Вы находитесь на странице: 1из 13

SPE 156937

Influence of Nanomaterials in Oilwell Cement Hydration and Mechanical


Properties
Ashok Santra, SPE, Peter J. Boul, and Xueyu Pang, Halliburton

Copyright 2012, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE International Oilfield Nanotechnology Conference and Exhibition held in Noordwijk, The Netherlands, 1214 June 2012.
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 necessar ily 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 prohi bited. 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
Nanotechnology encompasses a wide scope of disciplines and nanomaterials are now being used as commercially viable
solutions to technical challenges in industries ranging from electronics to bio-medicine. Recently, the application of
nanomaterials to solve problems in oilwell cementing has begun to be investigated by several different research groups in the
oil and gas industry. The following uses of nanomaterials have been presented by several independent laboratories as
possibilities in the oilwell cementing industry: (1) nanosilica and nanoalumina as potential accelerators; (2) nanomaterials
including carbon nanotubes (CNTs) with high aspect ratio to enhance mechanical properties; (3) nanomaterials to reduce
permeability/porosity; and (4) nanomaterials to increase thermal and/or electrical conductivity.
In this paper, a review of the aforementioned application concepts is presented with a focus on understanding the role of
multiwall CNTs (MWNTs), nanosilica, and nanoalumina in oilwell cement hydration chemistry. The influence of the
integration of MWNTs into oilwell cement on the physical properties of cement is discussed. Results from an isothermal
microcalorimetric study are presented to help understand the difference between the mode of acceleration of a typical cement
accelerator, like CaCl2, compared to nanosilica and/or nanoalumina.
Introduction
In recent years, the conceptual framework of nanotechnology has demonstrated utility across a wide variety of industries,
from textiles and defense to aerospace and energy. While the materials in these industries can vary greatly, the fundamental
concept that a superior and more functional organization of matter is achieved through an intelligent design from the bottom
up remains the same. This is no different in the cement industry, where self-sensing, self-healing, self-cleaning, and strength
enhancement are concepts now frequently discussed in the research community.
The inclusion of nanoscale particles into Portland cement paste, mortar, or concrete can impart functionality into cements
yielding a variety of different emergent property enhancements. Generally, nanomaterial induced property enhancements
have been observed and reported to include (1) early strength development, (2) increased long-term tensile-to-compressivestrength ratio, (3) viscosity enhancement, and (4) overall increases in the early-stage compressive strength. These
modifications and enhancements are attributed to factors such as densification, particle packing, acceleration, high surface
area, increased nucleation sites, and structural reinforcement. This paper focuses on the influence of nanoparticles,
specifically on the fresh and the hardened properties of the mixture (Senff et al. 2010; Jalal et al. 2012; Khaloo et al. 2011;
Leemann and Winnefeld 2007; Hosseini et al. 2010; Ozyildirim and Zegetosky 2010; Gaitero et al. 2010; Shih et al. 2006;
Ltifi et al. 2011; Gaitero et al. 2008). In particular, how nanosilica and nanoalumina can act as potential accelerators is
investigated. How CNTs with high aspect ratios can enhance the mechanical properties of cements is also discussed. Certain
nanomaterials that serve to reduce the cements permeability/porosity are described, and how some nanomaterials increase
the thermal and/or electrical conductivity in cement is briefly addressed.
The high temperature and pressure within an oil well often requires a particularly resilient cement system and the use of a
number of high-tech additives to help ensure adequate zonal isolation (Ravi et al. 2002). With regard to the desired
mechanical properties in oilwell cements, the Youngs modulus of cement should be as low as possible, while the tensile
strength of cement should be as high as possible. A set Portland cement specimen without any mechanical properties
modification generally has a tensile-to-compressive-strength ratio of ~10%; it has been suggested that some nanoscale
additives can alter this ratio. If this ratio were to be increased as high as possible, the life of the cement sheath in a given well
would be extended. Naturally, mechanical property enhancements would not be complete without tensile strength

