You are on page 1of 21

Journal of Building Engineering 63 (2023) 105434

Contents lists available at ScienceDirect

Journal of Building Engineering


journal homepage: www.elsevier.com/locate/jobe

Enabling sustainable rapid construction with high volume GGBS


concrete through elevated temperature curing and
maturity testing
Fragkoulis Kanavaris a, *, Marios Soutsos b, Jian-Fei Chen c
a
Technical Services, Materials, ARUP, London, UK
b
School of Natural and Built Environment, Queen’s University Belfast, Belfast, UK
c
Department of Ocean Science and Engineering, Southern University of Science and Technology, Shenzhen, China

A R T I C L E I N F O A B S T R A C T

Keywords: Nowadays, low carbon footprint concrete for construction relies heavily on ground granulated
Ground granulated blast furnace slag (GGBS) blast-furnace slag (GGBS) as a partial cement replacement material (CRM) in places where other
Temperature sensitivity CRMs are in short supply. However, it is relatively well-known that the low early-age strengths of
Compressive strength GGBS concretes discourage the maximisation of cement replacement in most applications; a
“Apparent” activation energy constraint which can be potentially overcome through exploitation of hydration acceleration
Maturity functions
under elevated temperature curing. Concrete and mortar mixes of 47 MPa 28-day target mean
cube strength were developed and investigated in this study with various percentages of GGBS (0,
20, 35, 50 and 70%) and cured under isothermal and non-isothermal regimes (20, 30, 40 and
50 ◦ C and adiabatic). Higher temperatures appeared to significantly accelerate the strength gain
of GGBS concretes, particularly those containing high GGBS percentages. In-situ strength devel­
opment may be estimated through maturity functions which were initially developed for neat
Portland cement concretes. The accuracy of several maturity functions, such as the Nurse-Saul,
Arrhenius, Weighted Maturity, Weaver-Sadgrove and Rastrup ones, were examined together
with two strength-maturity/time correlations. It was found that although maturity methods can
be used to optimise a concrete mix in terms of GGBS content and depending on the application, it
is not possible to obtain consistently reliable estimates for GGBS concretes from the current
functions. Nonetheless, from the current models considered, the Arrhenius, Weighted Maturity
and Rastrup functions appear as more appropriate for higher replacement levels of cement with
GGBS. Overall, the present study highlighted a need for further improving maturity functions to
account for the strength development of GGBS concrete.

1. Introduction
The increasing trend in utilising concrete incorporating Cement Replacement Materials (CRMs) has been evident over the last years
[1–8]. Nowadays, contractors and precast concrete manufacturers are required to replace the Portland cement (PC) in concrete which
is potentially detrimental for the environment since its manufacture accounts for 6–8% of worldwide anthropogenic CO2 emissions [9,
10], with alternative and more sustainable binders, such as ground granulated blast furnace slag (GGBS) which is a by-product from

* Corresponding author.
E-mail addresses: frag.kanavaris@arup.com (F. Kanavaris), m.soutsos@qub.ac.uk (M. Soutsos), chenjf3@sustech.edu.cn (J.-F. Chen).

https://doi.org/10.1016/j.jobe.2022.105434
Received 4 August 2022; Received in revised form 6 October 2022; Accepted 16 October 2022
Available online 1 November 2022
2352-7102/© 2022 Elsevier Ltd. All rights reserved.
F. Kanavaris et al.
Table 1
Maturity and equivalent age models considered.

Name and ref. Maturity/Equivalent age Age conversion factor (β) Notation Comments
model
∑ –
Nurse-Saul M = (T − T0 ) • Δt (Eq. 1) M is the maturity (◦ C-days), Saul [32] firstly introduced the maturity concept/index. T0 (temperature
t
[32] T is the average temperature (20 ◦ C for standard which below it is assumed that no strength gain occurs) taken as − 11 ◦ C,
curing) over the time interval Δt (◦ C), which is the average of what is recommended in the literature [33–35]. The
T0 is the datum temperature (◦ C), Nurse-Saul function and assumes that the rate of strength development is a
Δt is the time interval (days) linear function of temperature

Nurse-Saul [31, (T − T0 ) (T − T0 ) te is the equivalent age at the reference temperature The ratio β, which is called the “age conversion factor”, converts a curing
te = • Δt (Eq. 2) β= (Eq. 4)
(T − T ) (Tr − T0 ) (days), interval Δt to the equivalent curing interval at the reference temperature [31].
32] ∑ r 0
te = β • Δt (Eq. 3) Tr is the reference temperature (◦ C).
Rastrup [36] (T − Tr ) (T − Tr ) β is the age conversion factor Rastrup [36] introduced the concept of equivalent age.

te = 2 10 • Δt (Eq. 5) β = 2 10 (Eq. 6)
) ( T + 16 )2
Weaver and ∑ ( T + 16 2
te = • Δt (Eq. (Eq.8)
2

Sadgrove Tr + 16
β=
Tr + 16
[37] 7)
( ) ( )
Arrhenius [38] Ea 1 1 Ea 1 1 Ta is the average temperature of concrete during time The “apparent” activation energy is calculated according to ASTM Standard
∑ − R T − T R

T

T interval Δt (◦ K), C1074-11 [39], [27,40–42]
t = e
e
a s • Δt β=e a s (Eq.
(Eq. 9) 10) Ts is the specified reference temperature (◦ K),
Ea is the “apparent” activation energy (J/mol),
R is the universal gas constant (J/K⋅mol).

Weighted MW = t × T × Cn (Eq. 11) C0.1T− 1.245 − C− 2.245 Mw is the weighted maturity (◦ C⋅hours or ◦ C⋅days) t is
β=
Maturity C0.1Tr − 1.245 − C− 2.245 the age/time of concrete (hours or days)
[43–45] (Eq. 12) T is the average concrete temperature during time
interval Δt (◦ C) n is a temperature dependent
parameter

Journal of Building Engineering 63 (2023) 105434


C is a cement specific constant for which the strength
maturity curves for isothermal strength tests at 20 and
65 ◦ C coincide, C – cement specific value.
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

the steel industry. Apart from sustainability reasons, CRMs are also required for improving the durability of the concrete. However, the
addition of CRMs and more specifically, GGBS, significantly alters the early-age strength development of concrete [11,12] and thus, it
is not used in applications where high early-age compressive strength is required, i.e. removal of formwork/falsework as soon as
possible and in precast concrete factories for safely lifting of precast units as early as 16 h after casting.
The ambition for coal fired power stations to cease operation within the next few years in Europe and elsewhere, for sustainability
and climate change reasons, has led countries to close many of such facilities [13], e.g., in the UK it is expected that all coal-fired power
stations would be closed by 2025 whilst currently there are only four active. This has led to fly ash shortage in several regions, a CRM
which had traditionally been used in concrete for several decades. With low carbon concrete being of ever-increasing importance, the
concrete industry, designers and specifiers rely now heavily on GGBS with the intention consistently being to maximise the GGBS (e.g.,
up to 70%) percentage of the total binder in structural concrete, so that concrete’s carbon footprint is reduced [14–16]. This, however,
causes problems to contractors, e.g., potential issues with construction times, where such implications are more pronounced in densely
populated urban areas which require rapid construction and minimum disruption. Typical situations where such problems are
encountered are: (i) in core walls of high-rise buildings in which climbing formwork (jump form) is employed requiring early strengths
so that the formwork system advances to the next level, usually within 12–24 h from casting [17], (ii) in precast concrete elements
which usually require early strengths within 15–18 h from casting in order to be lifted from the casting “beds” and maintain an efficient
element production schedule [18], (iii) suspended slabs, floor slabs or ground slabs which require a certain surface treatment, e.g.,
power-floated/trowelled or application of other finishes, only few days after casting or even (iv) slip-forming operations which are
known to require very specific early concrete strength development [17]. Experience has shown that even nowadays, relatively small
GGBS quantities may be employed in these applications, for example up to 20% in (i) and (ii), up to 30–40% for (iii) and even 0% for
(iv), also because contractors cannot quantify the expected early-age strength deceleration once GGBS content is increased. This has a
significant negative impact on the carbon footprint of the concrete used in such applications and highlights the need to explore ways to
maximise GGBS addition without affecting the predominantly time-driven construction schedules.
Elevated temperature has been widely used as a method to accelerate the cement hydration process and therefore, the strength gain
rate [19,20], e.g., in precast concrete manufacturing. With GGBS being much more temperature sensitive than PC [21–27], there is
potential of exploiting in-situ concrete temperatures or early high curing temperatures in precast concrete production in order to
maximise the percentage of GGBS in the mix without necessarily compromising the strength of the concrete. To achieve this, it is
important to investigate and quantify the effect of temperature on the early-age strength development of concretes containing various
cement replacement levels with GGBS. Subsequent relevant guidance on the curing regime required to achieve high early-age strength
without negatively affecting the long-term strength of concrete may then be provided, as high early-age curing temperatures are
responsible for the “cross-over” effect which results in lower strengths at later ages [23].
The combined effects of temperature and time on concrete strength always created interest for engineers/researchers working on
rapid concrete construction. This, in addition to performing concrete strength-driven critical operations on site and avoiding con­
struction failures, such as those reported in Lew et al. Feld and Carper and Kaminetzy and Stivaros [28–30], has led to the development
and potential application of the so-called “maturity” methods, which account for the combined effect of temperature and time on the
strength development of concrete [19,31].

