It is assumed that the tendons exert an upward load only. (In a detailed analysis, the exact
tendon profile and both the upward and the downward curvatures are used). The drape of
the tendons fA = 1,37 m in both end spans; fC = 1,50 m in the midspan.
The working prestressing force is assumed to be constant along the bridge axis.
8 Pm, f C
at midspan: qp 0, 0068 Pm,
l22
Pm, Pm,
qP = 0,0085 Pm, qP = 0,0068 Pm, qP = 0,0085 Pm,
The condition that no tensile stresses are allowed due to permanent loads results in the
following requirements.
Section A:
Section B:
Section C:
The prestress losses caused by shrinkage and creep of the concrete, relaxation of the
prestressing steel and tendon friction are estimated to be 20%. The initial prestressing
force required then is:
Pm,
Pm,0
0,8
11, 28
Pm,0 14,1 MN
0,8
It is now assumed that the initial prestressing steel stress allowed follows from EN 1992-
1-1 cl. 5.10.3:
min. (0,75 fpk = 1328 N/mm2; 0,85 fp0,1k = 1335 N/mm2 ) = 1328 N/mm2
The minimum required amount of prestressing steel:
1410 4
Ap 10 10, 62 103 mm 2
1328
The required amount of prestressing steel can be calculated if the required bending
moment resistance is known. Bending moments from temperature and differential
settlements (20 mm) are also taken into account and have a load factor 1,0. The moments
from these imposed deformations are:
MA = 1,4 MNm
MB = -2,6 MNm
MC = 1,1 MNm
Also the bending moments from the prestressing loads have a load factor 1,0. The
following bending moments are obtained:
The axial prestressing force and the additional capacity of the prestressing steel must be
taken into account in a cross-sectional analysis.
The friction losses in section A are now assumed to be about zero because of some
overstressing of the steel to compensate for these losses, as well as for the wedge set at
the anchor. The total loss, including time-dependent losses, was assumed to be 20%. If it
is assumed that the friction losses are on average 10% for the bridge, the time-dependent
losses are on average 10% too.
This assumption has impact on the additional capacity of the prestressing steel. It is now
assumed that the friction loss is 0%, 7% and 14% in cross-sections A, B and C,
respectively. The time-dependent losses are estimated to be 10% in all cross-sections.
The working prestressing stress then is:
The additional capacity of the prestressing steel is given in the following scheme. It is
assumed that the stress can increase from the working stress to the stress at the 0,1%
strain limit, i.e. 1570 N/mm2. (A detailed analysis might demonstrate that this is a
conservative assumption).
Figure 11.20 presents the forces in cross-section A. It is assumed that the compressive
force in the upper flange is at 0,1 m from the top fibre. The internal lever arms of the
prestressing and reinforcing steel forces can now be calculated.
loads resistance
20,86
Nc
17,21
Ap
Ap·375
As·500
As
The required amount of reinforcing steel follows from bending moment equilibrium
around the point where the resulting concrete compressive force is:
where:
Pm,∞,A = 14400 · 1195 = 17,21 · 106 N = 17,21 MN and Ap = 14400 mm2 = 14,4 · 10-3 m2.
The lever arm of the axial prestressing force is 0,37 m; it is about 1,6 m for the resulting
force from the reinforcing steel stress and the prestressing steel stress increase, see fig.
11.20.
Result:
Section B:
where:
Pm,∞,B = 14400 · 1112 = 16,01 · 106 N = 16,01 MN and Ap = 14400 mm2 = 14,4 · 10-3 m2.
Result:
Section C:
where:
Pm,∞,C = 14400 · 1028 = 14,80 · 106 N = 14,80 MN and Ap = 14400 mm2 = 14,4 · 10-3 m2.
Result:
Figure 11.21 shows the reinforcement chosen. Small bar diameters are used. They are
distributed along the sides of the cross-section.
In section A, the 24 Ø 20 mm (As = 7540 mm2) are located near the outer fibre and they
surround the tendons (12 bars at the bottom part of each web). In the rest of the webs,
minimum reinforcement is applied to control crack widths.
In section C, minimum reinforcement can be used in the entire web, down to the outer
fibre.
The reinforcement in section B extends over the full flange width. Codes, such as EN
1992-1-1 for buildings (cl. 5.3.2.1) and EN 1992-2 for bridges, give design rules to
calculate the effective width of a flange. All reinforcement applied within the effective
width can be assumed to be effective in resisting the external load.
It is decided to concentrate almost all the required reinforcement close to the webs. Bars
Ø 20 mm are applied near the webs (66 Ø 20 mm: As = 20734 mm2, which is almost
equal to the total amount required, namely As = 22,5·103 mm2). The remaining parts of
the upper flange are provided with minimum reinforcement.
required
12 Ø20 minimum
Ø12-150
4 * 12/16,0
4 * 12/16,0
4 * 12/16,0
cross-section B-B
The minimum reinforcement chosen, namely Ø12 - 150 mm, provides a minimum
reinforcement ratio of 0,38% in the webs and 0,75% in the flanges.
The previous calculations refer to ULS; SLS is checked later. This might imply that
additional reinforcement is required to meet crack width requirements, see step 5.
It is assumed that the crack width must be restricted to wmax = 0,15 mm. To control if this
requirement is met, the stress in the reinforcing steel is calculated.
(As+ξ1Ap)σs
εs (As+ξ1Ap)σs
34,93
0,72
16,01 16,01
2,6
20,56
1,18
1,08
16,97
Nc εc Ec Aflange εc
Fig. 11.22 Calculation of the forces and moments in section B at SLS loading (forces in
MN, moments in MNm)
The bond factor of the prestressing steel is calculated using EN 1992-1-1 table 6.2 and
clauses 7.3.2 and 6.8.2:
Øs Øs 20
1 0,5 0,38
Øp 1,6 Ap 1,6 12 150
The concrete compressive force is assumed to be at the centre of the flange. Two
equilibrium conditions hold:
M 0:
A A E 1, 65 P
s s 1 p s m, 1, 08 20,56
H 0:
A A E
s s 1 p s Pm, c Ac,flange Ec
It appears that the reinforcing steel is in tension. This result indicates that the cross-
section is cracked in SLS. The concrete compressive strain follows from horizontal force
equilibrium:
εc = 0,52 · 10-3
Figure 11.15: In case bar spacing s = 100 mm (fig. 11.21), σs,max = 220 N/mm2 >
124 N/mm2.
H 0:
A A E
s s 1 p s Pm, c Ac,flange Ec
The reinforcing steel is in tension; the cross-section is cracked in SLS. The mean
concrete compressive strain:
εc = 0,26 · 10-3
Tension zone height:
s 0, 73
t (h 0,10 0, 25) (1,9 0,10 0, 25) 1,14 m
s c 0, 73 0, 26
ae 0, 20
0,18 (fig. 11.16)
t 1,14
Figure 11.15: In case bar spacing s = 100 mm (average value) (fig. 11.21), σs,max =
220 N/mm2 > 146 N/mm2.
H 0:
A A E
s s 1 p s Pm, c Ac,flange Ec
ae 0, 20
0,18 (fig. 11.16)
t 1,14
Figure 11.15: In case bar spacing s = 150 mm (average value) (fig. 11.21), σs,max =
160 N/mm2 > 146 N/mm2.
The degree of prestressing can be calculated for the cross-sections considered. The
decompression bending moment follows from the bending moment from prestressing and
the axial prestressing force.
Section A:
P 17, 21 1, 08
M D m, Wb M pm, 13, 46 17,1 MNm
Ac 3, 61 1, 43
Section B:
P 16, 01 2, 29
M D m, Wt M pm, 16,97 28, 4 MNm
Ac 4, 45 0, 72
Section C:
P 14,80 1, 08
M D m, Wb M pm, 5,96 9,1 MNm
Ac 3, 61 1, 43
The degree of prestressing K is the ratio between the decompression moments and the
bending moment from permanent loads:
17,1
KA 0,85
18,81 1, 4
28, 4
KB 0, 76
34,93 2, 6
9,1
KC 0, 64
13, 02 1,1
Note that the bending moments from differential settlements are included in the bending
moments from permanent loads. This is in accordance with EN 1992-1-1 cl. 2.3.1.3
which states that differential settlements must be regarded as permanent loads, whereas
loads from temperature gradients are variable loads (EN 1992-1-1 cl. 2.3.1.2).