SPE 156937

enhancement. This is one of the primary reasons for the investigation of CNT cement composites. Nanotubes have been
reported to yield a significant enhancement of tensile properties in cements. The effect of a specific chemical treatment of the
outer wall of MWNTs is investigated.
The acceleration of the hydration reaction in oilwell cementing is important because it decreases the wait-on-cement
(WOC) time, especially for lower-temperature applications. A shorter WOC time is desirable for saving rig time, which can
be very expensive in offshore wells. Materials, such nano-SiO2 and nano-C-S-H (calcium silicate hydrate), are known to be
good accelerators for cement hydration. Other properties that are affected by these nanoscale particles are
porosity/permeability, viscosity/rheology, strength, durability, shrinkage, and corrosion resistance (Ozyildirim and Zegetosky
2010; Gaitero et al. 2008). Additionally, it has been suggested that nanomaterials can fill the spaces between CSH gel and
act as a nanofiller to reduce effective porosity/permeability. Conversely, nanomaterials having tremendously high surface
area can unnecessarily increase the viscosity at a required slurry density, causing pumpability issues and air entrainment.
Therefore, proper optimization is necessary to achieve optimal porosity reduction and rheology.
To achieve this kind of optimization, silica fume is often used. The benefit of adding silica fume to Portland cement
mixtures has been known to the cement and concrete industry for decades. Silica fume is an amorphous silica with a typical
specific surface area of approximately 20 m2/g and an average particle size of about 150 nm (Senff et al. 2010; Jalal et al.
2012; Khaloo et al. 2011). New sources of nanosilica with much finer particle sizes have been developed in recent years and
are known to be much more efficient than silica fume in modifying the various properties of cement-based materials. The
specific surface area of the new nanosilica materials typically range from 60 to 650 m2/g, which is two to three orders of
magnitude higher than that of Portland cement. Therefore, the workability of fresh Portland cement mixtures generally
decreases with the amount of nanosilica used. The effect of nanosilica on the rheological properties of the fresh cement
mixture can be used to the benefit of cement-based materials in that it improves the consistency of certain types of concrete
(i.e., self-compacting concrete) and reduces the probability of bleeding and segregation (Jalal et al. 2012; Leemann and
Winnefeld 2007). In oilwell cementing, these can be used as extenders for designing lightweight slurries with a high waterto-cement ratio. Property enhancements attributed to the inclusion of nanosilica into cement-based materials mainly include
(1) stronger bonding between aggregates and cement paste matrix (Hosseini et al. 2010), (2) increased compressive and/or
tensile strength (Senff et al. 2010; Jalal et al. 2012; Khaloo et al. 2011; Hosseini et al. 2012; Ozyildirim and Zegetosky 2010;
Gaitero et al. 2010; Shih et al. 2006; Ltifi et al. 2011), (3) decreased permeability and hence increased resistance to calcium
leaching and various types of chemical attack (Jalal et al. 2012; Ozyildirim and Zegetosky 2010; Gaitero et al. 2006; Gaitero
et al. 2008), and (4) more uniform microstructure with reduced pore size and volume (Jalal et al. 2012; Ozyildirim and
Zegetosky 2010; Gaitero et al. 2008). However, the difficulty associated with uniformly dispersing the nanosilica increases
with its quantity and specific surface area, which is especially more apparent when a low water-to-cementitious-materials
ratio is used (Senff et al. 2010; Khaloo et al. 2011). Therefore, the optimal amount of nanosilica for enhancing the properties
of a certain cement mixture will vary with the particle sizes of the nanosilica.
The mechanism of silica fume for enhancing the property of cement-based materials is typically attributed to two factors.
The first is that silica fume can act as a filler material to fill the interstitial space between cement particles, resulting in a
higher packing density and lower porosity. The second factor (known as the pozzolanic effect) is that the amorphous silica
can slowly react with calcium hydroxide, one of the main hydration products of Portland cement, to form C-S-H, which is the
hydration product primarily responsible for the strength gain of cementitious materials (Senff et al. 2010; Mondal et al.
2010). Several studies have been reported that concrete strength improvement as a result of the addition of small-size
nanosilica is much higher than that of silica fume, even though the quantity used is much smaller for the former (Jalal et al.
2012; Khaloo et al. 2011; Ozyildirim and Zegetosky 2010). This phenomenon has been explained by a newly proposed
strengthening mechanism of nanosilica in which it acts as nucleation sites for C-S-H, which not only accelerates early
hydration of Portland cement, but also positively modifies the microstructure of C-S-H for stronger and more durable cement
(Gaitero et al. 2010; Thomas et al. 2009; Land and Stephan 2012).
The isothermal calorimetry test is the most commonly used method to evaluate cement hydration kinetics and
mechanisms. The accelerating effect of nanosilica on cement hydration has been confirmed by several different studies in that
the peak hydration rate is always noticeably increased (Leemann and Winnefeld 2007; Thomas et al. 2009; Land and Stephan
2012). Conversely, inconsistent reports appear to exist with regard to whether silica fume increases the cement hydration
peak (Leemann and Winnefeld 2007; Thomas et al. 2009; Land and Stephan 2012; Cheng-Yi and Feldman 1985; Mostafa
and Brown 2005; Senff et al. 2009). In this study, isothermal calorimetry tests were used to evaluate various types of microand nanoparticles on the hydration kinetics of oilwell cement at relatively elevated temperatures to better understand their
roles in cement hydration reactions. In addition, because the cumulative heat evolution data of cement correlates strongly
with its degree of hydration, which is one of the most important factors governing the mechanical properties of a given
cement mixture, its potential of being used to predict the mechanical properties of cement is also investigated.
Experimental
Calorimetry. A Setaram Calvet C-80 calorimeter was used for all isothermal calorimetry tests. The slurries were prepared at
ambient conditions in accordance with API guidelines. For each test, temperature of the sample was increased using a
constant heating ramp of 1.5 K/min to a target value of 60C, which was then held constant for the rest of the curing period.