2. Maturity functions
2.1. Maturity and equivalent age models
There have been several maturity functions developed over the last decades for predicting the in-situ strength of Portland cement
concrete. This section encompasses fundamental information of the most popular maturity functions, which are also considered in the
present study. There mainly are two types of maturity functions; those based on the calculation of the Maturity Index or Temperature-
Time Factor [32] and those based on the calculation of the Equivalent Age [31]. Equivalent age represents the duration of the curing
period at the reference temperature that would result in the same maturity as the curing period at other temperatures [31]. Table 2
summarises the characteristics of the maturity and equivalent age models considered herein.
With relevance to the Arrhenius function included in Table 1, the “apparent” activation energy, which is a temperature sensitivity
index of a certain mix, can be determined using “equivalent” mortar specimens, as described in ASTM Standard C1074-11 [39] and the
results may be applied to the concrete under investigation. This requires the determination of the compressive strength development
under at least three curing temperatures. The concept of “equivalent” mortars suggests that mortars having the same w/b ratio as the
concrete and the fine aggregate to binder ratio to equal to the coarse aggregate to binder ratio of the concrete, exhibit comparable
strength and hydration behaviour compared to that of the concrete under investigation. However, it has been reported in several cases
that “equivalent” mortar specimens are not necessarily representative of concrete specimens [27,40–42] which raises concerns when it
comes to their appropriateness in replicating concrete properties.

2.2. Strength-time relationships


Regression analysis using an appropriate descriptive mathematical formula is needed to relate compressive strength to age or
maturity index [46–53]. This is normally accomplished through certain strength-time relationships which are compatible with
maturity and equivalent age models. The following S-shape function proposed by Carino [52] (Equation (13)) is the one recommended
in the ASTM Standard [39]:

3
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

Su • k • (t − t0 )
S= (13)
1 + k • (t − t0 )

where: S is the compressive strength at age t (MPa),


Su is the ultimate compressive strength at temperature T (MPa),
k is the rate constant (1/days),
t is the test age (days),
t0 is the age at which strength development is assumed to begin (days).

Regression analysis is carried to obtain for each curing temperature the rate constant, k, the ultimate strength, Su, and the setting
time, t0, of the mortar mixture.
In order to calculate the apparent activation energy, Ea, the ASTM Standard’s recommendation [39] is to plot ln(k) against 1/Tabs
given in 1/Kelvin), where Tabs is the absolute curing temperature. The slope of the trend line is equal to -Q and the activation energy
(Ea) for the mixture will be equal to Q⋅R, where R is the universal gas constant equal to 8.31 J/K⋅mol.
Another widely used strength-maturity/age relationship is that developed by Freiesleben Hansen and Pedersen (FHP), which has
also become known as the Three Parameter Equation (TPE) [53], expressed as follows:
( )a
− τ
(14)
t
S = Su • e

where: S is the compressive strength at age t (MPa),


Su is the ultimate compressive strength at temperature T (MPa),
τ is the characteristic time constant (days),
t is the test age (days),
α is the shape parameter (taken as 0.7 in this study).
The “apparent” activation energy can be also determined using the FHP relationship, following the same procedure described
above, albeit by plotting ln(τ) against 1/Tabs (given in 1/Kelvin). It has to be noted that due to the fact that there does not seem to be
any consistent relationship of the shape parameter, α, with temperature, the shape parameter was set to be equal to 0.7 rather than
being left as variable, as this value was found to yield better fit at all temperatures [42,54].
It is recognised that the above is a non-exhaustive description of maturity models and that other strength-maturity relationships
and maturity functions have been proposed, such as those of Plowman [55] and Guo [56], however, these are not deemed particularly
popular or are known to exhibit certain flaws. Hence, the present study considers those which can be considered more popular in the
concrete engineering sector.

3. Research significance, methodology and objectives


3.1. Research significance
The aforementioned maturity functions were developed for predicting the compressive strength development of Portland cement
concrete. This generates doubts regarding their wider applicability in concretes with high quantities of GGBS. Gaining confidence on
the applicability of current maturity functions in predicting GGBS strength development will enable designers and contractors to adopt
more sustainable construction.
Furthermore, maturity methods currently lack standardisation, e.g., only the Nurse-Saul and Arrhenius functions are covered by
ASTM C1074-19 (USA) and the Weighted Maturity by NEN 5970 (Netherlands) [57] whilst official recommendations are given on the
use of the maturity method in Hong Kong by Ho et al. [58] and practical ones in France by IFFSTAR [59]. It is believed that this is
mainly associated with the uncertainty these methods exhibit with relevance to the prediction of concrete strength evolution,
particularly for concretes containing CRMs, such as GGBS. This, subsequently, restricts wider application of these methods in the
concrete industry. Therefore, further investigating the applicability of these methods in concretes other than Portland cement-based
will contribute to a better understanding and the capabilities and limitations of maturity functions and provide corresponding
recommendations.

3.2. Research methodology and objectives


The concrete mixes investigated in this study were designed to have 47 MPa 28-day target mean cube compressive strength, as this
class of normal strength concrete is widely used in a variety of structural concrete applications [2,4,60] (strength class C30/37
allowing for a strength margin of at least 10 MPa in accordance with the widely accepted BRE guidance for design of mixes [61]) and
several precast concrete applications [6,62,63]. “Equivalent” mortar mixes were also tested for compressive strength in order to: (a)
underline the main discrepancies with concrete strength development. (b) Provide guidance as to whether mortar specimens can be
used to evaluate early concrete strength development and (c) provide guidance as to whether they can be used calibrate maturity
models for concrete strength prediction. The effect of temperature on the compressive strength development of concretes with five
GGBS replacement levels, i.e. 0, 20, 35, 50 and 70%, was investigated and the applicability/accuracy of existing maturity functions in
estimating the compressive strength development of concrete with GGBS was assessed. The following were the objectives:

4
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

1. Investigate the appropriateness of using “equivalent” mortars to replicate/evaluate concrete properties, such as compressive
strength.
2. Investigate the effect of temperature on the compressive strength development of GGBS concretes, rather than mortars.
3. Provide guidance regarding curing regimes required for GGBS concretes to achieve high compressive strength at early ages without
compromising later age strength.
4. Determine the “apparent” activation energy of GGBS concretes, rather than mortars and compare results between calculation
methods, i.e. the ASTM and FHP-TPE methods.
5. Investigate the applicability/accuracy of existing maturity functions in estimating the compressive strength development of GGBS
concretes under non-isothermal curing regimes.
6. Quantify the extent to which the accuracy of the estimates is affected by the strength-maturity/age relationship used, i.e. the ASTM
(Carino) and FHP-TPE formulations.

4. Materials and experimental methods


4.1. Materials
Portland cement (CEM I 42.5 N), conforming to the requirements of BS EN 197–1:2011 [64], was supplied in bags by Quinn cement
and had 60 MPa 28-day compressive strength (tested in accordance with BS EN 196–1:2016 [65]). GGBS, which was used as partial PC
replacement material, conformed to BS EN 15167–1:2006 [66] and was also supplied in bags by Civil and Marine Ltd-Hanson. It must
be noted that the same materials were used throughout the experimental investigation of this study. The chemical compositions of PC
and GGBS are shown in Table 2.
Both of the aggregates used in this study were sourced locally in Northern Ireland, supplied by the quarry of Creagh Concrete
Products Ltd. The coarse aggregate used was 4–10 mm crushed basalt having specific gravity and water absorption of 2.76 and 1.85%,
respectively. The fine aggregate used was sand from Lough Neagh with specific gravity and water absorption of 2.67 and 1.02%,
respectively. The grading curves for the coarse and fine aggregates are shown in Fig. 1 alongside the grading limits described in BS EN
12620:2002 + A1:2008 [67].

