You are on page 1of 10

SPE-180038-MS

Nano-Sized MgO with Engineered Expansive Property for Oil Well Cement
Systems
Narjes Jafariesfad, Department of Petroleum Engineering and Applied Geophysics; Yi Gong and
Mette Rica Geiker, Department of Structural Engineering; Pål Skalle, Department of Petroleum Engineering and
Applied Geophysics, NTNU

Copyright 2016, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Bergen One Day Seminar held in Bergen, Norway, 20 April 2016.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
The failure of cement sheaths to perform as designed in oil wells can result in loss of zonal isolation. Bulk
shrinkage of the cement sheath in the annulus is one of the main causes compromising zonal isolation.
Conventional cement systems without compensation for bulk shrinkage have a high risk of failure during
all phases of well operation. However, when the volume reduction is compensated, without compromising
the mechanical properties, the risk of failure is significantly reduced.
In this study, nano-sized MgO with designed expansive properties has been introduced to the fresh
cement slurry. The expansive properties of nano-MgO were achieved by controlling the preparation
condition. A dilatometer with corrugated molds was used to measure the linear strain of samples cured at
40 °C.
The reactivity of nano-MgO played a main role in controlling the expansion performance at the
required time. The efficiency of nano-MgO with different reactivities for shrinkage compensation of
cement system, cured at 40 °C, was studied. Addition of only 2% nano-MgO with appropriate reactivity
was sufficient to maintain expansion in the cement system. Controlling the expansion performance of the
additive through designing its reactivity is a promising method to limit bulk shrinkage of cement sheaths
in oil wells.
The results presented show that nano-MgO with controlled expansive properties can be used to design
a cement system with short- and long-term zero-bulk shrinkage.

Introduction
Providing complete and effective zonal isolation for the life of the well is the main goal of primary
cementing. Without complete isolation in the wellbore, the well may never reach its full production
potential (Nelson and Guillot 2006).
From a cementing perspective, the factors contributing to loss of zonal isolation can be categorized
according to when a path for fluid migration is created. The path for fluid entry might be created during
the cementing process and/or after the cement placement. The loss-contributing factores originating from
the cementing process can be eliminated by propper casing centralization, adequate mud removal,
2 SPE-180038-MS

complete cement placement, no cement slurry contamination, and adequate cement-formation/cement-


casing bond. Fluid migration after cement placement can be devided into two main periods, short-term and
long-term. Short-term fluid migration occurs before the cement system sets, and long-term fluid migration
happens after setting. The factors influencing the short-term fluid migration also affect longterm fluid
migration due to the already created paths for fluid entry. Factors contributing to fluid migration are
summarized in Table 1. Analyzing the cause of problems involved in long-term zonal isolation, reveals
that it is mainly due to the low tensile properties and volumetric instability. These properties make the
cement sheath unable to withstand cyclic tensile stresses due to temperature and pressure fluctuation, and
shrinkage of cement sheath.

Table 1—Factors contributing to short-term and long-term fluid migration, after (Nelson and Guillot 2006, Ravi, Bosma, and
Gastebled 2002)
Short-term Long-term

● Fluid loss ● Bulk shrinkage (autogeneous deformation)


● Free-fluid (bleeding) ● Strength development of cement system
● Gel strength development ● Thermal/mechanical cycling condition
● Bulk shrinkage (autogenous deformation)
● Cement system permeability