The degree of prestressing is relatively high (a minimum of 0,64). This indicates that the
prestressing force might be reduced. However, to meet the bending moment resistance
requirement, additional reinforcement As = 7200 mm2 has to be provided in the webs at
midspan (section A). When applying less prestressing, the amount of reinforcement
required strongly increases and it might be difficult to apply all the reinforcement at the
bottom of the two webs (there is no bottom flange at midspan). On the other hand, the
bending moment resistance calculation is conservative, since it is assumed that the
prestressing steel reaches a stress fp0,1k. A more detailed analysis might demonstrate that
the prestressing steel stress is between fp0,1k and fpk at ULS, which reduces the amount of
additional reinforcing steel required.
It becomes clear that the design is iterative, balancing the prestressing force on one hand
and on the other ULS (additional As) and SLS requirements (crack width from σs).
11.7 Literature
11.1 Bachmann, H.: “From full to partial prestressing”, Edition “Prestressed concrete
in Switzerland”, 9th FIP congress, 1982, pp. 11-18 (in German).
11.4 Kupfer, H., Scholz, U.: “Economy as a criterion for the choice of the degree of
prestressing”, Concrete Plant + Precast Technology, No. 5 / 1986, pp. 289-293.
11.5 Peter, J.: “Structures with partial prestressing”, Beton- und Stahlbetonbau 6/1986,
pp. 150-152 (in German).
11.6 Walther, R.: “Partial Prestressing”, Beton- und Stahlbetonbau 4/1975, pp. 79-82.
In a number of cases, the use of unbonded tendons results in increased economy, for
example slabs in buildings, containment structures (fig. 12.2), and shell structures.
The pre-cracked situation is present again outside the influencing area of the crack.
There, new cracks can occur once the concrete tensile stress σct reaches the tensile
strength of the concrete. This will result in a well-distributed crack pattern, with cracks at
regular, relatively small, distances.
In a beam with unbonded prestressing steel, a different mechanism occurs after the
formation of the first crack. However, in spite of the absence of bond, also here cracks
will appear at regular distances. Figure 12.3b shows the stress trajectories for the area
adjacent to the crack and gives the stress distribution over a section where the original
stress distribution is restored. The distance required to have the original stress distribution
depends on de St.-Venant’s disturbance length (here denoted as le) which is about equal
to the height of the concrete cross-section. The elongation of the prestressing steel is very
large, because the prestressing steel is only fixed at the anchorages at both ends of the
beam. This implies that, in case of one crack, the elongation of the steel (increase of steel
strain multiplied with the tendon length between the anchors) concentrates in one crack.
Therefore, the crack widens considerably and substantially reduces the depth of the
uncracked compressive area. Furthermore, the compressive zone is highly curved and
tends to separate itself from the low-stressed concrete parts in between the cracks, see fig.
12.4.
Fig. 12.4 The formation of transverse cracks at the tip of bending cracks
This results in a crack pattern characteristic for beams with unbonded tendons, namely
the so-called “fork-shaped” cracks, see fig. 12.5. This figure also shows the failure mode
typical for beams with unbonded tendons: the concrete fails in compression before the
steel has reached its tensile strength or its 0,1%-strain limit.
dh
M with bonding
without bonding
Fig. 12.6 Qualitative comparison between two prestressed beams: one with bonded and
one with unbonded tendons
qG + 0,4 qQ
d.l.+0.4 p.l.
Pm
Fp Fp
f Pm
Fig. 12.7 Balancing part of the load by prestressing (in this example: the permanent
load plus 40% of the variable load)
For slabs, as a simplifying assumption for the calculation, it is assumed that the total
deformation concentrates at one crack, see fig. 12.8. The slab is modelled as two rigid
bodies that rotate about their intersection point.
w
Fig. 12.8 Simplified calculation of the increase of the stress in the prestressing steel in
a slab
The vertical displacement at midspan is denoted δ. From the geometry it follows that:
1
2 l
w 2 z
If the internal lever arm z (distance from the prestressing steel force to the resulting
concrete compressive force) is estimated to be 0,75 h, then
w 1,5 h
The total elongation l of the prestressing steel is equal to the crack width w, so that:
l 1,5 h 1,5 h3 h
1
2 l l
It was found experimentally that the deflection at failure is usually more than l / 50.
When using this conservative value
l
(12.1)
50
it is found that the elongation l of the tendon is at least
3h h
l (12.2)
50 17
This is a conservative estimation for slabs. For a more accurate calculation, reference is
made to [12.3] and [12.4].
Figure 12.9 shows a continuous slab over a number of supports, subjected to a special
load configuration. In this case it is assumed that the entire slab is subjected to a quasi-
permanent load qg + qq and that for one span the variable load is gradually increased.
Failure is assumed to occur at the ultimate load value qu at the highest loaded span.
Cracks are assumed to occur at the two supports of this span, as well as at its mid
position.
If it is assumed that the deflection in the highest loaded span is = l / 50, then the total
elongation of the prestressing tendon is 2 h / 17. This is two times the result from eq.
(12.2) since a crack develops at each of the two intermediate supports on each side of the
span. Both these two cracks have a width w / 2. This follows from compatibility of
rotating rigid bodies, assuming that the crack width at midspan is w. Moreover, it is
assumed that the internal lever arm z = 0,75 h at the supports too and that the adjacent
spans do not deform.
As a result of the elongation, the increase of the force in the prestressing tendon is:
2
h
P 17 Ep Ap
L
where L is the total length of the tendons between the anchorages.
Since the prestressing force is constant over the full slab length (note: unbonded
tendons!), the upward equivalent (balancing) prestressing load for each span is:
8 ( P P) f
qp (12.3)
l2
It now looks as if there is no difference between spans with and without cracks. However,
the tendon profile is forced to follow the slabs’ discrete rotations that occur at the cracks.
This causes kinks in the tendon profile. Their impact is discussed in detail in the chapter
“External prestressing”.
qplu.qp+vl.qvl
g qg + q qq
pl.qpl+vl..qvl
Pm + ΔPm
Fp + Fp
f
f
FpP).f
88.(FPp+
f
m m
F2
8.(Fp+Fp).(f+f) l
F
The minimum radius of curvature R of the prestressing tendons should, according to the
Dutch code NEN 6720, satisfy the following conditions:
R 20 ØT for strands in a smooth duct (unbonded)
R 40 ØT for strands in a ribbed duct (bonded)
where ØT is the diameter of the tendon (including the duct).
In practice, the radius of curvature at an intermediate support is about 1,5 m for unbonded
tendons and about 2,5 m for bonded tendons.
This conclusion is basically correct for slabs not provided with additional reinforcing
steel. However, additional (bonded) reinforcing steel is often provided. This might come,
for instance, from crack width reduction (SLS) or from bending moment resistance
(ULS) requirements. Thanks to the bonding effect of the reinforcing steel, inclined
cracking occurs.
Research [12.5, 12.9] demonstrated that shear reinforcement can also be efficient for
another reason, even if no additional reinforcing steel is provided. This is illustrated in
fig. 12.11a, showing the failure mode of an I-beam prestressed with unbonded tendons.
The beam has no additional reinforcement in the tensile zone and has no shear
reinforcement. The first inclined crack that appears in the web immediately results in
failure of the beam. This inclined crack is no bending tensile crack, starting from the
bottom of the beam, but is caused by splitting of the compression strut between the point
where the load is applied and the support (fig. 12.11b). Stirrups can be effective in
controlling crack formation in the strut.
a b
Figure 12.12 shows the behaviour of a similar beam provided with shear reinforcement.
Figure 12.12a shows the reinforcement. In fig. 12.12b it is shown that a bending crack
originates at the bottom of the beam, just at the position of the load. Well-distributed
inclined cracks originate in the web. After the formation of the first inclined crack, the
load could still be considerably increased until the compression strut failed (web crushing
failure, fig. 12.12c).
Fig. 12.12 Test on an I-shaped beam, prestressed with unbonded tendons and shear
reinforcement.
a. Reinforcement
b. Crack pattern
c. Failure mode
The behaviour displayed in fig. 12.12 shows a large similarity with the behaviour of a
beam with (bonded) reinforcing steel, as illustrated by fig. 12.13. Figure 12.13a shows a
truss model that simulates the behaviour of a beam reinforced with passive steel. Figure
12.13b shows a model to simulate the behaviour of a beam prestressed with unbonded
tendons. The only difference between both models is the force in the longitudinal tensile
tie at the bottom of the truss. In case (b), this force, that only slightly increases during
loading, acts as a concentrated force at the beam ends. In case (a), on the contrary, there
is a tensile force inside the beam. In both cases, the system is in equilibrium; in case (b)
with a compressive force at the bottom of the beam, and in case (a) with a tensile force
provided by the reinforcing steel. This implies that the shear resistance of structural
members with unbonded tendons can be calculated using the same procedure as for
members with bonded reinforcement.