4.2. Mixing, casting, curing and testing procedures


All concrete mixes were batched using a 0.1 m3 capacity horizontal pan mixer while all mortar mixes were batched using a 0.005
m3 capacity mixer. Aggregates, which were oven-dried for at least 24-h, were placed first together with the vinder in the mixing pan.
Following dry-mixing for 1 min, water was added and the mixing continued for another 5 min. The workability (consistency) of
concrete was determined by carrying out the slump test, in accordance with BS EN 12350–2:2009 [68]. Concrete cube (100 mm)
specimens conforming to BS EN 12390–1:2012 [69] and mortar cube (50 mm) specimens as per ASTM C1074-19 [39] were the cast in
two layers and each layer was compacted using a vibrating table.
The following curing procedures were adopted (see Fig. 2 for set ups):
• Standard curing, for which the concrete and mortar specimens were kept inside the moulds for 1 day after casting, covered with
damp hessian and polyethelene sheet under room temperature (approximately 20 ◦ C) and were placed into 20 ◦ C water tanks at 1
day after casting onwards.
• Elevated isothermal curing, for which the concrete specimens were placed into water tanks with water temperatures set at 30, 40
and 50 ◦ C immediately after casting. The specimens in the moulds were wrapped with cling film and tape to avoid paste/mortar
leaching. They were demoulded at 1 day after casting, or earlier depending on the testing age and immediately placed back into the
water tanks.
• Adiabatic curing – The adiabatic temperature rise is produced by the cement hydration and will occur if concrete is placed in a
perfectly insulated environment, where there are no heat exchanges or losses, between concrete and surrounding environment. To
achieve such state, special testing arrangements are required. The adiabatic tests were performed using a temperature matched

Table 2
Chemical compositions of PC and GGBS.

Chemical Composition

Chemical Constituent PC GGBS

% by weight of dry mass

SiO2 21 29.38
Al2O3 5.5 11.23
Fe2O3 3 0.36
CaO 62.5 43.72
MgO 2.25 6.94
SO3 2.6 1.76
K2 O – 0.93
Na2O – 1.01
Equiv. Alks Na2Oe 0.525 –
Chloride 0.045 –
Loss of ignition 2 –

5
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

Fig. 1. Sieve analysis of coarse and fine aggregate.

Fig. 2. Equipment and tanks used for curing specimens under (a) standard, (b) elevated isothermal and (c) adiabatic conditions.

curing TMC and an adiabatic curing tank developed in The University of Liverpool and Queen’s University Belfast, and their
operating principles have been outlined in Soutsos and Kanavaris [70]. Concrete cube samples were cured under the adiabatic
temperature profiles of the concretes under investigation and remained in the TMC tanks until testing.
The testing ages for all types of specimens and curing methods are shown in in Table 3. As an attempt to effectively quantify the
effect of temperature on the early age compressive strength development of GGBS concretes, testing commenced as early as 3 h (50 ◦ C),
6 h (30, 40 ◦ C) and 12 h (20 ◦ C) after casting. Concrete compressive strength was also assessed at 16 h after casting as it is usually a
target time set by the precast concrete industry for lifting structural elements and for jump-form operations in in-situ concrete. At each
testing age, 2 specimens were tested from each curing regime in terms of compressive strength with procedures according to BS EN
12390–3:2009 [71].

4.3. Concrete and mortar trial and final mixes


Trial concrete mixes with five different PC replacement levels with GGBS and 47 MPa target mean cube compressive strength (0,
20, 35, 50 and 70% of total binder referred to as PC47, 20GGBS47, 35GGBS47, 50GGBS47 and 70GGBS47, respectively). This required
to initially establish the relationships between w/b ratio and compressive strength by performing trial mixes, considering five different

Table 3
Testing ages for all types of specimens and tests.

Curing regime Specimen type

Concrete cubes (100 × 100 × 100 mm) Mortar cubes (50 × 50 × 50 mm)
a
Isothermal 3, 6, 12, 16 h ,1, 2, 3, 7, 14, 28, 56, 91 and 182 days 1, 2, 3, 7, 14 and 28 days
Adiabatic 12 h,1, 2, 3, 14 and 28 days –
a
Note that testing at 3 h occurred only for 50 ◦ C and at 6 h for 30, 40 and 50 ◦ C due to insufficient concrete hardening at other temperatures at these two testing ages.
Specimens at all temperatures were tested for strength from 12 h onwards.

6
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

w/b ratios (according to Teychenne et al. [61]). The cube compressive strength results of the investigated trial mixes at 3, 7 and 28
days after casting are shown in Fig. 3. The 28-day cube compressive strengths were used to finalise the mix proportions of the
investigated mixes so as they all had similar cube compressive strengths. As anticipated, compressive strength is adversely affected by
GGBS replacement level, especially at early ages and high w/b ratios. This is generally attributed to the decelerated hydration rate with
the inclusion of GGBS.
Results from Fig. 3 (at 28 days after casting) were used to derive the final concrete mix proportions, which are shown in Table 4. As
the mixes were designed to attain a 28-day cube compressive strength of at least 47 MPa, the free w/b ratio decreased and binder
content increased with GGBS addition. This indicates the compromise in binder content associated with higher GGBS replacement
levels and higher strength GGBS concretes. The mix proportions of the “equivalent” mortars were determined according to ASTM
C1074-19 [39] based on the mix proportions of the corresponding concretes. The mortars were required to have the same free w/b
ratio as the concrete and the fine aggregate to binder ratio to be equal to the coarse aggregate to binder ratio of the concrete. The final
mortar mix proportions are also shown in Table 4.

5. Results and discussion


5.1. Strength development of concrete and mortar at 20 ◦ C
The experimentally determined concrete and mortar compressive strength developments under standard curing conditions
(approximately 20 ◦ C) are shown in Fig. 4 for comparison. Only two distinct replacement levels, i.e. 0% (neat PC) and 70% GGBS, are
shown as to provide more context on the concerns that equivalent mortars may not be used to accurately represent the early-age
properties, particularly strength, of concretes. The “equivalent” mortars underestimated the concrete compressive strength
throughout the investigated curing period. This varied from 5% to 34% for the investigated mixes (including the rest of the mixes). At
early ages, the mortars represented relatively accurately the compressive strength of concretes, while the most significant underes­
timation was observed at later ages potentially due to void spacing and surface area of the aggregates. In addition to the compressive
strength development, the rate of compressive strength gain (dS/dt) has also been calculated and is shown in the same figure. The
regression line through experimental data obtained from FHP-TPE (Equation (14)) [53] was used in order to convert the compressive
strength development of concrete and mortar to rate of compressive strength gain. Although the compressive strength development of
concrete and mortar might appear similar, when that is converted to strength gain rate, the discrepancies between them become more
evident. As shown in Fig. 4, the rate of compressive strength gain of the “equivalent” mortar samples is significantly lower from that of
the similar concrete sample. More specifically, it was found that the peak of the rate of compressive strength gain of the “equivalent”
mortar was lower by 25, 33, 36, 42 and 39% for 0, 20, 35, 50, and 70% GGBS replacement level respectively.
It is important to assess the applicability and suitability of “equivalent” mortar mixes in representing accurately concrete prop­
erties, such as compressive strength or “apparent” activation energy. Use of mortars instead of concretes is often preferable, as mixing,
handling, curing, storing and testing of mortar samples is less labour intensive and more convenient than concrete samples, due to their
decreased size [72]. However, the obtained results indicate that mortar mixes cannot replicate accurately the compressive strength
development of concrete and thus, they should not be used for, e.g., obtaining the ultimate (limiting) strengths at standard curing

Fig. 3. Relationship between w/b ratio and cube compressive strength for GGBS mixes.

7
F. Kanavaris et al.
Table 4
Mix proportions of concrete and their equivalent mortars.

Mix ID PC47 20GGBS47 35GGBS47 50GGBS47 70GGBS47

Material Concrete Equivalent mortar Concrete Equivalent mortar Concrete Equivalent mortar Concrete Equivalent mortar Concrete Equivalent mortar

PC [kg/m3] 372 457 330 396 281 334 226 267 139 163
GGBS [kg/m3] – – 83 99 151 180 226 267 325 379
Total binder content [kg/m3] 372 457 413 495 432 514 452 534 464 542
8

Gravel [kg/m3] 1239 – 1217 – 1205 – 1193 – 1188 –


Sand [kg/m3] 616 1500 596 1475 589 1450 581 1435 574 1416
Free water [kg/m3] 190 233 190 228 190 226 190 224 190 222
Total water [kg/m3] 219 248 219 243 219 241 219 239 219 236
Free w/b [¡] 0.51 0.51 0.46 0.46 0.44 0.44 0.42 0.42 0.41 0.41
Total w/b [¡] 0.59 0.54 0.53 0.49 0.51 0.47 0.48 0.45 0.47 0.44
Concrete slump [mm] 105 – 80 – 75 – 75 – 65 –

Journal of Building Engineering 63 (2023) 105434


F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

Fig. 4. Comparison between concrete and mortar compressive strength development.

temperatures which are required by the ASTM-Carino [39,52] and FHP-TPE [53] strength-maturity/time relationships. Moreover, the
different compressive strength development of mortar samples would also affect other parameters inherent in the aforementioned
relationships, such as the shape parameter, α, the characteristic time constant, τ, the rate constant, k and the time (or maturity) at
which strength gain is assumed to begin, t0 (or M0). The above could arguably result in inaccurate compressive strength estimates for
the corresponding concrete. It is therefore important to investigate the effect of temperature on compressive strength of concrete

Fig. 5. Adiabatic temperature rise of investigated concretes cast at 20 ◦ C (±1.5 ◦ C).