The reaction between water and cement is associated with a reduction in volume (‘chemical shrink-
age’), because the reaction products fill a lesser volume than the reactants. After setting, chemical
shrinkage results in the pore structure being partly emptied. This so-called self-desiccation creates
capillary forces and an under-pressure in the pore water. The pressure is transferred to the solid and results
in external/bulk contraction of the cementitious material called bulk or autogenous shrinkage (Lura,
Jensen, and van Breugel 2003). Bulk shrinkage might lead to debonding and microannulus formation
between the cement sheath and the casing or the formation, and/or tensile cracks and subsequent increased
permeability (Reddy et al. 2009, Ghofrani and Plack 1993).
Conventional cement systems without shrinkage compensation have a high risk of failure during all
phases of the well operations. However, when the volume reduction is compensated without significantly
modifying the mechanical properties, the risk of loss of zonal isolation is significantly reduced. Moreover,
increasing the tensile strength of conventional systems without compensating for shrinkage can provide
some reduction of the risk of failure (McCulloch et al. 2003). Therefore, in addition to mechanical
properties, the cement system shrinkage should be compensated effectively to maintain the desired zonal
isolation.
Several strategies, including internal curing, shrinkage reducing admixtures (SRA), fibers and expan-
sive additives have been developed in order to mitigate shrinkage of concrete and to some extent cement
sheaths in oil wells. Internal curing through addition of well distributed internal water reservoirs results
in reduced shrinkage stresses and strains (Kovler et al. 2007). Internal curing can e.g. be provided by
superabsorbent polymers or hydrogels (Jensen 2013, Jensen and Hansen 2001), partrially saturated
lightweight aggregate (Bentur, Igarashi, and Kovler 2001, Bentz and Weiss 2011) and, as recently
proposed, nanoclays (Mangadlao, Cao, and Advincula 2015). SRA reduces the surface tension of the pore
solution, decreasing the development of capillary stresses and maintaining a higher relative humidity (Sant
et al. 2011, Bentz, Geiker, and Hansen 2001). Addition of fibers to cement-based materials mitigate
undesirable effects of shrinkage (Polat, Demirboga, and Khushefati 2015).
Expansive additives have been applied in oil well cement systems to prevent shrinkage for more than
50 years. The expansion is based on two principal mechanisms: crystal growth, e.g. of MgO, CaO or
sulfoaluminate-type additives; and gas generation, using e.g. Mg, Al or Zn powder. Gas generating
systems are to be carefully stabilized; otherwise, gas bubbles may coalesce and create channels for
SPE-180038-MS 3

formation fluid to flow (Nelson and Guillot 2006). In the crystal growing systems the rate of reaction and
the products’ stability under downhole conditions are the key parameters required to be considered. A
major limitation of sulfoaluminate-type expansive additives is the lack of stability of ettringite above 76
°C.
The reactivity of expansive additive must be controllable to produce the required expansion at the right
time. If crystallization occurs when the cement slurry is still in the fluid phase; the cement system is
unable to sustain the expansive forces. For additieves with very low reactivity, the crystallization pressure
of expansive additive might result in cracking and loss of mechanical properties, so-called unsoundness.
Consequently, the reactivity of the additive must be designed somehow that its expansion occurs mainly
in the late plastic phase and at the beginning of cement hardening phase. In the initial phase of hardening,
sufficient unhydrous cement capable of hydration is still available for mutually recombining the particles
of binding agent which have been driven apart by the crystallization pressure of expansive additive
(Ghofrani and Plack 1993).
Industrially, CaO-type and MgO-type expansive additives are usually manufactured by calcination of
calcium and magnesium carbonates. In contrast to other expansive additives, the reactivity of MgO and
CaO is influenced by the manufacturing process. In comparison with MgO, CaO exhibits a considerably
higher free enthalpy and thus a higher reactivity, even in the dead-burnt state (manufacture at a
temperature above 1600°C) (Ghofrani and Plack 1993). Therefore, CaO-type expansive additives cause
expansion mainly at an early stage (Mo, Deng, and Wang 2012). MgO reactivity has been devided into
three cathegories based on reactivity value, specific surface area (SSA) and agglomeration ratio (AR):
highly reactive, medium reactive and less reactive (Jin and Al-Tabbaa 2013). These categories permit the
use of MgO with appropriate reactivity according to the bottomhole conditions. Recently, there have been
several studies on controlling MgO expansion properties through preparation and curing conditions (Mo,
Deng, and Tang 2010, Mo et al. 2014, del Valle-Zermeno et al. 2012, Li et al. 2010, Tang et al. 2014,
Salomao et al. 2014). However, still there is a lack of knowledge on expansion of MgO as a function of
curing pressure and temperature, which is essential for designing expanding oil well cement systems.
The size of the expansive additieves has a significant impact on the autogenous shrinkage of the cement
system. Nano-MgO appeared more effective in compensating autogeneous shrinkage than micro-sized
MgO (Polat, Demirboga, and Khushefati 2015). As the nano-MgO is synthesized compared to micro-
MgO which are typically produced through calcination of MgO sources, there is a high potential to control
the reactivity of nano-MgO via sysnthesis process and later in heat treatment process (Sundrarajan,
Suresh, and Gandhi 2012). Therefore, the reactivity of nano-MgO is more designable to provide the
effective expansion at well condition compared to micro-MgO.Nano-MgO is a promising additive for
preserving mechanical properties of the cement system (Moradpour et al. 2013), as well as it can be
designed with appropriate reactivity to prevent shrinkage of cement sheath under downhole condition.
The current research focused on the use of nano-MgO for shrinkage compensation. Three versions of
nano-MgO with different reactivities were prepared by pre-treatment at various temperatures.