- - - -
a - + - + - - + - + -
+- + + + +-
- - - -
b - + - + - - + - + -
- - - -
Fig. 12.13 Truss models for a beam with passive reinforcement (a) and for a beam
prestressed by unbonded tendons (b)
Prestressed slabs are predominantly applied for structures like office buildings, parking
houses, schools, hospitals and warehouses. These structures, with spans from 7 to 10 m
and live loads up to 5 kN/m2, mostly have flat slabs provided with concentrated tendons
in column strips in one direction. In the case of larger spans and/or higher loads, the
prestressed slabs are provided with drop panels, have columns strips with concentrated
tendons in two directions or are designed as waffle slabs.
Compared with other structural solutions, prestressed flat slabs (both with bonded and
unbonded tendons) have the following advantages:
Larger spans and smaller slab thickness, see fig. 12.14.
The selfweight of the slab structure is reduced, which has a favourable effect on the
dimensions of the structure and its foundation.
Thanks to the smaller height of a storey, the total height of the building can be
reduced, or the number of storeys in a building with a fixed height can be increased.
This also holds for underground structures.
d c b
300
250 a
floor thickness [mm]
a
b
200
c
d
150
100
5 6 7 8 9 10 11 12 13 14 15
span [m]
Figure 12.15 shows four options for arranging the prestressing tendons.
In tendon configuration (a) the same number of tendons is provided in each direction.
The distance between the tendons in the span is relatively large and the tendons are
predominantly concentrated in the column strips since this is advantageous for the
transmission of the balanced load (exerted by the tendons) to the column. A simple
equilibrium model of the punching cone shows that the part of the load directly
transferred to the column, is 2 Pm,∞ sinα per tendon, see fig. 12.16. The smaller the radius
of curvature, the more the tendon contributes to the punching shear resistance.
In tendon configuration (b), see fig. 12.15, all prestressing tendons are concentrated in the
column strip. Between the columns strips, the slab is reinforced with (passive) reinforcing
steel.
In the configurations shown in fig. 12.15 (c) and (d), in one direction column strips are
provided with concentrated tendons. In the other direction, the slab is prestressed by
distributed tendons (c) or is only reinforced (d).
a b
prestressing tendons
prestressing cables
reinforcing steel in
in the
in the field
span steel reinforcement
the span
in the field
prestressing
prestressingcables
tendons
in the collumn strip
in the column strip
c d
prestressing tendons
prestressing cables
in the span
in the field
Fig. 12.15 Various options to arrange unbonded prestressing tendons in a flat slab
[12.13]
Pm sin FpP.msin
sin
Fp.sin
punch cone
punching shear cone
Fig. 12.16 Direct transfer of the balanced load from the prestressing tendons to the
column
A prestressed flat slab has to be checked for the ultimate limit states (ULS). One of them
is the bending moment resistance check. The bending moments in the slab caused by
loading can be determined on the basis of the theory of elasticity. EN 1992-1-1 does not
provide detailed information about the bending moments in flat slabs. EN 1992-1-1
annex I presents some simplified rules to distribute the negative (hogging) and positive
(sagging) moments over the column and middle strips (table I.1).
NEN 6720 cl. 7.5.3 presents more detailed information. Bending moments were
calculated using the theory of elasticity. The locally very high negative moments at the
columns were distributed over the column strip. The reinforcement moments (bending
moments plus torsion moments) were calculated and transformed into coefficients given
in tables (NEN 6720 tables 19 - 26).
The tendon configurations from fig. 12.15 do not in all cases perfectly match with the
distributions of the moments that follow from the theory of linear elasticity. In particular
the arrangements (c) and (d) give deviating results. This might imply that there is some
redistribution of forces when, under increasing load, the slab cracks.
Especially the areas with high concentrated moments (near the columns), are sensitive to
this bending moment redistribution. Therefore, NEN 6720 cl. 9.9.2.2.c specifies a
minimum reinforcement required in those areas. This reinforcement should be provided
in both main load bearing directions and extended up to a distance of 0,25 lx and 0,25 ly
in x and y direction, respectively, from the centre of the column, see fig. 12.17. The
minimum reinforcement ratio for loading in bending has to be applied, see Chapter 7.
x
0.25 x 0.25 x 0.25 x
0.25 y
0.5 x
0.25 y
y
0.25 y
ar 0.25 x
ar 0.25 y 0.5 y
x
Fig. 12.17 Minimum reinforcement in the column area in a flat slab prestressed with
unbonded tendons
The parking house has only one storey. Expansion joints subdivide it into three units. In
the design, a top layer of 0,4 m soil on top of the concrete roof has to be accounted for.
Furthermore, it is assumed that vehicles class 30 (total vehicle weight 300 kN) can pass
and load the roof.
By prestressing the slab instead of only reinforcing it, the slab thickness is reduced from
500 to 350 mm, whereas, simultaneously, the span increases from 5,0 to 8,75 m.
In the project, unbonded monostrands Ap = 139 mm2, Y1860, are applied. The ULS
design value for the ULS (design load) of such a tendon is 139 · 1860 / 1,1 = 235 · 103 N.
The code used in the design allows a force directly after prestressing of 188 kN, that can
be temporarily increased to 204 kN. For the calculation of the (small) prestressing losses
by friction μ = 0,06 is used, in combination with k = 0,009 rad/m for the Wobble effect.
Thanks to these low values, it is possible to even prestress the longest tendons, which
pass five spans, from one side, without too high losses.
NEN 6720 cl. 7.5.3 was used to determine the moments in the column strip and the
middle strip from the selfweight and the soil layer. The tendons are arranged cross-like,
with a concentration of tendons in the column strip and an increased tendon spacing in
the middle strip (see fig. 12.15a). The prestressing system can effectively balance the
load.
For the roof of the parking house, the balanced (equivalent) load qp is equal to the
selfweight of the roof qg,1 and the soil layer qg,2. The traffic load is not yet considered
because the structure is designed in partially prestressed concrete. This implies that the
requirements from ULS (failure safety) and from crack width limitation (under maximum
load in SLS) will be met by adding reinforcing steel if required.
At a volumetric weight of 25 kN/m3 for the reinforced concrete and 19 kN/m3 for the soil,
the total load to balance is:
The prestressing force required to provide this balancing load is calculated using the
relation:
qp l 2
Pm,
8f
At first, the slab strip between axes 14 and 16 (see fig. 12.18) is considered. The main
span in this direction is 8,75 m. The distance from the tendon to the outer concrete fibre
is assumed to be 0,04 m, both at midspan and support.
The drape of the tendon profile then is f = 0,35 – 0,04 – 0,04 = 0,27 m (neglecting the
downward curvatures at the supports).
581
Pm,0 726 kN/m
0,8
726
This corresponds to a theoretical number of tendons of 3,9/m1 .
188
For a width of 7,50 m (see fig. 12.18 on line F), the number of tendons is n = 7,50 · 3,9 =
29,30 30.
About half of these tendons are arranged in the column strips, whereas the other half is
uniformly distributed over the middle strip (between the column strips).
The mean compressive stress in the concrete σcm,p is:
30 0,8 188 103
cm,p 1, 72 N/mm 2
7,5 1000 350
This stress is relatively low (much lower than, for instance, in beams). This demonstrates
that the most important component of the prestressing is the balancing part.
Friction, shrinkage and relaxation losses, and losses by wedge slip, can be determined in
the usual way. Depending on the degree of prestressing, they will amount up to,
approximately, 20%. The elastic shortening of the slab by the axial compressive force is
small.
The resistance check in ULS should be carried out following the method presented in
section 12.4. Furthermore, crack width control should be carried out according to the
method presented in Chapter 9. This is not further presented here.
Special attention should be paid to the behaviour of the slab in the vicinity of the
columns. As pointed out already, the punching shear resistance is considerably increased
by the component Vp = Pm,∞ sinα. It should be emphasized that this component is
automatically taken into account when using the equivalent prestressing load (balancing)
method, in which the prestressing loads are directly transferred to the column by the local
downward pressure of the curved tendons. So, additionally reducing the punching shear
force by Vp = Pm,∞ sinα would be erroneous, because then the effect of prestressing
would be taken into account twice! The effect of prestressing can be optimized by
positioning the inflexion point of the tendon (where the upward curvature changes into a
downward curvature and vice versa) in the punching cone. In that case the maximum
effect of load balancing is achieved. In the case considered, no shear reinforcement was
required because of this effect.
13 14 16 17
7.50 7.50 4.875
F
4.705
E
a a
8.75
D
section a-a
4
10
= =
35
4
2e bundle 3e bundle 4e bundle
Fig. 12.18 Arrangement of the prestressing tendons in the design example (dimensions
m (top) and cm (bottom))
For easier handling during positioning, the tendons can be bundled at the construction
site. According to clause 9.4.2c of NEN 6720 a bundle can consist of a maximum of two
monostrands. In other countries larger bundles are allowed. For instance in Germany, it is
allowed to bundle up to four monostrands in the column strip (fig. 12.19) and up to two
in the middle strip. In the anchorage region, the distance between the strands has to be
increased to 120 -140 mm, to provide space for the anchors, see fig. 12.18.