9
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

samples rather than mortar samples. This is contrary to what is recommended by ASTM C1074-19 [39].

5.2. Effect of temperature on the compressive strength development of GGBS concretes


Concrete specimens were cured under isothermal and adiabatic conditions in order to determine the effect of temperature on their
compressive strength development. Fig. 5 shows the adiabatic temperature rises of the investigated concretes. The casting temperature
was approximately 20 ◦ C (±1.5 ◦ C) for all mixes. The partial replacement of PC with GGBS by 20 and 35% of total binder appeared not
to be very effective in reducing the maximum adiabatic temperature of concrete, e.g., 20GGBS47 and 35GGBS47 had maximum
temperature rises of 51.0 ◦ C and 48.5 ◦ C whilst that of PC47 was approximately 50.0 ◦ C. It appears that for GGBS replacements of up to
35%, the increased binder content resulted in small decreases of adiabatic heat outputs per kg of binder. These were calculated to be
273, 262, 235, 181 and 144 kJ/kg of binder for PC47, 20GGBS47, 35GGBS47, 50GGBS47 and 70GGBS47, respectively [42] suggesting
that high levels of GGBS replacement are required to effectively reduce peak adiabatic temperature rise, also reported by Barnet et al.
and Soutsos et al. [22,73]. Replacing Portland cement by 50 and 70% with GGBS resulted in reducing the peak adiabatic temperature
rise by 11.0 ◦ C and 18.0 ◦ C, respectively. This suggests that these replacement levels would not only be particularly beneficial in mass
concrete applications where maintaining the peak and rate of temperature rise to a minimum level will reduce the risk of thermal
cracking occurrence [42,60], but also in thin elements where rapid heat dissipation occurs.
As mentioned earlier, all of the GGBS mixes, i.e. PC47, 20GGBS47, 35GGBS47, 50GGBS47 and 70GGBS47, were designed to have
47 MPa target mean 28-day cube compressive strength. Results of the compressive strength development of all mixes under the
investigated curing regimes are shown in Fig. 6. The target mean cube compressive strength of the 47 MPa concretes was successfully
achieved at 28-days, ranging from 47.6 to 48.5 MPa, at standard curing temperature (approximately 20 ◦ C). This work was necessary
in order to identify the effect of temperature on the compressive strength development of concretes, rather than mortars, with and
without GGBS and determine the values of the “apparent” activation energies which will be used later for the maturity estimates.
At standard curing temperature, i.e. approximately 20 ◦ C, PC 47 exhibited faster strength development at early ages when
compared to the mixes contained GGBS. This is a relatively well-known effect of GGBS addition in concrete, namely, leading to slower
early-strength gain [22,42]. Nevertheless, while PC47 seems to be reaching a “plateau” in strength development at later ages, e.g.,
from 28-days onwards, whilst mixes that contain GGBS continue to gain strength even at 182 days since casting and possibly later. This
is particularly the case for high GGBS mixes, such as 50GGBS47 and 70GGBS47 at standard curing temperature. This is attributed to the
pozzolanic reaction and GGBS reacting with Ca(OH)2 resulting to additional CSH formation which contributes to additional strength
gain at later ages [11,74].
The compressive strength development of the investigated concretes is highly dependent on the curing temperatures, as can be seen
from Fig. 6. At early ages, the compressive strength of concretes cured at elevated temperatures was significantly greater than those
cured at lower temperatures. This is attributed to the accelerated cement hydration rate and pozzolanic reaction, resulting from the
increase in curing temperature. However, the elevated curing temperatures, although they effectively increase the early-age
compressive strength development, they also have a detrimental effect on later age strength, i.e. the compressive strength at early
ages at 50 ◦ C is higher than that at 20 ◦ C, but conversely at later ages the compressive strength at 20 ◦ C is higher than that at 50 ◦ C. This
effect is also known as the “cross-over” effect (first reported by McIntosh [75]) and it has been attributed to the formation of dense
hydrated phases around the un-hydrated cement particles, inhibiting further hydration [75,76]. The suggestion is that at elevated
curing temperatures, reaction products are not distributed uniformly within the pores of the hardening paste resulting in coarser and
more porous microstructure of cement paste [76–78].

Fig. 6. Effect of temperature on the compressive strength development of investigated concretes.

10
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

The cross-over effect was significantly delayed by incorporating GGBS. For PC47 concrete, the cross-over effect occurred at 7 days
after casting, at all curing regimes, while for GGBS mixes it only occurred between 14 and 28 days after casting. For all of the mixes
containing GGBS, when cured at 30 ◦ C, the cross-over effect was delayed even later than 56 days after casting, which is a considerable
improvement when compared to the 7-day of PC47.
GGBS concretes cured under adiabatic conditions showed also better improvements in compressive strength at early ages than PC
concretes. Although the rate of compressive strength gain of GGBS concretes was slower than that of PC concretes at early ages
(attributed to their slower rate of adiabatic temperature rise), as long as the nearly-maximum temperature rise was reached, the
compressive strength development was more significant for GGBS concretes. The 28-day compressive strength of adiabatically cured
GGBS concretes was found to vary between 47 and 51 MPa, which was higher than the 43 MPa of PC47 concrete, indicating that the
detrimental effect of temperature on the compressive strength development is delayed with the addition of GGBS. Notably, for
35GGBS47 (also for 20GGBS47), no cross-over effect was observed up to 28 days of adiabatic curing, however, results suggest that it
will occur at later ages. It is also noted that the adiabatic strength development at early ages, e.g., up to 2 days, is similar to that of the
corresponding concretes cured at 30 and 40 ◦ C. Conversely, once the near-maximum temperature was achieved the adiabatic
compressive strength development becomes more comparable with that of the concretes cured at 50 ◦ C.
GGBS has not traditionally been popular in “fast-track” construction and in precast concrete applications, regardless of the need to
reduce the carbon footprint of modern concrete. This is mainly attributed to the strength development of GGBS concretes at room
temperature being slower compared to that of neat Portland cement. For example, for a precast concrete contractor to maintain an
optimised production schedule, sufficient concrete strength should be attained as early as 16 h after casting [18]. As such, in the
present study, the very early-age strength development of GGBS concretes cured at different temperatures was investigated. With
testing commencing as early as 3, 6, 12, 16 and 24 h, it was possible to quantify the effect of elevated curing temperature on GGBS
concrete strength gain at very early ages. 16-hour (0.67-day) compressive strength tests were conducted in order to provide guidance
as to whether the investigated mixes achieve the lifting criterion of 15 MPa, as specified by PD CEN/TR 15728:2016 [79].
It appears that high curing temperatures, 40 and 50 ◦ C were required to obtain a compressive strength of at least 15 MPa at 16 h
(0.67 days) after casting. At 40 ◦ C, the 16-h strength was 16, 15, 15, 11 and 10 MPa and 16, 16, 19, 16 and 16 MPa at 50 ◦ C, for 0, 20,
35, 50 and 70% GGBS replacement levels respectively. This suggests that for the mixes and strength class investigated, GGBS content
should not exceed 35% in precast concrete, unless high temperature curing is applied, which will consequently increase the cost and
embodied carbon of the curing process. A traditional precast PC concrete curing cycle will consist of a “delay” period before the
temperature rise and maximum curing temperature periods [20]. However, this curing regime, may not necessarily be applicable to
GGBS concrete, especially for high replacement levels, due to its low compressive strength development gain rate at low curing
temperatures. In addition to that, high temperature curing can have more beneficial impact on compressive strength at very early ages,
i.e. even as early as immediately after casting, than at several hours after casting, due hydration reaction being more temperature
sensitive at early ages. Relatively higher temperatures could therefore be applied in precast GGBS concrete, compared to PC concrete,
to accelerate its strength development and enable its wider adoption in precast applications. It is also noted that since the detrimental
effect of early-age temperature on later age concrete strength is not as pronounced for GGBS as is it for PC concrete, the aforementioned
“delay” period could be somewhat reduced for GGBS concretes.
The above was quantified in a previous study by the authors [18], and it was demonstrated that it is possible to maximise the GGBS
replacement level in precast concrete without adversely affecting the production schedule and carbon footprint of the concretes.
Higher curing temperatures lead to higher energy demand and consumption; it was, however, suggested that if the GGBS content is
maximised, the embodied carbon saving from cement substitution can potentially outweigh the embodied carbon increase from the

Fig. 7. Relative compressive strengths, i.e. ratio of actual compressive strengths (S) with 20 ◦ C curing compressive strength (S20), for investigated concretes.