Experimental

Materials
Cement pastes used in this study were prepared from API Class G oil well cement, deionized distilled
water and selected additives. The pastes’ composition and the chemical properties of cement and nano-
MgO are described in Table 1 and 2 respectively. Morphology of untreated MgO particles was ellipsoidal
and spherical with an average size of 20 nm and specific surface area of 60 m2/g.
4 SPE-180038-MS

Table 2—Cement paste composition (amounts in g)


Nano-MgO
Bentonite Defoamer Dispersant
Id. Cement water (0.5% BWOC) (0.1% BWOC) (3% BWOC) PEG* NM-A NM-B NM-C

Reference 761.20 336.60 3.81 0.76 22.83 - - - -


S-A 743.00 335.10 3.72 0.74 22.29 3.72 14.86 - -
S-B 743.00 335.10 3.72 0.74 22.29 3.72 - 14.86 -
S-C 743.00 335.10 3.72 0.74 22.29 3.72 - - 14.86
S-A-B 743.00 335.10 3.72 0.74 22.29 3.72 7.43 7.43 -

*Poly ethylene glycole

Nano-MgO (NM) was pre-treated at different temperatures to regulate its reactivity A titration method
was used to characterize the reactivity of NM, as described by (Mo, Deng, and Tang 2010). The reactivity
of pre-treated NMs was as follow: NM-A⬎ NM-B⬎ NM-C. Slurries named S-A, S-B and S-C contained
2% BWOC NM-A, NM-B and NM-C, respectively. S-A-B contained a combination of 1% BWOC NM-A
and 1% BWOC NM-B.

Table 3—The chemical composition (wt.%) of API Class G oil well cement and nano-MgO
Al O
Compositions(wt.%) SiO2 2 3 Fe2O3 CaO MgO SO3 P2O5 K2O Na2O LOI Density (g/cm2)

Cement 21.67 4.11 5.24 64.32 0.65 2.77 0.08 0.65 0.13 1.48 3.15
Nano-MgO - - - - ⱖ 99 - - - - 6.00* 3.58

*Loss on ignition (850 °C/2h)