Prestressing with unbonded tendons also offers advantages for the design of containment
structures and shell structures. In those cases failure safety does not play a dominant role,
but crack width control is mostly governing. The elastic ducts fully function as corrosion
protection, if the concrete is cracked whereby the tendon is exposed to aggressive
substances. Particularly with regard to this aspect, a number of special applications are
possible, such as prestressing of bridge decks and containers with aggressive fluids.
For an extended example of the design and detailing of an office building slab,
prestressed with unbonded tendons, reference is made to [12.15].
12.8 Literature
12.1 Gerber, C., Özgen, W.: “Flachdecke mit teilweiser Vorspannung ohne Verbund”,
Beton- und Stahlbetonbau, Juni 1980, Heft 6., Seiten 129-132.
12.2 Grasser, A.: “Bemessung von Beton- und Stahlbetonbalken bei Biegung mit
Längskraft, Schub und Torsion”, Deutscher Ausschuss für Stahlbeton, Heft 240,
1979.
12.3 Ivanyi, G., Buschmeyer, W.P.: “Biegerißbildung bei Plattentragwerken mit
Vorspannung ohne Verbund”, Beton- und Stahlbetonbau 9/1981, S. 215-220.
12.4 Ivanyi, G., Buschmeyer, W.P.: “Additional strains in unbonded tendons during
loading”, Magazine of Concrete Research, Vol. 37, No. 37, March 1985.
12.5 Ivanyi, G., Samol, J.: “Versuche zur Schubtragfähigkeit an Balkentragwerken mit
Vorspannung ohne Verbund”, Forschungsbericht No. 31, Universität Essen, Juli
1985.
12.6 Ivanyi, G., Buschmeyer, W.P., Müller, R.A.: “Entwurf von vorgespannten
Flachdecken”, Beton- und Stahlbetonbau 4/1987, S. 95-105.
12.7 Ivanyi, G., Buschmeyer, W.P.: “ Kontrollierte Biegerißbildung zur Sicherung der
Verformungsfähigkeit bei teilweiser Vorspannung ohne Verbund”, Der
Bauingenieur 62, 1987, S. 339-343.
12.8 Kordina, K., Hegger, J., Teutsch, M.: “Anwendung der Vorspannung ohne
Verbund”, Deutscher Ausschuß für Stahlbeton, Heft 335, 1984.
12.9 Kordina, K., Hegger, J.: “ Zür Ermittlung der Biegebruch-Tragfähigkeit bei
Vorspannung ohne Verbund”, Beton- und Stahlbetonbau 4/1987, S. 85-90.
12.10 Marti, P., Ritz, P., Thürlimann: “Prestresssed Concrete Flat Slabs”, Institut für
Baustatik und Konstruktion ETH Zürich, Bericht No. 68, Feb. 1977.
12.11 Matt, P.: “Vorspannung ohne Verbund: Beispiele und Möglichkeiten der
Anwendung”, Beton- und Stahlbetonbau 9/1981, S. 212-215.
12.12 Schütt, K.: “Vorspannung ohne Verbund: Möglichkeiten und Beispiele
ausgeführter Bauwerke”, Beton- und Stahlbetonbau 6/1968, pp. 153-154.
12.13 Wölfel, E.: “Flachdecken mit Vorspannung ohne Verbund”, Der Bauingenieur 55,
1980, S. 185-195.
12.14 Leeuwen, J., Tukker, T., Veenstra, P.: “Voorspanning zonder aanhechting”, CUR
Rapport 95, December 1979.
12.15 Bouquet, G.Chr., Groeneveld, J., Keusters, A.C.A.M., Pauw, J.H., Veen, C. van,
Zielinski, A.J.: “Toepassingen van VZA en VMA bij vloeren in de
utiliteitsbouw”, Stuvo-rapport 95, mei 1992.
The interest in external prestressing has, however, grown again over the last decades. For this
increased interest a number of reasons can be mentioned [13.1], [13.2]:
The necessary repair of bridges with corroded prestressing tendons.
The need for methods to strengthen structures because of increased traffic loads.
To simplify execution, by avoiding a complicated layout of the tendons inside the
concrete. This offers advantages with regard to casting (fig. 13.1), prestressing (no or
small friction losses) and the injection of the tendon ducts (no leakage). Furthermore,
unintended curvature pressures are avoided (fig. 13.2).
The need to be able to repair (bridge-) structures, without temporarily closing them.
To reduce the web width, by getting rid of the tendon ducts (fig. 13.3).
To be able to control the durability of the structure and to eventually replace corroded
tendons.
The possibility to apply additional prestressing in the case of unexpected deformations.
Fig. 13.3 Reduced strength in the construction phase from empty ducts in the web
Figure 13.4 shows an application of external prestressing. It concerns the Long Key Bridge in
Florida (USA). After placing the box-girder type of segments on a temporary steel truss, they
were prestressed together by external tendons. The tendons were guided through concrete
anchor blocks (saddles) at the bottom of the cross-section. At those saddles, the prestressing
forces were transmitted to the structure.
35.96
Fig. 13.4 Long Key Bridge: schematic representation of the construction sequence and the
tendon profile [13.3]
Another possibility is to apply transverse beams (fig. 13.5a) or stiffeners (fig. 13.5b) for
changing the direction of the tendons.
Especially in France (70 bridges) and the USA (30 bridges), experience is acquired with
externally prestressed bridges. Reference is made to publications of Virlogeux [13.4]-[13.8]
and Müller [13.9]. Nelissen [13.13] and Bruggeling [13.14] prepared State-of-the-Art papers
in the Dutch journal “Cement”. Vermeulen [13.22] demonstrated that external prestressing
can be of interest in Dutch conditions.
Principally, for external prestressing the same prestressing steel can be used as for “normal”
prestressing. An important point is, however, the way in which external prestressing steel is
protected against corrosion. There are various possibilities.
Galvanising:
In this case the corrosion resistance depends on the type of galvanising and the thickness of
the coating. Galvanised prestressing steel has been applied in France several times. However,
there is some doubt about the long-term corrosion resistance. Furthermore, the coating can be
damaged during mounting and replacing of the tendons.
Epoxycoating:
The technology of applying melted polymers on the steel has been developed in the USA,
especially for reinforcing steel. It is questionable whether it is a good solution for prestressing
steel. At the anchorage the coating is locally interrupted, so that problems might occur here.
Special measures are necessary as well, to avoid damage during transport and mounting.
Protective ducts:
Suitable materials are steel and plastic (High Density PolyEthylene, HDPE). In order to
achieve a complete protection system, good solutions are needed for coupling, the anchorage
region and the saddles (where the deviations of the tendons occur). Injection of the ducts with
a cement grout is a good and economic solution. If the possibility of second prestressing (for
compensating unexpected deformations) should be kept open, no cement grout can be used.
Other possibilities are grease, paraffin, tar-epoxy or other products with plastic properties.
These products are not easy to inject; sometimes heating up to 100 ºC is required.
Furthermore, special measures are necessary to avoid leakage.
In this category monostrands, produced in factories, offer many advantages. Those strands are
surrounded by a plastic duct, which is filled with grease (fig. 2.20). In this way, the
prestressing steel is effectively protected from corrosion during transport, storage at the
building site and assembling. Monostrands can be applied as single units or in bundles. In the
last case they are mostly placed in an additional coating of steel or plastic, see fig. 13.6 (left).
The space between the strands is filled with grease or with cement grout. In the case of filling
with grease, the tendons, consisting of many monostrands, can be readily delivered at the
building site. Often internal spacers are used, guaranteeing the distance between the
monostrands, also at the saddles. The individual strands can, if necessary, be stressed or
released one by one.
“Prestressing bands” are composed of a limited number of monostrands (fig. 13.6, right).
They need only a small volume of grease, also when they are composed to large units, see fig.