11
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

energy required for elevated temperature curing. Alternatively, to implement higher GGBS contents in precast concrete, the mixes
would be designed with a higher target mean strength and thus a higher cementitious content than what is actually required. This is
undesirable for both sustainability-wise and cost-wise.
Elevated curing temperatures have a much more pronounced effect on the compressive strength development of GGBS concretes
than neat Portland cement concrete, and this has also been reported by other researchers from carrying out tests on mortars [24,27,47,
80]. Fig. 7 shows the effect of temperature on the relative strength development of concretes with and without GGBS. The relative
strength is the ratio of compressive strengths at elevated curing temperatures (S) to the compressive strength at standard curing
temperature (S20), i.e. (S/S20), as shown in Fig. 7. The distance between relative strength lines at different temperatures is increased
with GGBS replacement level with the maximum strength enhancement occurring at early ages, as expected. The compressive strength
at 50 ◦ C at 1 day after casting for 20, 35 and 50% replacement levels was between 2.5 and 4.7 times greater than the 20 ◦ C compressive
strength, whilst for 70GGBS47 the compressive strength was as high as 5.5 times greater. The relative strengths were even greater for
high GGBS mixes at 0.5 day; however, the relative strength at that early age may be somewhat unreliable due to low 0.5-day
compressive strengths of mixes cured at standard curing temperature. The relative compressive strengths (S/S20) were gradually
decreased with curing age surpassing the threshold of 20 ◦ C due to the occurrence of the “cross-over” effect. Regardless, Fig. 7 indicates
that GGBS concretes are significantly more temperature sensitive than PC concretes at early and very early ages. Such
temperature-driven strength enhancements should be exploited in several applications in order to promote the use of higher GGBS
containing concretes, e.g., precast, prestressed and post-tensioning applications and/or jump-forming applications.

5.3. Determination of “apparent” activation energies and mix-specific C-values


The determination of “apparent” activation energy becomes necessary as it is required for the Arrhenius equivalent age method. In
the present study, the “apparent” activation energy was determined based on concretes, rather than “equivalent mortars” as suggested
by ASTM C1074-19 [39] due to the previously demonstrated discrepancies in concrete and mortar compressive strength development.
The activation energy was calculated according to both the ASTM C1074-19 (Carino) [39,52] method (Equation (13)) and the
FHP-TPE [53] method (Equation (14)). This required regression analysis at all curing temperatures using Equations (13) and (14) and
resulting regression curves are shown in Fig. 8. The obtained rate constant (k) from Equation (13) and characteristic time constant
values (τ) from Equation (14) were plotted against the reciprocal of absolute temperature (1/Tabs), as shown in Fig. 9, so that the
activation energies could be determined. The regression parameters (obtained through regression analysis using Sigmaplot; a

Fig. 8. Regression lines through experimental data at different temperatures with the ASTM-Carino and FHP-TPE formulations.

12
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

Fig. 9. ln(k) and ln(τ) against the reciprocal absolute temperature.

commercially available statistical analysis software) and “apparent” activation energies obtained from both methods are shown in
Table 5.
As shown in Table 5, increasing the GGBS content results in higher “apparent” activation energies indicating the greater tem­
perature sensitivity of concretes containing GGBS compared to that of neat PC. The obtained values of “apparent” activation energies
from compressive strength data are in good agreement with those reported in the literature [24,27,80–83]. It should also be noted that
the calculated values of the “apparent” activation energies of the mixes based on the two methods are numerically different. Although
the obtained values of the “apparent” activation energy for PC47 was similar from both methods, it appears that the ASTM C1074-19
method yields greater values of “apparent” activation energy than the FHP-TPE method for mixes containing GGBS whilst these
differences may even reach up to 10 kJ/mol. This suggests that the maturity method used to predict concrete strength should be the
same with that used to estimate the “apparent” activation energy of the mix.
The C-value which is used in the (Dutch) Weighted-Maturity method is essentially a factor which accounts for the temperature
sensitivity of a particular concrete mix, similar to the “apparent” activation energy used in the Arrhenius function. Recommended
values in the literature may be found based on type of cement used, e.g. 1.25 for CEM I and CEM II/B–V, 1.60–1.65 for CEM III/B and
1.40 for CEM III/A and CEM V/A [42–44]; however, these may be considered generalised and may not compensate for the various
levels of GGBS used in this study. In order to calculate the C-value individually for every mix, the early-age compressive strength at the
lowest and highest curing temperatures, in this case 20 and 50 ◦ C, is considered. Compressive strength results at these two curing
temperatures should be distributed uniformly over a specified compressive strength range, i.e. approximately 5–35 MPa. The calcu­
lated Weighted Maturity is then plotted on logarithmic scale against the early age compressive strength. By performing linear
regression analysis, the C-value is determined by trial and error until achieving the highest coefficient of correlation [42–44], as
illustrated in Fig. 10. Evidently, the C-value increases with GGBS replacement level as this was estimated to be 1.12, 1.32, 1.41, 1.57
and 1.84 for PC47, 20GGBS47, 35GGBS47, 50GGBS47 and 70GGBS47, respectively. This is attributed to the increased temperature
sensitivity of concretes containing GGBS at different levels, which cannot be accurately accounted from the recommended C-values in
existing literature [42–44].
The effect of temperature on the age conversion factor inherent in the investigated maturity functions is shown in Fig. 11. The
assumption of the Nurse-Saul function that the rate of strength gain varies linearly with temperature seems to be insufficient for
describing the temperature sensitivity of the mixes which particularly contain GGBS. Conversely, the Arrhenius model which takes into
account the temperature sensitivity of a specific mix by considering its “apparent” activation energy (which is the primary advantage
of this function), assumes that there is an exponential relationship between strength development and temperature, resulting in a more
adequate description of the effect of temperature on the strength development of concretes with GGBS. Equally for the age conversion
factor inherent in the Weighted Maturity function, the C-value accounts for the temperature sensitivity of a particular mix and can
reflect the higher temperature sensitivity of GGBS mixes in the age conversion factor. The effect of temperature on the strength
development in Weaver-Sadgrove model becomes more pronounced than the Nurse-Saul but still not enough to account for the

13
F. Kanavaris et al.
Table 5
Regression parameters for strength-time relationships.

ASTM C1074-19 (Carino) (Equation (13)) [39,52]

Mix ID PC47 20GGBS47 35GGBS47 50GGBS47 70GGBS47

Regression parameters Su k t0 Su k t0 Su k t0 Su k t0 Su K t0

Curing 20 58.7 0.225 2.42E-09 63.3 0.139 4.08E-09 65.0 0.119 0.421 64.3 0.084 0.084 61.0 0.080 0.272
Temperature [◦ C] 30 53.0 0.325 0.488 58.8 0.287 3.87E-09 61.8 0.279 0.104 62.1 0.224 0.006 55.7 0.230 0.185
40 46.7 0.560 0.721 53.7 0.551 0.174 57.9 0.498 0.027 55.5 0.404 0.342 48.1 0.482 0.167
14

50 44.0 0.776 0.009 49.6 0.740 0.059 55.8 0.777 0.052 54.6 0.652 0.068 47.9 0.666 0.045
Ea [kJ/mol] 33.4 44.8 49.1 53.3 56.0
FHP-TPE (Equation (14)) [53]
Mix ID PC47 20GGBS47 35GGBS47 50GGBS47 70GGBS47
Regression parameters Su τ a Su τ a Su τ a Su τ a Su Τ a
Curing 20 60.2 2.120 0.7 65.5 2.939 0.7 65.5 3.133 0.7 69.3 4.342 0.7 61.3 5.052 0.7
Temperature [◦ C] 30 55.4 1.394 59.6 1.797 63.1 1.794 65.3 2.653 57.3 2.432
40 48.1 0.880 55.1 0.963 59.5 1.056 57.7 1.344 49.4 1.087
50 45.2 0.636 51.5 0.724 57.2 0.712 55.5 0.882 48.6 0.793
Ea [kJ/mol] 32.1 38.1 39.2 43.0 50.2

Journal of Building Engineering 63 (2023) 105434


F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

Fig. 10. Determination of the C-value inherent in Weighted Maturity for the 47 MPa PC and GGBS concretes.

increased temperature dependence of high GGBS replacement levels, in contradiction with the Rastrup function where its age con­
version factor is increasing considerably with temperature, making it more suitable probably only for mixes with high temperature
sensitivity.