All cement pastes had a water-to-solid ratio by mass (w/s) of 0.44 and were prepared according to API
Recommended Practice RP 10-B (API-RP-10B-2 2012). For the systems containing NM, cold water was
used to diminish the temperature rising due to NMs dispersing process in water. Bentonite was mixed with
water and left to prehydrate for 30 min before cement addition to enhance its efficiency to maintain a
homogenous slurry and prevent the development of excessive free water (i.e. limit bleeding) (Nelson and
Guillot 2006). First, bentonite was mixed with water at low speed for 15 sec, followed by 15 sec at high
speed, then left 15 min for prehydration. To prepare a well-dispersed suspension of NMs, poly ethylene
glycole (PEG) was added to the mixture (Moradpour et al. 2013), stirred and then, NM was added and
mixed for 15 sec at low speed, followed by 5 min mixing at high speed. At this stage, the suspension
contained lots of air bubbles; therefore, defoamer was added to the suspension, and then was treated with
ultra sound for 5 min. Then, the suspension was transferred to the blender and again mixed 1 min at high
rate; up to this time, it took about 30 min from the time bentonite mixed with water. Subsequently,
dispersant was added to the suspension which was mixed with a spatula. Then, cement was added to the
suspension during 15 sec while mixing at 4000 rpm, followed by 35 sec mixing at 12000 rpm.
For free-fluid and autogeneous shrinkage measurements, the prepared paste was preconditioned at
27 °C at a stirring rate of 150 rpm for 20 min.
Bulk shrinkage (autogeneous shrinkage) measurement
Linear autogenous strain of the cement pastes was measured in a technique in which the fresh paste was
placed in watertight corrugated polyethylene molds with a length-to-diameter ratio of approximately
420:30 mm according to ASTM C 1698-09 which is a modified version of teqnique proposed by Jensen
and Hansen (1995). The technique is designed to encapsulate the fresh cement paste while minimizing
restraint, thus permitting measurements to start 30 min after water addition. The corrugated mold
transforms volumetric deformation into linear deformation when the paste is in a liquid state because the
SPE-180038-MS 5

mold has a greater stiffness in the radial direction than it does in the longitudinal direction. The corrugated
molds were specially designed to minimize restraint on the paste (Jensen and Hansen 1995).
The corrugated tube was filled in four equal layers and each layer was compacted 5 times with a
tamping rod before each subsequent layer was cast. The specimen was then placed in a dilatometer kept
in an oven at 40°C. The dilatometer was equipped with automatic data-logging and electronic linear
displacement transducers (Fig. 1); the measuring accuracy was ⫾ 5 ␮m/m (Sant, Lura, and Weiss 2006).
The measurements were performed at intervals of 15 min and started at about 50⫾10 min after water
addition. They were later referenced to zero-time wich was obtained from the curve of rate of strain vs.
time, further details in the result section.

Figure 1—Bulk shrinkage (autogeneous shrinkage) testing set up based on ASTM C 1698-09

Free-fluid (bleeding) measurement


In order to verify free-fluid content, 250 ml of precondisiond paste was placed in a conical flask, sealed
and left in a place without any vibration for 2 h. After 2 h, the volume of developed supernatant fluid was
measured by an accuracy of ⫾0.1 ml and it was recorded as milliliters free-fluid (API-RP-10B-2 2012).
Later the milliliters free-fluid was converted to a percentage of the initial paste volume representing the
value as percent free-fluid, utilizing the following formula:
(1)

Where ␸ is the volume fraction of free-fluid, expressed as a percent; VFF is the volume of free-fluid
(supernatant fluid) collected (ml); and Vi is the initial volume of the paste (ml).
6 SPE-180038-MS