13.7. The prestressing bands can be bundled by piling up.
At deviation points, a solution is necessary which enables tendon deviation and the
introduction of local prestressing loads into the structure without damage. This pertains to
both vertical and horizontal deviations. Horizontal deviations can occur in bridges with a
curved axis. The transition of the curved part of the tendons into the straight part should be
inside the area of the saddle, since otherwise kinks occur in the tendons. This concerns again
both the vertical and the horizontal curvature, see fig. 13.8. For the detailing of a saddle there
are various possibilities, like the ones shown in fig. 13.9.
section B-B
A A
section A-A
undesirable deviation
B B
PE
a ST
steel tube
PE tube
ST = PE
socket
b steel tube
PE tube
ST < PE
socket
c steel tube
PE tube
ST > PE
Fig. 13.9 Some solutions for the junction of the external tendons to the saddle [13.15]
a & b: Discontinuous PE tubes connected to a steel tube
c: Continuous PE tube through a (larger diameter) steel tube
In most saddles a pre-deformed steel tube is applied, cast into the concrete or, in the case of a
steel saddle, connected by stiffening plates. The connection between the free part of the
tendon and the saddles should be detailed carefully not to damage the prestressing steel by
kinking. If the exchangeability of the tendons is a design criterion, this should be regarded in
the design.
Relative displacements between the prestressing steel and the saddle should be allowed to
occur. If the prestressing tendon and the surrounding tube have been rigidly connected, for
instance by a mortar, another solution is necessary. In this case the relative displacement can
be facilitated by a neoprene intermediate layer that allows a displacement of the tendon and
tube relative to the saddle, see fig. 13.10.
For the design of the saddles, the minimum radius of curvature has to be taken into account to
avoid damage to the prestressing steel or duct. In [13.16] allowable radii of curvature are
given as a function of the tendon dimensions. These radii of curvature are between 2,5 m and
5,0 m.
mortar
morter
elastomeric bearing
elastomer
Fig. 13.10 Solution for the relative displacements between the saddle and a rigidly injected
prestressing tendon
In analogy with conventional prestressing systems, friction losses are calculated with eq.
(4.23). The Wobble-effect can be ignored, since the tendons have a straight alignment
between the saddles. On the basis of experimental results and experiences at the building site,
the following values apply for the friction coefficient :
Figure 13.11 shows a beam on two supports with two deviation points. The prestressing
forces are regarded as being loads on the beam. At the deviation points the loads are applied
on the structure in the direction of the bisector of the angle of deviation (dashed arrow). If the
friction in the saddle is also taken into account, the resulting force indicated by the arrow
rotates over a certain angle.
The prestressing forces can be split into horizontal and vertical components. The calculation
can then be carried out in the usual way.
By a right choice of the tendon profile, a prestressing load can be generated, which balances
the loads (dead weight and part of variable load) on the structure in the optimum way.
Sometimes more deviation points are created to have the optimum balancing effect, see fig.
13.12.
V2v V1v
Zv
Dv
e H2v
H1v
Normal force
A B - friction-
In finding the optimum design, there is a strong dependence on the construction method. If it
is possible to provide reinforcing steel complementary to the external tendons, a solution in
partially prestressed concrete is possible. Then, the dead weight and a part of the variable load
are balanced. At relatively high loads cracks occur, which are well distributed and small, as a
result of the action of the reinforcing steel. By choosing a favourable ratio between the
amounts of prestressing steel and reinforcing steel, an economic solution can be achieved.
If, however, the structure is built in segments, the situation is totally different. The joints are
mostly profiled (fig. 13.13) to transmit the shear forces, and are bonded together by epoxy
glue. In this case no reinforcing steel crosses the joint, so that the crack width control is a
critical aspect. Since there is not sufficient knowledge about the long term behaviour of the
epoxy, it is mostly required to design the structure in such a way, that in SLS the joints are in
permanent compression. In such a case due notice should be taken of the effects of imposed
deformation (temperature gradients, differential settlement).
With regard to the behaviour in ULS, the best approach is to use a kinematic failure model. In
fig. 13.14 this is shown for a simple case. The elongation of the prestressing tendon is a
function of the deflection. There is also a relation between the elongation of the tendon and
the load. In section 13.4 an analysis is given for a specific case.
5000
300
3x500 450 400 4x500
100
500
1000 1000
35
25
A
50
25
50
25
150
50
section A-A
25
A A
f
w = 4f .h w = f .2
L L
Fig. 13.14 Kinematic model for an externally prestressed beam in the failure stage (two
anchors and two deviation points)
For a numerical analysis of a segmental bridge with external prestressing, reference is made to
Huang [13.19]. Figure 13.15 shows the calculated deformations in the failure stage. The
asymmetric behaviour is caused by torsional moments, corresponding to the eccentric loading
case considered. The analysis shows a distinct change of behaviour when the joints open, see
fig. 13.16. The calculation shows as well, that the transmission of shear forces close to the
support can become critical.
Fig. 13.16 Principal stresses in the web of a box girder bridge at the failure load [13.19]
Because of the small bottom flange, locally a combination of very high compressive stresses
and shear stresses occurs. Furthermore, fig. 13.16 shows that the transmission of forces in the
segments between the (opening) joints needs appropriate attention. Moreover, fig. 13.17
shows truss models for a segment with (fig. 13.17b) and without a tendon deviation point.
compression bars
Survey
tensile bars
a section bb cc
p p p p p p p p p p
P2
P1 P2 P1
Zv Zv
flow of
Flow of forces
forcesinin flow of
Flow of forces
forces in
in
segment
transversewithdiafragm
open segment withdiafragm
transverse open
joints
in theand deviation
span with joints andspan
in the no deviation
without
forceforce
deviation force
deviation force
T1 T2 T1 T2
p p p p p p
D1 D2 D1 D2
Zv
Fig. 13.17 Truss models for a segment with and without tendon deviation point [13.19].
Expansion joints have been provided at any 8 spans (distance 288 m). For the assembling of
the segments a moveable steel truss was used, see fig. 13.19. At first, all segments of a span
were put in position. In order to have a levelled bridge deck, the steel truss had a camber.
Between the pier segment and the first segment in a span, a joint having a width of 150 mm
was provided, filled with concrete at the site. Subsequently, the span was prestressed in one
operation. Per span always six tendons were used. These were anchored above the piers,
where the tendons of the adjacent spans overlap, see fig. 13.20.
35.96
In the saddles, the tendons were led through steel ducts. In between the saddles, they are
placed in plastic ducts. At the junction between the two types of ducts, neoprene sleeves take
care of the tightness and the required elasticity, see fig. 13.21. After prestressing, the ducts
were injected with a grout. In this way the upper structure had its full bearing resistance
directly after prestressing. By using this assembling system, two to three spans per week (72
to 108 m) could be finalized. In some periods the production even increased to five spans per
week (180 m).
A particular aspect was that no structural topping was applied on the deck: the traffic moved
directly on the concrete of the segments. Another particularity was, that dry joints were
applied between the concrete segments (concrete to concrete, without intermediate layer).
This requires a high production accuracy. Figure 13.22 shows a cross-section at a joint.
Polyethylene
duct
2500 mm
P P
c.a Ac = 4,07 m2
Ic = 3,51 m4
The selfweight of the upper structure is 4,07 · 25 = 102 kN/m = 0,102 MN/m.
The variable (traffic) load is 500 kg/m2, which is 8 · 5 = 40 kN/m = 0,040 MN/m.
In situation 1, the full load is applied over the full length of the bridge. For that situation, the
resulting bending moments are:
When there is no variable load on the bridge, the bending moments are:
As a simplification, the cases “fully loaded” and “fully unloaded” are regarded as being
governing for the design.
The loads by external prestressing are determined using fig. 13.24. Only the vertical force
components FV are presented. The horizontal force components at the deviation points
(P·(1-cos)) are not indicated.
P P
FV FV
FV =Psin
P
F =2Psin(/2)
-FV a2 / l
FV a (l-a) / l
If the deviation points of the tendon profile are applied at a = l / 3, then the bending moments
generated by prestressing are:
Support: Ms = 2/9 FV l
Midspan: Mm = -1/9 FV l
FV = P sin = 0,144 P
Note: = 0,144 rad is a small angle. This implies that sin = can be used.
Now, the value for FV is determined to balance the dead weight and 50% of the variable load.
At the support, a bending moment of 0,5 · (-12,9 + -9,3) = -11,1 MNm has to be balanced by
prestressing. The force required follows from:
2
/9 FV l = 11,1 MNm
or:
2
/9 FV l = 2/9 · 0,144 P · 33 = 11,1 MNm
so that:
P = 10,5 MN
With pm,∞ = 1000 N/mm2 this would require a cross-sectional area of the prestressing steel
Ap = 10,5 · 106 / 1000 = 10500 mm2, or four tendons with Ap 2600 mm2 each.
However, the mean compressive stress in the concrete by the axial prestressing force has not
been taken into account yet:
The SLS stress distribution in the governing cross-sections can now easily be calculated:
M 1,8
ct cm 2, 6 2,1 MN/m 2
I 3,51
yt 0,91
M 1,8
cb cm 2, 6 3, 4 MN/m 2
I 3,51
yb 1,59
In a fully loaded situation in SLS, the remaining moment at midspan M = 6,4 – 11,1 / 2 =
0,85 MNm.