5.4. Concrete strength estimates


Numerous maturity functions currently exist in the literature [32,36–38,52,53,84–89] which are used to estimate the compressive
strength development of concrete under variable temperature conditions. These functions use the experimentally determined
compressive strength-age relationship at a reference temperature (usually at 20 ◦ C) to estimate the strength at any other temperature.
However, the aforementioned were developed for PC concrete. With GGBS being more temperature sensitive than PC concrete, it is
necessary to assess the accuracy of existing maturity functions and strength-maturity/time relationships in estimating GGBS concrete
strength.
The adiabatic temperature history was used to calculate the equivalent age during a time interval, Δt, based on each maturity
function investigated herein, i.e. Nurse-Saul (Equation 2), Rastrup (Equation 5), Weaver-Sadgrove (Equation 7), Arrhenius (Equation
9) and Weighted Maturity (Equation 11) functions. As mentioned earlier, the reference temperature in all cases was 20 ◦ C (293.15 K).
Once the equivalent age, te, from each function was calculated, it was then substituted for t in Equations (13) and (14) with constants
(Su, k, t0 and Su, τ, α for the ASTM-Carino and FHP-TPE methods respectively) as previously determined for the strength results ac­
quired for the concretes cured at 20 ◦ C (Table 5). The values of the “apparent” activation energies, which were required by the
Arrhenius function, were those calculated as explained previously (also shown in Table 5) and were used in the corresponding
strength-maturity/time relationship considered, i.e. “apparent” activation energy values calculated based on the ASTM-Carino and
FHP-TPE method were used to derive maturity estimates using the ASTM-Carino and FHP-TPE formulations respectively.
The adiabatic compressive strength estimates using different maturity functions based on the ASTM-Carino formulation for PC, 20,
35, 50 and 70GGBS47 are shown in Fig. 12, alongside those based on the FHP-TPE formulations. Generally, both ASTM-Carino and
FHP-TPE method, incorporating different maturity functions, predicted the adiabatic compressive strength development rather
inaccurately throughout the investigated curing period, especially at later ages. This is associated with the deficiency of existing
maturity functions in terms of accounting for the detrimental effect of temperature on the later age compressive strength of concrete.
In order to investigate the accuracy of the different models used, the estimated and actual adiabatic compressive strengths are
shown in terms of a ratio, where values below one indicate an underestimation of compressive strength and values above one indicate
its overestimation. These are illustrated at Fig. 13 for estimates using the ASTM-Carino method and the FHP-TPE method.
For the neat Portland cement concrete using the ASTM-Carino correlations, the maturity estimates were somewhat inaccurate. At
an early age, i.e. 12 h after casting, only the Arrhenius and Weaver-Sadgrove functions produced accurate estimates. These functions,
however, overestimated the actual adiabatic compressive strength, together with the Rastrup function which considerably over­
estimated the adiabatic compressive strength throughout the investigated curing period. The Nurse-Saul function was the one that
estimated the adiabatic compressive strength more accurately throughout the investigated curing period due not only to its lower age
conversion factor but also to the fact that the adiabatic compressive strength development of PC47 was close to that at standard curing
temperature. Predictions of similar accuracy were observed when the FHP-TPE formulation was used, where the accuracy of the
Rastrup, Arrhenius, Sadgrove and Nurse-Saul functions was slightly decreased.
For GGBS concretes, the accuracy of the adiabatic compressive strength estimates improved for the maturity functions with higher
age conversion factors but not to a reliable extend. More specifically, the Nurse-Saul function using both the ASTM-Carino and the
FHP-TPE formulations underestimated the early age compressive strength which was then followed by a slight overestimation of the
strength at later ages. The Nurse-Saul function does not sufficiently account for the increased temperature sensitivity of concretes with
GGBS, hence, the early age underestimations. There were not significant differences between the two compressive strength-maturity

15
F. Kanavaris et al.
16

Journal of Building Engineering 63 (2023) 105434


Fig. 11. The effect of temperature on the age conversion factors inherent in different maturity functions.
F. Kanavaris et al.
17

Journal of Building Engineering 63 (2023) 105434


Fig. 12. Adiabatic compressive strength estimates for investigated concretes based on the ASTM-Carino the FHP-TPE formulations.
F. Kanavaris et al.
18

Journal of Building Engineering 63 (2023) 105434


Fig. 13. Ratios of estimated and actual strengths based on the Carino-ASTM and FHP-TPE formulation.
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

correlations used, other than that the FHP-TPE relationship resulted in slightly more accurate estimates.
The Weaver-Sadgrove function produced more accurate estimates for GGBS concretes than for PC47, especially for 20GGBS47. This
is attributed to the fact that it has a higher age conversion factor than the Nurse-Saul function. However, it still prodiced relatively
unreliable early age estimates for concretes with higher replacement levels of PC with GGBS, i.e. 35, 50 and 70%, where the adiabatic
compressive strength of these mixes was underestimated and was followed by an overestimation at later ages. Generally, whenever the
FHP-TPE formulation was used the accuracy of the Weaver-Sadgrove estimates slightly increased, especially for 35GGBS47.
With regards to the Arrhenius function estimates, these did not improve to a significant degree than the previously mentioned
functions. The Arrhenius function, which is usually considered as the most accurate maturity function due to the incorporation of the
“apparent” activation energy of a mix, generally overestimated the adiabatic compressive strength development of GGBS concretes,
especially from 1 day onwards. The adiabatic compressive strength was even overestimated at early ages for the 20GGBS47 mix, which
is probably related to the detrimental effect of temperature occurring at an early age. Estimates of similar accuracy were calculated
with both the ASTM-Carino and FHP-TPE compressive strength-time relationships when the Arrhenius equivalent age was used.
Finally, the Rastrup function could estimate the adiabatic compressive strength development more accurately for GGBS concretes
than for the neat Portland cement concrete, due to the increased temperature sensitivity of concretes containing GGBS. However, the
compressive strengths were still overestimated for all mixes, and were very similar to the Arrhenius function estimates, especially for
35, 50 and 70GGBS47. Only for 35 and 70GGBS47 at early ages, i.e. up to 2–3 days after casting, the estimates were of acceptable
accuracy (<15%), whenever the ASTM-Carino formulation was used.
Overall, none of the existing maturity functions considered herein was capable of reliably predicting the adiabatic compressive
strength development of concretes with and without GGBS, stressing even more the need for the development of a reliable model for
early age and long-term strength development. The existing and also investigated maturity functions at their present state should not
be incautiously used as a decision tool to perform critical operations on site for high GGBS concrete, especially at early ages, unless a
conservative approach is adopted, i.e. consider a high factor of safety and error margin for the estimate. This would arguably result in
overdesigning reinforced concrete structures and inefficient construction/production time management. However, there is great
potential of using maturity functions to optimise curing regimes and exploit the beneficial effects of temperature on strength devel­
opment of GGBS concrete. This will allow maximisation of GGBS in concrete mixes, which is favourable for reducing their embodied
carbon, without adversely impacting the construction schedule. Consequently, modification in existing maturity functions should be
proposed in order to improve their reliability/accuracy towards predicting the strength development of GGBS containing concrete.
Further work is currently ongoing on addressing the aforementioned deficiencies in maturity functions and enable an accurate in-situ
strength prediction for concrete with and without CRMs and particularly GGBS [70,90].

6. Conclusions
In this study the effect of temperature on the compressive strength development of concretes containing GGBS was investigated.
The conclusions drawn are:
• There are some discrepancies between concrete and “equivalent” mortar compressive strength development and strength gain rate
as the mortar mixes underestimated both. This potentially indicates that “equivalent” mortars are unsuitable in replicating concrete
properties and “apparent” activation energy should be preferably calculated based on concrete samples.
• High replacement levels of GGBS are required to reduce the adiabatic temperature rise of concrete (i.e. 50 and 70%).
• While the compressive strength development of GGBS concrete is slow at standard curing temperature, it is significantly accel­
erated with elevated temperature curing. Applying elevated temperature curing soon after casting may potentially enable the use of
GGBS concrete in applications where high early-age strength is required, i.e. in precast concrete applications.
• The values of “apparent” activation energies increased with GGBS content, but also depend on their determination method, i.e. the
ASTM method resulted in higher “apparent” activation energies than the FHP-TPE method.
• Existing maturity functions, although they encompass limitations, can be used to estimate the compressive strength development of
concrete. The Nurse-Saul function, produced the most accurate estimates for neat PC concrete; however, this was not also the case
for the mixes containing GGBS which are more temperature sensitive. The Arrhenius and Weighted Maturity functions resulted in
improved estimates at early ages as they consider the temperature sensitivity of a mix. The Weaver-Sadgrove model is slightly more
accurate than the Nurse-Saul one, while the Rastrup function mostly overestimated the development of strength. It is important to
mention that none of the functions consider the detrimental effect of temperature on the strength development resulting in
overestimations of strength, especially at later ages. Generally, none of the examined functions produced satisfactory and
consistently reliable estimates, where the FHP-TPE correlations resulted in slightly more accurate results than the ASTM-Carino
ones.

CRediT authorship contribution statement


Fragkoulis Kanavaris: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Writing –
original draft. Marios Soutsos: Conceptualization, Methodology, Writing – review & editing, Project administration, Resources,
Supervision. Jian-Fei Chen: Supervision, Writing – review & editing, Project administration.

19
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements
The authors are grateful to the Department of Employment and Learning for the financial support received, as well as to the School
of Natural and Built Environment, Queen’s University Belfast, for the facilities provided. The contents of this paper reflect the views of
the authors, who are responsible for the validity and accuracy of presented data, and do not necessarily reflect the views of their
affiliated organisations.