Result and discussion


Time-zero determination
Evolution of the autogenous deformations has to be expressed from the time-zero defined as the time when
stresses develop inside the cement paste. Several methods based on experimental and modeling techniques
to identify the time of solidification and the extent of structure development in cement based materials are
described (Sant et al. 2009). Determination of time-zero using the rate of autogenous shrinkage devel-
opment over time has been proposed to be a reliable approach for determination of time- zero (Meddah
and Tagnit-Hamou 2010, Darquennes et al. 2011, Sant 2007).
The rate of deformation is the first derivative of deformation function with respect to time . The
rate of deformation as a function of time for cement paste containing 2% NM-C is illustrated in Fig. 2.
Rate of deformation evolution has been divided into three main phases: plastic phase (Z1), transition phase
(Z2, Z3 and Z4) and hardening phase (Z5) (Meddah and Tagnit-Hamou 2010). In the plastic phase, Z1,
cement paste is unable to sustain the stresses being built up, and autogeneous shrinkage equals the
chemical shrinkage development. In the second zone, Z2, the rate of deformation starts to increase
considerably providing evidence of accelerated hydration reaction and start of transition phase. In Z2, the
cement phase could be considered as multiphase material (liquid phase ⫹ solid phase), and the rate of
deformation indicates a significant increase until the peak value is reached. It could be expected that the
initial and final setting time occur during this zone. The peak time indicated with the red line in Fig. 2 was
determined as time-zero which is representative of setting time. Z3 indicates a sharp and linear decrease,
which corresponds to the progress of the cement stiffening process (Meddah and Tagnit- Hamou 2010).
Z4 presents the zone with positive rate of deformation which corresponds to a period of expansion. In Z5,
the rate of deformation tends to stabilize which presents the hardening phase of the cement paste.

Figure 2—Evolution of rate of deformation as a function of time for cement paste containing 2% NM-C, see text for explanation of zones
(Z1-Z5)

Described method was used to determine time-zero (setting time) for all the samples. The time-zero for
the reference system, S-A, S-B, S-C and S-A-B samples was determined to 11.8, 6.7, 8.7, 9.6 and 7.5 h,
respectively.
SPE-180038-MS 7

Pre-treatment condition of NM has a remarkable effect on time-zero. Addition of NM to the cement


paste, regardless of its reactivity, decreased the plastic phase length. The system containing NM-A
revealed a significant reduction in Z1 length compared to the system containing NM-C. The rate of
deformation evolution would be more similar to the reference system for the cement system containing
NM-C. Addition of NM particles to the cement paste reduced the setting time. These results approve that
incorporation of NMs increases the hydration rate, and reactivity of NMs has a noticable impact on their
accelerating role in cement hydration. In contrast to nano-MgO, it has been observed that addition of
micro-MgO to the cement paste led to increased setting time (Zheng, Xuehua, and Mingshu 1992).
Bulk shrinkage
Fig. 3 illustrates the development of autogeneous shrinkage of the reference system and systems
containing 2% of NMs; each system is referenced to its time-zero. Addition of NMs reduced the shrinkage
strains compared to the reference system. The extent of shrinkage reduction depends on the NMs
pre-treatment. In Table 4, the percentage of shrinkage reduction after 1, 3, 7 and 14 days is presented for
the systems containing NMs. sample S-C showed the minimum reduction in shrinkage, but it started
displaying a slow monotonic expansion after 28 days of curing. 28 days of curing is an ample period for
the cement matrix to gain strength. Therefore, expansion forces at this time can be detrimental to the
cement paste.

Figure 3—Autogeneous shrinkage result referenced to time-zero for reference system and cement pastes containing NMs

Table 4 —Percentage of shrinkage reduction in systems


containing NMs compared to the reference system
% shrinkage reduction

1d 3d 7d 14 d

S-A 200 152 137 122


S-B 34 48 47 26
S-C 0 27 14 8
S-A-B 43 56 56 24

NM-A was the most effective additive to compensate shrinkage at 40°C, and preserved the positive
strain up to three weeks. At higher curing temperature, NM with lower reactivity might be required and
8 SPE-180038-MS

at lower curing temperatures NM with higher reactivity can be more efficient to maintain the desired
expansion.
In addition to designable expansion, it has been reported that mechanical performance of cement
system containing NMs are improved. According the ASTM standards for soundness evaluation of
cement, which limit the autoclave expansion up to 0.8 % (Helmuth and West 1998), the cementitious
system contained different amount of MgO nanoparticles, from 1 % to 5 % by the weight of binder
showed sound expansion. Moreover, the compressive and flexural strengths improved in cement mortars
containing NMs (Moradpour et al. 2013). Flextural and compressive strength increased significantly at
low concentrations of NMs (⬍ 3%). At high concentrations of NMs (ⱖ 3%), the strength reduced
compared to the systems contained lower content of NMs; however, the strengths were still higher than
the reference system. It can be concluded that addition of nanoscale expansive agent has a noticeable
impact on mechanical properties of cement composite resulting in a system much less prone to experience
unsoundness compared to cement system containing microsized (Gao et al. 2007) expansive agents.
Work is uner progress to determine the effect of NMs with different reactivities on tensile strength,
compressive strength and Young’s modulus.