M 0,85
ct cm 2, 6 2,8 MN/m 2
I 3,51
yt 0,91
M 0,85
cb cm 2, 6 2, 2 MN/m 2
I 3,51
yb 1,59
Now the situation in case no variable load is applied, is dealt with. Only the dead weight and
the prestressing are active.
M 1,8
ct cm 2, 6 3,1 MN/m 2
I 3,51
yt 0,91
M 1,8
cb cm 2, 6 1,8 MN/m 2
I 3,51
yb 1,59
M 0,95
ct cm 2, 6 2, 4 MN/m 2
I 3,51
yt 0,91
M 0,95
cb cm 2, 6 3, 0 MN/m 2
I 3,51
yb 1,59
All values for both the fully loaded and the unloaded situation are given in fig. 13.25. It is
seen that the joints are fully compressed under all loading combinations.
-12,9 -12,9
-9,3 -9,3 -2,1 -3,1 -2,4 -2,8
4,6
G
G+Q
fully loaded Q=0
6,4
G+Q G
support span
from prestressing
11,1 11,1
Fig. 13.25 Bending moments [MNm] and stresses [N/mm2] at support and midspan
To open the joint, the support moment has to introduce an additional 2,1 N/mm2 tensile stress
at the top. The increase of the bending moment required is:
I 3,51
M ct 2,1 8,1 MNm
yt 0,91
relative to the fully loaded situation. The corresponding increase of the uniformly distributed
load q follows from:
For the span the stress increase is 2,2 N/mm2 to have joint opening. This requires:
I 3,51
M cb 2, 2 4,9 MNm
yb 1,59
It turns out that joint opening occurs at an additional load of 0,089 MN/m = 89 kN/m (the
support cross-section is governing).
The total load at joint opening at the support then is (selfweight + variable load + additional
load):
If the load is further increased, the midspan cross-section joint opens too and finally, the
mechanism shown in fig. 13.26 occurs.
l/6
Fig. 13.26 Failure mechanism with opening joints at supports and at midspan
It is now assumed that the parts between the opening joints are rigid, and that the force in the
prestressing tendons starts to increase only at joint opening. So, when the kinematic model
starts to work, the stress in the prestressing steel is still pm,∞ = 1000 N/mm2, with a
corresponding prestressing force P = 10,5 MN.
At the ends of the structure, the prestressing force P is applied centrically. At a deviation
point, a vertical force FV = P sin = 0,144 P = 1,51 MN is exerted. The situation directly
after opening of the joints is shown in fig. 13.27, where the rotation of the bridge part is
exaggerated since actually there is almost no vertical displacement in this stage. At increasing
deformation, the compressive force in the concrete gradually moves downwards at the support
and upwards at midspan.
It is assumed that in the first stage after joint opening the deflection of the bridge is that small,
that the angle of the kink does not change. Moreover, it is assumed that joint opening results
in a shift of the concrete compressive forces to the centre of the flanges. The vertical positions
of the two horizontal forces at the right hand side of fig. 13.27 (10,5 MN tendon force and
10,5 MN top flange concrete compressive force) are not influenced by the small deflection.
The small deflection also makes that tendon elongation is small. It is, therefore, assumed that
the prestressing force does not change.
10,5
10,5 10,5
1,51
1,51
1
/2ql
10,5
Fig. 13.27 Equilibrium of a rigid part during joint opening (dimensions MN and m;
displacements not to scale)
The corresponding external load is found on the basis of the equilibrium of moments, for
instance about point S. As stated before, it is assumed that the compressive forces in the
flanges are centric (distance to outer fibre is half the flange height). The forces acting on the
concrete rigid body are:
A 10,5 MN compressive force at the centre of both top and bottom flange;
a 1,51 MN vertical force at both kinks (at the deviator at the support and the deviator at
11 m from the support);
a uniformly distributed load q;
a vertical reaction force at the support (1/2 q l).
2 2
so:
q = 0,299 MN/m = 299 kN/m
Note: In the previous calculation only the dominant bending moment contributions are
accounted for.
The stress in the ultimate fibre of a joint (at the support) is zero for q = 231 kN/m. The load at
which the joints are open appears to be q = 299 kN/m. The difference between both can be
explained as follows:
Also at midspan the situation of zero stress in the joint must first be reached.
The compressive forces in the concrete gradually shift towards the outer fibres (increased
internal lever arm).
At further increase of the load, the deflection increases rapidly and the tendons stretch. The
design value that follows from the characteristic tensile strength of the prestressing steel is
1860 / 1,1 = 1691 N/mm2. This stress is reached at a relatively high strain. It is therefore
assumed that the stress in ULS is somewhat lower: the stress in the tendons is assumed to
increase to the design value, which is related to fp0,1k and is about 90% of the design value
derived from the characteristic strength. The material factor γs has to be taken into account,
which implies that the stress is assumed to increase to fp0,1k / γs = 0,90 · 1860 / 1,1 =
1522 N/mm2. Up to failure occurs, the tendon force can therefore increase with a value:
Note that using a stress of 1522 N/mm2 implies that the prestressing steel is assumed to be
still in the linear elastic stage (just at the kink in the stress-strain relationship from EN 1992-
1-1 fig. 3.10).
In the kinematic model shown in fig. 13.26, with rigid moving parts, the tendon elongation is:
l = 2 w1 + w2 = 2 z1 + z2 2 (13.1)
where
= / ( 1 /2 l ) (13.2)
and
It is now assumed that the failure mechanism from fig. 13.26 is active in all spans. This
implies that the increase of the strain of a tendon follows from the elongation over a span and
the span length.
Note: In case only one span would show a failure mechanism, the elongation of the tendon
has to be distributed over a much larger distance. The result is a much smaller strain increase.
42
l 15, 6
l 3 l
The increase of strain in the prestressing steel (which is assumed to be in the linear elastic
stage):
l E
Es Es 2s (15, 6 1,33 2 ) (13.4)
l l
To reach p = 522 N/mm2 = 522 MN/m2 (Ep = 195 · 103 MN/m2 and l = 33 m), it is found
that = 0,190 m is required.
Note that the 2/ l component has hardly any impact on the result of the l expression:
This implies that z2 = 1,975 m in (13.3b) would give almost the same elongation l.
The ULS equilibrium situation is shown in fig. 13.28. The figure presents the forces exerted
on the concrete. The increase of the axial prestressing force is 5,48 MN. Note that the initial
prestressing force is 10,5 MN, which makes that the total prestressing force now is 15,98 MN.
δ = 0,190
0,79
10,5+5,48=15,98
16,5 m
Fig. 13.28 Equilibrium situation at the onset of failure (forces in MN, displacements in m)
The deflection at midspan is 0,190 m, which is relatively high when compared with the height
of the cross-section. The rigid body rotation from the deflection is therefore taken into
account, which influences two parameters:
1 – The angle over which the tendon is kinked at the deviation point.
2 – The vertical position of the horizontal forces exerted at midspan.
The deflection at the deviator at 11 m from the support is 11 / (0,5 · 33) = 2/3 of the deflection
at midspan:
Bending moment equilibrium about the centre of the bottom flange at the support results in:
1
/2 q l · 1/4 l = 1/8 q l 2 = 15,98 · (2,50 - 0,25/2 - 0,15/2 - 0,190) + 11 · 2,48 = 61,0 MNm
In this expression only the dominant bending moment contributions are taken into account.
The first component after the = sign is the contribution of the (axial) tendon force
(15,98 MN); the second component is from the vertical component of the tendon force at the
deviator. It is noted that the internal lever arm between both compressive forces in the flanges
(bottom flange at support and top flange at midspan) is reduced by the deflection δ (0,19 m)
that is required to have the assumed increase of the tendon force.
To find the global load-deflection relation (fig. 13.29), the deformation in the “uncracked
stage” (i.e. no joint opening) is calculated. For q = 231 kN/m1 the deflection is:
1 q l4 1 0, 231 334
6,8 103 m 6,8 mm
384 EI 384 30 103 3,51
For q = 299 kN/m it is found that = 8,8 mm when using the “uncracked stage” stiffness.
This is too optimistic since joint opening at the support then has already started (the joint in
the span is still closed at this load; opening this joint too requires app. 108 – 89 = 19 kN/m
additional loading).
500
450
400
350
q load [kN/m]
300
250
200
150
100
50
0
0 50 100 150 200
displacement [mm]
The safety with regard to failure turns out to be very large. The requirement is:
This high safety is especially due to the short free length of the tendon, which is assumed to
be only 33 m (cracks at all spans and supports are assumed, which makes that an elongation
results in a relatively high strain increase), and the relatively high prestressing force, which
creates high compressive forces in the joints, see fig. 13.25 (right).