References
[1] ACI (American Concrete Institute), ACI 233R-03: Slag Cement in Concrete and Mortar, ACI, Farmington Hills, MI, USA, 2003. Report by ACI Committee 233.
[2] M.L. Berndt, Influence of concrete mix design on CO2 emissions for large wind turbine foundations, Renew. Energy 83 (2015) 608–614.
[3] J. Bijen, Blast-furnace Slag for Durable Marine Structures, Stichting Beton Prisma, Den Bosch, Netherlands, 1996.
[4] K. Golden, S. Wong, Sustainable construction: mount lucas windfarm, Concrete Magazine 50 (No. 1) (2016) 33–36.
[5] T.C. Ling, S.C. Kou, C.S. Poon, Precast architectural tiles produced by double-layer casting method, Cement Concr. Compos. 66 (2016) 73–81.
[6] J.P. Won, H.H. Kim, S.J. Lee, S.J. Choi, Carbon reduction of precast concrete under marine environment, Construct. Build. Mater. 74 (2015) 118–123.
[7] V. Pachta, E.K. Anastasiou, Utilisation of industrial byproducts for enhancing the properties of cement mortars at elevated temperatures, Sustainability 13
(2021).
[8] B. Lothenbach, K. Scrivener, R.D. Hooton, Supplementary cementitious materials, Cement Concr. Res. 41 (No. 12) (2011) 1244–1256.
[9] J.M. Crow, The Concrete Conundrum, Chemistry World, 2008, pp. 62–66.
[10] NRMCA (National Ready Mixed Concrete Association), “Concrete CO2 Fact Sheet”, Publication Number 2PCO2, USA.
[11] J. Ahmad, K. Kontoleon, A. Majdi, et al., A comprehensive review on the ground granulated blast furnace slag (GGBS) in concrete production, Sustainability 14
(14) (2022).
[12] X. Zhou, J.R. Slater, S.E. Wavell, O. Oladiran, Effects of PFA and GGBS on early-ages engineering properties of Portland cement systems, J. Adv. Concr. Technol.
10 (2) (2012) 74–85.
[13] UN Environment, K.L. Scrivener, M.J. Vanderley, E.M. Gartner, Eco-efficient cements: potential economically viable solutions for a low-CO2 cement-based
materials industry, Cement Concr. Res. 114 (2018) 2–26.
[14] F. Kanavaris, N. Ferreira, B. Marsh, Low carbon concrete: specification and practical recommendations, Concrete 55 (Issue 2) (2021) 35–37.
[15] J. Parker, Marginal Gains – Carbon in Concrete Buildings, The Structural Engineer, 2021, pp. 20–21.
[16] The Green Construction Board, Low Carbon Concrete Routemap, The Institution of Civil Engineers, 2022.
[17] Formwork, A Guide to Good Practice, third ed., Report of a Concrete Society Working Party, 2012.
[18] F. Kanavaris, M. Soutsos, J.F. Chen, S.V. Nanukuttan, An investigation on appropriate curing regimes for precast concrete structural elements with GGBS using
maturity functions, in: International RILEM/COST Conference on Interdisciplinary Approaches for Cement-Based Materials and Structures - SynerCrete 18,
2018. At: Funchal, Madeira, Portugal.
[19] T. Harrison, Concrete properties: setting and hardening, in: J. Newman, B.S. Choo (Eds.), Advanced Concrete Technology – Concrete Properties, Butterworth-
Heinemann/Elsevier, Oxford, UK, 2003, pp. 4/1–4/33.
[20] A.M. Neville, J.J. Brooks, Concrete Technology, second ed., Pearson Education, 2010.
[21] M. Soutsos, A. Hatzitheodorou, J. Kwasny, F. Kanavaris, Effect of in-situ temperature on the early-age strength development of concrete with supplementary
cementitious materials, Construct. Build. Mater. 103 (2016) 105–116.
[22] S.J. Barnett, M.N. Soutsos, J.H. Bungey, S.G. Millard, Fast-track construction with slag cement concrete: adiabatic strength development and strength prediction,
ACI Mater. J. 104 (4) (2007) 388–396.
[23] M. Soutsos, A. Hatzitheodorou, F. Kanavaris, J. Kwasny, Compressive strength estimates of adiabatically cured concretes using maturity methods, J. Mater. Civ.
Eng. 31 (7) (2019).
[24] A.G. Brooks, A.K. Schindler, R.W. Barnes, Maturity method evaluated for various cementitious materials, J. Mater. Civ. Eng. 19 (12) (2007) 1017–1025.
[25] M. Soutsos, A. Vollpracht, F. Kanavaris, Applicability of fib model code’s maturity function for estimating the strength development of GGBS concretes,
Construct. Build. Mater. 264 (2020), 120157.
[26] A. Vollpracht, M. Soutsos, F. Kanavaris, Strength development of GGBS and fly ash concretes and applicability of fib model code’s maturity function – a critical
review, Construct. Build. Mater. 162 (2018) 830–846.
[27] M. Soutsos, A. Hatzitheodorou, F. Kanavaris, J. Kwasny, Effect of temperature on the strength development of mortar mixes with GGBS and fly ash, Mag. Concr.
Res. 69 (15) (2017) 787–801.
[28] H. Lew, S. Fattel, J. Shaver, T. Reinhold, B. Hunt, Investigation of Construction Failure of Reinforced Concrete Cooling Tower at Willow Island, National
Engineering Lab (NBS), Washington, 1979. West Virginia Final Report (No. PB-80-192883).
[29] D.V. Kaminetzy, P.C. Stivaros, Early-age concrete: construction loads, behaviour and failures, Concr. Int. 16 (1) (1994) 58–63.
[30] J. Feld, K. Carper, Construction Failure, John Wiley and Sons, New York, 1997.
[31] N.J. Carino, The maturity method, in: V.M. Malhotra, N.J. Carino (Eds.), Handbook on Nondestructive Testing of Concrete, second ed., CRC Press, 2004,
5.1–5.47.
[32] A.G.A. Saul, Principles underlying the steam curing of concrete at atmospheric pressure, Mag. Concr. Res. 2 (No. 6) (1951) 127–140.
[33] M.L. Gambhir, Concrete Technology, fifth ed., McGraw Hill Education, New Delhi, India, 2013.
[34] V.M. Malhotra, Nondestructive tests, in: J.F. Lamond, J.H. Pielert (Eds.), Significance of Tests and Properties of Concrete and Concrete-Making Materials, ASTM
International, West Conshohocken, PA, USA, 2006, pp. 314–334.
[35] N. Han, Maturity method, in: H.W. Reinhardt, C.U. Grosse (Eds.), Advanced Testing of Cement-Based Materials during Setting and Hardening, 2005,
pp. 277–296. RILEM Publications, Bagneux, France, Final report of RILEM TC 185-ATC.
[36] E. Rastrup, Heat of hydration in concrete, Mag. Concr. Res. 6 (17) (1954) 79–92.
[37] J. Weaver, B.M. Sadgrove, Striking Times of Formwork-Tables of Curing Periods to Achieve Given Strength”, Construction Industry and Information Association,
October, 1971. Report 36.
[38] Hansen, F., P. and Pedersen, E., J., “Maturity computer for controlled curing and hardening of concrete”, Nord Betong, Vol. 1, p. 19.
[39] ASTM International, ASTM C1074-19 Standard Practice for Estimating Concrete Strength by the Maturity Method, ASTM International, West Conshohocken,
Pennsylvania, USA, 2019.