Free-Fluid
The amount of free-fluid (bleeding) was measured to characterize the stability of the pastes. Addition of
NM decreased the free-fluid content; the results are presented in Table 5. NMs with higher reactivity
demonstrated lower free-fluid content. The sample NM-A displayed the lowest free-fluid, about 70%
reduction in free-fluid content compared to the reference system.

Table 5—Free-fluid content (bleeding ␸) of the reference paste


and the pastes containing NMs
Sample ␸ (%)

Refrence system 2.2


S-A 0.7
S-B 1.8
S-C 2.0
S-A-B 1.1

Conclusion
Nano-MgO (NM) was used for shrinkage compensation of Class G oil well cement paste. Three versions
of nano-MgO with different reactivities (NM-A, NM-B and NM-C) were prepared by pre-treatment at
various temperatures.
● The development of expansive properties of NM could be controled through heat treatment.
● High reactivity NM provided early expansion in the cement system, eliminated shrinkage for three
weeks.
● Incorporation of NM reduced the setting time of the cement systems. The setting time of system
contained high reactivity NM decreased approximately 40% compared to the reference system.
● The systems containing NMs with high reactivity exhibited lower free-fluid compared to the
reference system.
● Use of hybrid additives with various reactivities might be effective to compensate short- and
long-term bulk shrinkage.
SPE-180038-MS 9

Acknowledgements
The authors are very grateful for the financial support by The Research Council of Norway (project
number 10412401).