Bending moment equilibrium about the centre of the bottom flange at the support resulted in
the following expression:
1
/2 q l · 1/4 l = 1/8 q l 2 = 15,98 · (2,50 - 0,25/2 - 0,15/2 - 0,190) + 11 · 2,48 = 61,0 MNm
At a small angle , tan = sin = . The upward force at the deviator then is:
2, 48 MN = Pm P
where
2,5 0, 4 0,5 0,190 23
11
0,15 0, 25
1
8 ql 2 = Pm P 2,50 - 0,50 - 0,50 2,50 0, 40 2,50 0, 40
2 2
2,50 0, 40 0,50 23
13 l Pm P
3l
1
0,15 0, 25 1
1
8 ql 2 = Pm P 2,50 - 0,50 - 2,50 0, 40 3
2 2
1
8 ql 2 = Pm P dsup port d midspan 13
The expression contains the total bending moment from the external q-load ( 18 ql 2 ). The
external bending moment is resisted by both the support and midspan cross-section. It appears
that their internal lever arms as defined in the stage of joint opening (dsupport, dspan) must be
reduced ( 13 ) This reduction follows from the vertical displacement of some of the
deviators:
Rewriting the bending moment equilibrium expression, as presented before, is not that
obvious. Therefore, this expression will be derived by using a number of variables, see the
following figure. Thanks to symmetry, only half of a span has to be modelled.
Bending moment equilibrium around the position of the concrete compression zone force at
the support cross-section:
1 1 1
Pm P zB epB epA zA Pm P sin( ) l ql l
3 2 4
e e
pA pB
1 1
l l
3 2
Pm P zA zB ql 2
1 1
3 8
Note that three bending moment components were not taken into account since they have only
minor impact (both a small force and a small lever arm).
The deviators present at the supports prevent vertical tendon displacement to occur. The
internal lever arm at the support therefore does not change compared with the lever arm at
initial joint opening. At midspan, however, there is no deviator; the deviators are at 5,5 m
from the midspan cross-section. A vertical displacement at midspan implies a vertical
displacement of 23 at these deviators. The tendon follows the deviator and, as a result, there
is a 13 reduction of the midspan internal lever arm (the hinge in the midspan compression
zone has a vertical displacement , whereas the tendons has a displacement of only 23 ).
It is noted that this calculation is based on assumed plasticity, where plastic hinges develop at
the support and midspan cross-section and resist the total bending moment exerted.
Redistribution of bending moments from support to midspan or vice versa is assumed to
occur, which requires sufficient rotational capacity.
T
(13.5)
h
where:
However, the curvature is restrained by a bending moment M. This moment follows directly
from the well-known relationship:
M
(13.6)
EI
Fig. 13.30 Imposed deformations by a temperature gradient (a) and differential settlement (b)
At a temperature cycle of about 12 hours (sun) the short term E-modulus is not valid (EN
1992-1-1 table 3.1). Therefore, a reduction of 15% on this E-modulus is applied.
So one finds:
This implies that, due to this temperature gradient alone, no joint opening would occur
(compare the value 1,6 N/mm2 with the compressive stresses at the bottom of the beam from
fig. 13.25).
In the case of a differential settlement z (fig. 13.30b) in the uncracked stage, the bending
moment is:
6 EI z
Mz
l2
In general, settlements occur gradually. According to section 6.7.4 a gradual increase of the
imposed deformation results in a reduction of the magnitude of the bending moments.
The reduction factor is:
(13.7)
(1 )
Mz = 1,7 MNm
To have joint opening at the support in the unloaded situation (fig. 13.25; “support’, “G”,
bottom; -1,8 N/mm2), a differential settlement of (1,8 / 0,77) · 10 = 23 mm is required. In the
fully loaded situation (fig. 13.25; “support”, “G+Q”, bottom; -3,4 N/mm2), (3,4 / 0,77) · 10 =
44 mm leads to joint opening.
If the mean prestressing is lower, which is possible on the basis of the traffic load, see fig.
13.25, the differential settlement required to have joint opening, would be lower.
At the bottom of the beam joint opening occurs already at a differential settlement of 23 mm
(unloaded situation). It should be noted that, if there is a significant joint opening, this causes
a reduction of the shear resistance of the structure. This shows that a good estimation of the
effect of imposed deformations is of major importance.
In a structure without segmental joints, where reinforcing steel can continue along the full
structure, the sensitivity to imposed deformation is much smaller.
A structure that cracks is discontinuous at the location of the crack; there is locally a discrete
separation between the structural parts at both crack faces. As a result of this discontinuity, an
unbonded tendon that crosses the crack and first had a smooth (linear or curved) profile, now
has a discontinuous profile at the crack. This causes kinks in the tendon profile. This is
demonstrated in fig. 13.31 for simply supported beams with a linear and a curved tendon
profile. As discussed in chapter 12, the increase of the stress in the prestressing steel follows
from the crack width at the level of the tendon and the tendon length. The bending moment
resistance of this type of structure is now discussed.
Fig. 13.31 Failure mode of simply supported uniformly loaded beams containing linear (left)
and curved (right) unbonded tendons
The simply supported beam with linear tendons is discussed first, see fig. 13.32. The beam
has a rectangular cross-section. The tendon eccentricity relative to the centroidal axis is
denoted as ep. The beam is uniformly loaded to failure and cracks at midspan only.
Fig. 13.32 Deformations and force equilibrium of a simply supported uniformly loaded beam
with linear unbonded tendons
At a deflection δ at midspan, the angle θ, the elongation of the tendon and the increase of
prestressing steel strain are:
0,5
p 2 z
p
pm
The increase of the prestressing force ΔP follows from the total steel strain εp,tot (initial strain
εpm∞ plus strain increase Δεp) and the stress-strain diagram of the steel. The total prestressing
steel force is denoted as Pm∞ + ΔP. At a relatively small angle θ (sin = and cos = 1), the
horizontal and vertical component of the prestressing force can be expressed as indicated in
fig. 13.32b. The figure presents only the left half of the beam, including the forces acting on
it. Note that at the crack, there is only the vertical component of the prestressing force caused
by the kink in the tendon. The force at the anchor is split in a horizontal and a vertical
component.
The concrete compressive force at the assumed "hinge" in the concrete compression zone
follows from the horizontal component of the total prestressing force:
N cu Pm P
The "hinge" is assumed to be at a distance ηh from the outermost concrete compression fibre.
Bending moment equilibrium about the support:
1
8 ql 2 = Pm P h h 12 l 12 h+ep Pm P 12 h h ep
At the left hand side of the = - sign is the total bending moment in the midspan cross-section
from the uniform load on the beam; at the right hand side is the internal bending moment in
this cross-section. It appears that the bending moment capacity is not reduced by the
deflection of the beam as should be the case when it is an external tendon. This becomes clear
when looking at fig. 13.32: The deflection of the beam not only influences the working line of
the concrete compressive force, but is also imposed on the tendons. These two aspects have
the same quantitative effects, but are opposite. There is only one effect from the deflection at
midspan, which is the increase of the prestressing force from tendon elongation.
The following example is similar to the previous one. The only difference is that the
unbonded tendons now have a parabolic profile, see fig. 13.33a. The tendons are anchored at
the level of the centroidal axis. The tendon drape is f.
Fig. 13.33 Deformations and force equilibrium of a simply supported uniformly loaded beam
with parabolic unbonded tendons. The beam is assumed to be cracked at midspan
Again, the elongation of the tendons results in a prestressing steel strain increase:
0,5
p 2 z
p
pm
The total prestressing steel force at failure is again denoted as Pm∞ + ΔP. At their anchors, the
tendons are at an angle α relative to the centroidal axis of the beam. This angle follows from
the tendon profile. In case of a parabolic profile α = 4f / l if the angle α is relatively small. The
tendons are kinked over an angle θ at both crack faces. At small angles α and θ (sin(α + θ) =
α + θ and cos(α + θ) = 1) the horizontal and vertical component of the prestressing force can
be expressed as shown in fig. 13.33b. At the crack, there is again only the vertical component
of the prestressing force. At an anchor, the angle between the tendons and a horizontal line
has increased from α to α + θ.
The concrete compressive force at the assumed "hinge" in the concrete compression zone
follows from the horizontal component of the total prestressing force:
N cu Pm P
The "hinge" is again assumed to be at a distance ηh from the outermost concrete compression
fibre.