20
F. Kanavaris et al. Journal of Building Engineering 63 (2023) 105434

[40] A. Hatzitheodorou, In-situ Strength Development of Concretes with Supplementary Cementitious Materials, PhD thesis, The University of Liverpool, 2007.
[41] G. Turu’allo, Early Age Strength Development of GGBS Concrete Cured under Different Temperatures, PhD Thesis, The University of Liverpool, 2013.
[42] F. Kanavaris, Early Age Behaviour and Cracking Risk of Concretes Containing GGBS, PhD Thesis, Queen’s University of Belfast, 2017.
[43] R. De Vree, T, R. Tegelaar, Gewichtete reife des betons, beton 48” (11) (1998) 674–678.
[44] CUR, CUR-recommendation: Determining the Strength of Young Concrete on the Basis of Weighted Maturity, The Dutch centre for civil engineering, codes and
specifications, Gouda, the Netherlands, 1986.
[45] M. Papadakis, J. Bresson, Contribution to the study of the maturity of hydraulic binders. Application to the prefabricated concrete industry, Revue Mater.
Constr. & Trav. Pub. “Ciments & Betons” (678) (1973) 18–22.
[46] M. Soutsos, F. Kanavaris, A. Hatzitheodorou, Critical analysis of strength estimates from maturity functions, Case Stud. Constr. Mater. 9 (2018), e00183.
[47] T. Boubekeur, K. Ezziane, E. Kadri, Estimation of mortars compressive strength at different curing temperature by the maturity method, Construct. Build. Mater.
71 (2014) 299–307.
[48] I. Galobardes, S. Cavalaro, C.I. Goodier, S. Austin, A. Rueda, Maturity method to predict the evolution of the properties of sprayed concrete, Construct. Build.
Mater. 79 (2015) 357–369.
[49] M. Sofi, P. Mendis, A, D. Baweja, Estimating early-age in situ strength development of concrete slabs, Construct. Build. Mater. 29 (2012) 659–666.
[50] T. Yikici, A, H. Chen, Use of maturity method to estimate compressive strength of mass concrete, Construct. Build. Mater. 95 (2015) 802–812.
[51] M.N. Soutsos, G. Turu’allo, K. Owens, J. Kwasny, S. Barnett, J, P. Basheer, M. A, Maturity testing of lightweight self-compacting and vibrated concretes,
Construct. Build. Mater. 47 (2013) 118–125.
[52] N.J. Carino, R.C. Tank, Maturity functions for concretes made with various cements and admixtures, ACI Mater. J. 89 (No. 2) (1992) 188–196.
[53] F.P. Hansen, E.J. Pedersen, Curing of concrete structures, CEB Information Bulletin 166 (May) (1985).
[54] M.N. Soutsos, The effect of temperature on the rate of strength development of mortar mixes, in: M.C. Forde (Ed.), Structural Faults & Repair – 2010, 13th
International Conference, 2010, p. 13. Engineering Technics Press, 46 Cluny Gardens, Edinburgh EH10 6BN, UK, ISBN: 0-947644- 67-9, June.
[55] J.M. Plowman, Maturity and the strength of concrete, Mag. Concr. Res. 8 (22) (1956) 13–22.
[56] G. Cheng-ju, Maturity of concrete: method for predicting early-stage strength, ACI Mater. J. 86 (4) (1989) 341–353.
[57] NEN 5970: Determination of Strength of Fresh Concrete with the Method of Weighted Maturity, 2001.
[58] CICR/04/20 G. Ho, F. Kanavaris, N. Yiu, L. Caine, N. Ferreira, S. Ip, Reference Materials on Maturity Method for Estimation of Concrete Strength – Practical
Guideline, Hong Kong Construction Industry Council, 2022 [In press].
[59] IFFSTAR, Guide technique: Resistance du beton dans l’ouvrage – La maturometrie, Techniques et methodes des laboratoires des points et chaussees, Mars 2003
[in French].
[60] ACI (American Concrete Institute), ACI 207.1R-05: Guide to Mass Concrete”, Farmington Hills, MI, USA, 2005. Report by ACI Committee 207, ACI.
[61] D.C. Teychenne, R.E. Franklin, H.C. Erntroy, Design of Normal Concrete Mixes”, second ed., BRE Press, 1997.
[62] G.K. Munkelt, The Durability Factor, Precast Solutions Magazine, National Precast Concrete Association, Carmel, Indiana, USA, 2010.
[63] W. Xue, X. Yang, Seismic tests of precast concrete moment-resisting frames and connections, PCI J. 55 (3) (2010) 102–121.
[64] BSI (British Standards Institution), BS EN 197-1:2011 – Cement. Composition, Specifications and Conformity Criteria for Common Cements, BSI, London, UK,
2011.
[65] BSI (British Standards Institution), BS EN 196-1:2016 – Methods of Testing Cement. Determination of Strength, BSI, London, UK, 2016.
[66] BSI (British Standards Institution), BS EN 15167-1:2006 – Ground Granulated Blast Furnace Slag for Use in Concrete, Mortar and Grout. Definitions,
Specifications and Conformity Criteria, BSI, London, UK, 2006.
[67] BSI (British Standards Institution), BS EN 12620:2002+A1:2008 – Aggregates for Concrete, BSI, London, UK, 2002.
[68] BSI (British Standards Institution), BS EN 12350-2:2009 – Testing Fresh Concrete. Slump-Test, BSI, London, UK, 2009.
[69] BSI (British Standards Institution), BS EN 12390-1:2012 – Testing Hardened Concrete. Shape, Dimensions and Other Requirements for Specimens and Moulds,
BSI, London, UK, 2012.
[70] M. Soutsos, F. Kanavaris, Compressive strength estimates for adiabatically cured concretes with the Modified Nurse-Saul (MNS) maturity function, Construct.
Build. Mater. 255 (2020), 119236.
[71] BSI (British Standards Institution), BS EN 12390-3:2009 – Testing Hardened Concrete. Compressive Strength of Test Specimens, BSI, London, UK, 2009.
[72] A.M. Neville, Properties of Concrete, fifth ed., Pearson Education, Essex, UK, 2011.
[73] M.N. Soutsos, S.J. Barnett, J.H. Bungey, S.G. Millard, Fast track construction with high strength concrete mixes containing ground granulated blast furnace slag,
in: H.G. Russell (Ed.), Proceedings of ACI Seventh International Symposium on High Strength/High Performance Concrete, 2005, pp. 255–270. ACI SP-228, 1.
[74] X. Li, J. Bizzozero, C. Hesse, Impact of C-S-H seeding on hydration and strength of slag blended cement, Cement Concr. Res. 161 (2022).
[75] J.D. McIntosh, The effects of low-temperature curing on the compressive strength of concrete, in: Proceedings of RILEM Symposium on Winter Concreting,
Session BII, Danish Institute for Building Research, Copenhagen, 1956.
[76] E. Gallucci, X. Zhang, K.L. Scrivener, Effect of temperature on the microstructure of calcium silicate hydrate (C-S-H), Cement Concr. Res. 53 (2013) 185–195.
[77] J.L. Escalante-Garcia, J.H. Sharp, The microstructure and mechanical properties of blended cements hydrated at various temperatures, Cement Concr. Res. 31
(No. 5) (2001) 695–702.
[78] K. Ezziane, A. Bougara, A. Kadri, H. Khelafi, E. Kadri, Compressive strength of mortar containing natural pozzolan under various temperatures, Cement Concr.
Compos. 29 (No. 8) (2007) 587–593.
[79] BSI (British Standards Institution), PD CEN/TR 15728:2016 – Design and Use of Insert of Lifting and Handling of Precast Concrete Elements, BSI, London, UK,
2016.
[80] S.J. Barnett, M.N. Soutsos, S.G. Millard, J.H. Bungey, Strength development of mortars containing ground granulated blast-furnace slag: effect of curing
temperature and determination of apparent activation energies, Cement Concr. Res. 36 (No. 3) (2006) 434–440.
[81] J.L. Poole, K.A. Riding, M. Juenger, G. C, K.J. Folliard, A.K. Schindler, Effects of supplementary cementitious materials on apparent activation energy, J. ASTM
Int. (JAI) 7 (No. 9) (2010).
[82] S.J. Barnett, M.N. Soutsos, S.G. Millard, J.H. Bungey, Temperature rise and strength development in laboratory-cast structural elements containing slag, in: V.
M. Malhotra (Ed.), Ninth CANMET/ACI International Conference on Fly Ash, Silica Fume, Slag and Natural Pozzolans in Concrete, ACI SP-242, Warsaw, Poland,
2007, pp. 37–50. ISBN 978-0-87031-241-0.
[83] J.L. Poole, K.A. Riding, K.J. Folliard, M. Juenger, G. C, A.K. Schindler, Methods for calculating apparent activation energy for Portland cement, ACI Mater. J.
104 (No. 1) (2007) 303–311.
[84] K.G. Babu, G. Rao, N. S, Early strength behaviour of fly ash concretes, Cement Concr. Res. 24 (No. 2) (1994) 277–284.
[85] K.O. Kjellsen, R.J. Detwiler, Later-age strength prediction by a modified maturity model, ACI Mater. J. 90 (No. 3) (1993) 220–227.
[86] Y.A. Abdel-Jawad, The maturity method: modification to improve estimation of concrete strength at later ages, Construct. Build. Mater. 20 (2006) 893–900.
[87] G. Chanvillard, L. D’Aloia, Concrete strength estimation at early age: modification of the method of equivalent age, ACI Mater. J. 94 (No. 6) (1997) 520–530.
[88] T.J. Parsons, T.R. Naik, Early age strength determination by maturity, Concr. Int. 7 (No. 2) (1985) 37–43.
[89] W.-C. Liao, B.J. Lee, C.W. Kang, A humidity-adjusted maturity function for the early age strength prediction of concrete, Cement Concr. Compos. 30 (2008)
515–523.
[90] M. Soutsos, F. Kanavaris, The Modified Nurse-Saul (MNS) maturity function for improved strength estimates at elevated curing temperatures, Case Stud. Constr.
Mater. 9 (2018).

21

You might also like