References
API-RP-10B-2. 2012. Recommended Practice for Testing Well Cements. API Publishing Services.
Bentur, Arnon, Shin-ichi Igarashi, and Konstantin Kovler. 2001. ⬙Prevention of autogenous shrinkage in high-strength
concrete by internal curing using wet lightweight aggregates.⬙ Cement and concrete research 31 (11): 1587–1591.
Bentz, D. P., M. R. Geiker, and K. K. Hansen. 2001. ⬙Shrinkage-reducing admixtures and early-age desiccation in cement
pastes and mortars.⬙ Cement and Concrete Research 31 (7): 1075–1085. doi: http://dx.doi.org/10.1016/S0008-
8846(01)00519-1.
Bentz, Dale P, and W Jason Weiss. 2011. Internal curing: a 2010 state-of-the-art review: US Department of Commerce,
National Institute of Standards and Technology.
Darquennes, Aveline, Stephanie Staquet, Marie-Paule Delplancke-Ogletree, and Bernard Espion. 2011. ⬙Effect of autog-
enous deformation on the cracking risk of slag cement concretes.⬙ Cement and Concrete Composites 33 (3): 368 –379.
doi: http://dx.doi.org/10.1016/j.cemconcomp.2010.12.003.
del Valle-Zermeno, R, JM Chimenos, J Formosa, and AI Fernandez. 2012. ⬙Hydration of a low-grade magnesium oxide.
Lab-scale study.⬙ Journal of Chemical Technology and Biotechnology 87 (12): 1702–1708.
Gao, Pei-wei, Sheng-xing Wu, Xiao-lin Lu, Min Deng, Ping-hua Lin, Zhong-ru Wu, and Ming-shu Tang. 2007.
⬙Soundness evaluation of concrete with MgO.⬙ Construction and Building Materials 21 (1): 132–138. doi: http://
dx.doi.org/10.1016/j.conbuildmat.2005.06.033.
Ghofrani, Reza, and Heiko Plack. 1993. ⬙CaO- and/or MgO-Swelling Cements: A Key for Providing a Better Annular
Sealing?⬙ SPE/IADC Drilling Conference, Amsterdam, 1993/1/1/.
Helmuth, R, and Presbury B West. 1998. ⬙Reappraisal of the autoclave expansion test.⬙ Cement, concrete and aggregates
20 (1): 194 –219.
Jensen, O Mejlhede. 2013. ⬙Use of superabsorbent polymers in concrete.⬙ Concrete international 35 (1): 48 –52.
Jensen, O Mejlhede, and P Freiesleben Hansen. 1995. ⬙A dilatometer for measuring autogenous deformation in hardening
Portland cement paste.⬙ Materials and Structures 28 (7): 406 –409.
Jensen, Ole Mejlhede, and Per Freiesleben Hansen. 2001. ⬙Water-entrained cement-based materials: I. Principles and
theoretical background.⬙ Cement and Concrete Research 31 (4): 647–654. doi: http://dx.doi.org/10.1016/S0008-
8846(01)00463-X.
Jin, Fei, and Abir Al-Tabbaa. 2013. ⬙Characterisation of different commercial reactive magnesia.⬙ Advances in Cement
Research 26 (2): 101–113.
Kovler, K, OM Jensen, K Kovler, OM Jensen, OM Jensen, O Bjntegaard, TA Hammer, EJ Sellevold, DP Bentz, and EAB
Koenders. 2007. ⬙Internal curing of concrete, state-of-the-art Report of RILEM Technical Committee 196-ICC.⬙
RILEMReport. 41.
Li, FX, YZ Chen, SZ Long, B Wang, and GG Li. 2010. ⬙Research on the preparation and properties of MgO expansive
agent.⬙ Advances in Cement Research 22 (1): 37–44.
Lura, Pietro, Ole Mejlhede Jensen, and Klaas van Breugel. 2003. ⬙Autogenous shrinkage in high- performance cement
paste: An evaluation of basic mechanisms.⬙ Cement and Concrete Research 33 (2): 223–232. doi: http://dx.doi.org/
10.1016/S0008-8846(02)00890-6.
Mangadlao, Joey Dacula, Pengfei Cao, and Rigoberto C. Advincula. 2015. ⬙Smart cements and cement additives for oil
and gas operations.⬙ Journal of Petroleum Science and Engineering 129 (0): 63–76. doi: http://dx.doi.org/10.1016/
j.petrol.2015.02.009.
McCulloch, J, J Gastineau, DL Bour, and K Ravi. 2003. ⬙Life Cycle Modeling of Wellbore Cement Systems Used for
Enhanced Geothermal System Development.⬙ TRANSACTIONS- GEOTHERMAL RESOURCES COUNCIL: 147–154.
Meddah, Mohammed Seddik, and Arezki Tagnit-Hamou. 2010. ⬙Evaluation of rate of deformation for early-age concrete
shrinkage analysis and time zero determination.⬙ Journal of Materials in Civil Engineering.
Mo, L., M. Deng, and A. Wang. 2012. ⬙Effects of MgO-based expansive additive on compensating the shrinkage of cement
paste under non-wet curing conditions.⬙ Cement and Concrete Composites 34 (3): 377–383. doi: http://dx.doi.org/
10.1016/j.cemconcomp.2011.11.018.
Mo, Liwu, Min Deng, and Mingshu Tang. 2010. ⬙Effects of calcination condition on expansion property of MgO-type
expansive agent used in cement-based materials.⬙ Cement and Concrete Research 40 (3): 437–446. doi: http://
dx.doi.org/10.1016/j.cemconres.2009.09.025.
10 SPE-180038-MS