Figure 13.33b presents the left half of the beam. Forces exerted on the concrete are presented,
namely:
1. the uniformly distributed load on the beam;
2. the uniformly distributed upward curvature pressure from the tendons;
3. at the anchor: a horizontal and a vertical component of the prestressing force;
4. at the cracked cross-section: a horizontal concrete compressive force and a vertical force
caused by the kink in the tendon;
5. at the support: a vertical support reaction;
2:
The prestressing force has increased from Pm∞ to Pm∞ + ΔP. The curvature of the tendon
profile has not changed; the two parts in which it is split by the crack, have rotated only over
an angle θ. The curvature pressure is (< 0, which is upward):
Pm P
qp
R
The curvature pressure is not vertical; it is at an angle α + θ relative to a vertical line at the
support, gradually reducing to an angle θ at midspan. The angles are small. As a result, the
vertical component of the curvature pressure is equal to the curvature pressure. There is also a
horizontal component of the curvature pressure. This force follows from sin(α + θ) = α + θ at
the support, gradually reducing to sin(θ) = θ at midspan. The resulting horizontal force,
however, is relatively small and has a small lever arm since the lever arm is related to the
beam height and the vertical position of the tendon. It will therefore have a negligible
contribution to the bending moment equilibrium.
3:
Assume that the original tendon profile was at the supports at an angle α relative to the
centroidal axis. When the beam is not loaded, it is the angle relative to a horizontal line. The
deflection of the beam causes its two parts to rotate over an angle θ. The angle between a
horizontal line and the tendon profile now is α + θ at the anchors. The forces at an anchor:
5:
The vertical support reaction is in equilibrium with the load on the beam. Note that the
uniform upward load from prestressing does not result in a support reaction.
2f 4f
f 14
1
2
4 ( Pm P) 14
Pm P Pm P 0 ; OK
Note that it is a statically determinate beam. Prestressing then does not introduce vertical
support reaction forces. On the other hand, in case of a statically indeterminate structure,
vertical reactions forces are often introduced.
8 ( Pm P) f 1
N cu h h Pm P 12 l+ 12 4 l 12 ql 14 l+ Pm P 12 h (13.9)
2
1
8 ql 2 = Pm P h h 12 l 12 h+f Pm P 12 h h f
Again, the deflection has no influence on the internal lever arm of forces, just as in case of the
beam with linear tendons. The tendons follow the vertical displacement of the beam and the
vertical distance between the concrete compressive force and the tendon remains unchanged.
The parabolic tendons are now applied in a statically indeterminate beam. The tendon profile
is modelled using upward curvatures only; the downward curvatures at the supports are
replaced by kinks, see fig. 13.34. The drape of the tendons is f, which is the distance between
the upper and lower point of the parabolic tendon profile. The angle between the tendon
profile and a horizontal line at the supports is again denoted as α. The distance from the
outermost concrete compression fibre to the resulting compressive force (at the "hinges") is
η1 h and η2 h at the midspan and support cross-section, respectively.
Figure 13.34 presents the forces that are relevant from the point of view of bending moment
equilibrium.
Fig. 13.34 Statically indeterminate beam prestressed with unbonded tendons. The tendon
profile is assumed to have upward curvatures only. The beam is uniformly
distributed loaded and cracked at midspan and support cross-sections
8 ( Pm P ) f 1
N cu h 1h Pm P 12 l+ 12 4 l 12 ql 14 l+N cu 2 h (13.10)
2
1
8 ql 2 = Pm P h 1h 2 h 12 l+f Pm P h 1h 2 h f
As expected, the deflection does not influence the internal lever arms of forces. If the distance
from the top fibre of the beam to the tendons is denoted as y at the support, the distance
between the concrete compressive force and the tendons is:
The summation of both internal lever arms, namely h 1h 2 h+f , is used in expression
(13.10) which also contains the total bending moment from the external q-load that has to be
resisted by the support and midspan cross-section ( 18 ql 2 ). This indicates that this calculation
is based on assumed plasticity: plastic hinges develop at the support and midspan cross-
section and together, they must resist the total bending moment exerted. Redistribution of
bending moments from support to midspan or vice versa is assumed to occur. This implies
that there must be sufficient rotational capacity.
Within the scope of such “service” activities, Seltenhammer described the upgrading
operations for the Wangauer Bridge [13.20], see also Wicke [13.21]. As a second example of
the use of external prestressing, namely the renovation of the Ruhr Bridge Essen/Werder, will
be considered [13.15].
span support
axis
Fig. 13.35 Cross-section of the Wangauer Ach Bridge before refurbishment (dimensions m).
25 Years after its completion, it was decided to refurbish the bridge. Since meanwhile also the
traffic loads were increased and the design codes were changed, it was decided to provide
additional prestressing. Since it is almost impossible to provide additional bonded prestressing
in an existing structure, it was decided to use external prestressing.
The next decision was to apply the prestressing tendons axially, at the level of the gravity line
of the structure, and not to apply them according to the bending moment curve. This decision
was predominantly taken on the basis of technical arguments. Although a tendon profile, that
is adapted to the distribution of moments along the structure, is most effective and economic
with regard to the use of prestressing steel, a problem is the introduction of the high
prestressing forces in the structure at the deviation points. To apply a saddle in a structure as
shown in fig. 13.35, is a difficult and expensive task. In a bridge having a length of 385 m, the
number of saddles, intermediate anchorages and couplers is very large. As shown in fig.
13.32, the tendons have been placed at the outside of each web. An additional prestressing
force of 5000 kN per web was applied on the structure (fig. 13.36).
refurbished
additional
prestressing
Fig. 13.36 External prestressing at the outside of the webs of the Wangauer Ach Bridge
Special attention was given to anchoring the tendons at the bridge ends. The tendons are
distributed over the height of the webs to have a more uniform load introduction and to create
sufficient space for the anchorages. To introduce the prestressing force in the structure and to
spread it over its width, a force (or a tensile tie) is required in transverse direction (compare
fig. 10.5 and 10.6). Therefore, the old transverse end girder was extended with an additional
new part. In this newly cast part, three additional tendons were placed to balance the splitting
force from load introduction (fig. 13.37).
additional concrete
Fig. 13.37 Strengthening of the transverse end beam in front view (top) and in horizontal
cross-section (bottom), including the position of the old and new prestressing
tendons.
Crack formation had occurred in the bottom flange and in the beam webs of the largest span.
The cracking was that severe, that repair was inevitable because the post-tensioned steel in the
cracked area had to be protected against corrosion. Injecting the cracks (which locally showed
widths up to 0,4 mm) was not regarded as being a satisfactory solution, because calculations
demonstrated that these cracks were caused by temperature gradients. It was therefore
expected that, very soon after injection of the old cracks, new cracks would occur for the
same reason. Furthermore, it was shown that the stress variations in the post-tensioned steel
were far beyond values allowed. It was clear that measures were not only necessary for
corrosion protection, but from the point of view of structural safety as well. Also here,
strengthening with external straight tendons appeared to be the best solution.
longitudinal cross-section
cross-section
construction
joint
plan view
At the abutments, a total of 24 longitudinal tendons (VSL Type 5-16 with a breaking load of
2833 kN), each about 75 m long, were placed. At the abutments, they were anchored in
anchorage blocks that were cast at the longitudinal beam (web) ends, see fig. 13.39. Also the
transverse beam at the intermediate support was provided with additional vertical prestressing
to sustain the forces from the external longitudinal prestressing tendons.
plan view
tendon
tendon
a detail A
detail B
anchor tendon
tendon anchor
Fig. 13.39 Longitudinal external prestressing applied in the longest span (a), anchored in
newly cast in-situ parts of the longitudinal beams (b). Strengthening of the
transverse girder at the intermediate support by external vertical prestressing (c)
(dimensions mm)
13.8 Literature
13.1 Bruggeling, A.S.G.: “External Cables: State of the Art Report”, Conference on Partial
Prestressed Concrete Structures.
13.2 Combault, J.: “Evolution et développement des ponts modernes a précontrainte
totalement extérieure au béton”, FIP-Congres New Delhi, 1985.
13.3 Müller, J.: “Construction of the Long Key Bridge”, PCI-Journal, Nov-Dec. 1980.
13.4 Virlogeux, M.P.: “Die externe Vorspannung”, Beton- und Stahlbetonbau 83, Heft 5, p.
121-126.
13.5 Virlogeux, M.P.: “External prestressing: From Construction History to Modern
Technique and Technology”, Naaman, A., Breen, J. (eds.): External Prestressing in
Bridges, ACI-SP 120, Detroit 1990.
13.6 Virlogeux, M.P.: “La Précontrainte Extérieure. Le point de la question aujourd’hui. La
Conception et la Construction des Ponts a Précontrainte Extérieure au Béton”, Annales
de l’Institut Technique du Bâtiment et des Travaux Publics 1991, No. 498, p. 1-47,
No. 499, p. 1-75.