Mo, Liwu, Min Deng, Mingshu Tang, and Abir Al-Tabbaa. 2014. ⬙MgO expansive cement and concrete in China: Past,
present and future.⬙ Cement and Concrete Research 57 (0): 1–12. doi: http://dx.doi.org/10.1016/j.cem-
conres.2013.12.007.
Moradpour, Reza, Ehsan Taheri-Nassaj, Tayebeh Parhizkar, and Masoud Ghodsian. 2013. ⬙The effects of nanoscale
expansive agents on the mechanical properties of non-shrink cement-based composites: The influence of nano-MgO
addition.⬙ Composites Part B: Engineering 55 (0): 193–202. doi: http://dx.doi.org/10.1016/j.compos-
itesb.2013.06.033.
Nelson, Erik B., and Dominique Guillot. 2006. Well Cementing. Second ed: Schlumberger.
Polat, Riza, Ramazan Demirboga, and Waleed H. Khushefati. 2015. ⬙Effects of nano and micro size of CaO and MgO,
nano-clay and expanded perlite aggregate on the autogenous shrinkage of mortar.⬙ Construction and Building
Materials 81 (0): 268 –275. doi: http://dx.doi.org/10.1016/j.conbuildmat.2015.02.032.
Ravi, K., M. Bosma, and O. Gastebled. 2002. ⬙Improve the Economics of Oil and Gas Wells by Reducing the Risk of
Cement Failure.⬙ IADC/SPE Drilling Conference, Dallas, Texas, 2002/1/1/.
Reddy, B. R., Y. Xu, K. Ravi, D. Gray, and P. D. Pattillo. 2009. ⬙Cement-shrinkage measurement in oilwell cementing
- A comparative study of laboratory methods and procedures.⬙ SPE Drilling and Completion 24 (1): 104 –114.
Salomão, Rafael, Cezar C. Arruda, Adriane D. V. Souza, and Leandro Fernandes. 2014. ⬙Novel insights into MgO
hydroxylation: Effects of testing temperature, samples volume and solid load.⬙ Ceramics International 40 (9, Part B):
14809 –14815. doi: http://dx.doi.org/10.1016/j.ceramint.2014.06.074.
Sant, Gaurav, Mukul Dehadrai, Dale Bentz, Pietro Lura, Chiara F Ferraris, Jeffrey W Bullard, and Jason Weiss. 2009.
⬙Detecting the fluid-to-solid transition in cement pastes.⬙ Concrete international 31 (6): 53–58.
Sant, Gaurav, Barbara Lothenbach, Patrick Juilland, Gwenn Le Saout, Jason Weiss, and Karen Scrivener. 2011. ⬙The origin
of early age expansions induced in cementitious materials containing shrinkage reducing admixtures.⬙ Cement and
concrete research 41 (3): 218 –229.
Sant, Gaurav, Pietro Lura, and Jason Weiss. 2006. ⬙Measurement of volume change in cementitious materials at early ages:
review of testing protocols and interpretation of results.⬙ Transportation Research Record: Journal of the Transpor-
tation Research Board 1979 (1): 21–29.
Sant, Gaurav Niteen. 2007. ⬙Examining volume changes, stress development and cracking in cement based systems.⬙
Purdue University.
Sundrarajan, M, J Suresh, and R Rajiv Gandhi. 2012. ⬙A comparative study on antibacterial properties of MgO
nanoparticles prepared under different calcination temperature.⬙ Digest Journal of Nanomaterials and Biostructures
7 (3): 983–989.
Tang, Xiaojia, Lin Guo, Chen Chen, Quan Liu, Tie Li, and Yimin Zhu. 2014. ⬙The analysis of magnesium oxide hydration
in three-phase reaction system.⬙ Journal of Solid State Chemistry 213 (0): 32–37. doi: http://dx.doi.org/10.1016/
j.jssc.2014.01.036.
Zheng, Liu, Cui Xuehua, and Tang Mingshu. 1992. ⬙Hydration and setting time of MgO-type expansive cement.⬙ Cement
and Concrete Research 22 (1): 1–5. doi: http://dx.doi.org/10.1016/0008-8846(92)90129-J.

You might also like