SPE 156937

Because it typically takes about 1.5 hr for the temperature of the sample to stabilize, the zero point of cumulative heat
evolution data was set at 1.5 hr for the purpose of consistency. A cement retarder was used in the slurry design to help ensure
the main hydration occurred after 1.5 hr for proper measurement (i.e., induction period 1.5 hr).
Slurry Designs. All the slurries prepared for this study were mixable, pourable, non-settling, and were prepared by using
standard API RP 10B-2 (2010). Slurry designs used for the calorimetric measurements are specified in Table 1. The average
sizes of the various silica and alumina used in the slurries in this study are presented in Table 2.
TABLE 1SLURRY DESIGNS USED FOR CALORIMETRIC STUDY
Slurry
No.

Cement

Free-Water-Controlling

Cement

Class H

Agent

Retarder

Anti-Air-Entrainment

Accelerators

Water

Density

Agent

(bwoc)

(% bwoc)

(lbm/gal)

(%)

(% bwoc)

(% bwoc)

100

0.05

0.2

0.03 gal/sk

None

56

14.86

100

0.05

0.2

0.03 gal/sk

2% Nanoalumina

57

14.86

100

0.05

0.2

0.03 gal/sk

10% Microsand

60

14.86

100

0.05

0.2

0.03 gal/sk

10% Silica fume

60

14.86

100

0.05

0.2

0.03 gal/sk

0.45% Nano-SiO2

57

14.86

100

0.05

0.2

0.03 gal/sk

0.9% Nano-SiO2

58

14.86

100

0.05

0.2

0.03 gal/sk

2% CaCl2

57

14.86

TABLE 2AVERAGE PARTICLE SIZE OF THE SILICA AND ALUMINA USED IN THE SLURRIES FROM TABLE 1
Additive

Average Particle Size

Nanosilica

30 nm

Silica fume

~1 micron

Microsand

7 micron

Nanoalumina

140 nm

With regard to the nanotube cement studies, a total of five slurries with a uniform density of 15 lbm/gal were prepared for
the mechanical properties study. Table 3 shows their compositions. For each slurry (except for the control slurry), the
nanomaterial and surfactant (when applicable) were premixed with mixing water in a jar and sonicated for 30 min to produce
a liquid suspension. The liquid suspension was then mixed with the dry blend (Class H cement and a free-water-controlling
agent). The mixing water of the control slurry was not sonicated because no nanomaterials were mixed. Five 1- 3-in.
cylinders were cast from each slurry for mechanical property tests. All the samples were cured in an autoclave at 88C and
2,000 psi for 7 days.
TABLE 3SLURRY DESIGNS USED FOR MECHANICAL PROPERTIES STUDY
Composition

Control

Slurry A

Slurry B

Slurry C

Slurry D

MWNT (% bwoc)

None

0.2

0.1

0.2

0.2

Anti-air-entrainment agent (% bwoc)

0.27

0.27

0.27

0.27

0.27

Cement Class H

100

100

100

100

100

Free-water-controlling agent (% bwoc)

0.45

0.45

0.45

0.45

0.45

Water (% bwoc)

54

54

53

53

53

Surfactant (% bwoc)

None

None

0.5

None

0.5

Mechanical Test Procedures (Table 3 Slurries). One cylinder of each slurry was cut into four 0.5-in. long disks for
Brazilian tensile tests; the other four cylinders were cut and ground to a final length of 2 in. for uniaxial compression tests.
All cylinders were kept under water during preparation and wiped dry with a paper towel just before testing to help ensure
they maintained the same moisture content at the time of testing. Brazilian tensile tests were performed at a fixed loading rate
of 150 lbf/min. Two pieces of cardboard were used as padding material between the disk sample and the loading platen.

SPE 156937

Uniaxial compression tests were performed at a constant displacement rate of 0.0001 in./sec. Axial strain was measured by
two axial extensometers (with a gauge length of 1 in.) directly attached to the sample. Radial strain was obtained by
measuring the circumferential change of the sample.
Isothermal Calorimetry and Early Cement Hydration
Cement Heat of Hydration (HOH) and Calorimetry. Fig. 1 shows a typical HOH plot as a function of elapsed time and
the effects of measurement temperature, which represents the bottomhole temperature in real oil wells. It is clear from the
plot that, with increasing bottomhole temperature, the peak hydration rate is increased, while the time to reach the peak is
decreased. In other words, the rate of cement hydration increases (or accelerates) with increasing bottomhole temperature,
which is consistent with chemical kinetics laws. Such dependence of hydration rate on bottomhole temperature is associated
with the activation energy of the cement. It can also be observed from the plot that the peak width decreases with increasing
bottomhole temperature, indicating faster decrease in hydration rate. This is associated with the particular mechanism of
cement hydration, namely the hydration rate is dependent on the total degree of reaction.

Fig. 1HOH for a Class G (15.8-lbm/gal) mixed with water, suspending aid, and anti-air-entrainment agent, and measured at 30, 45,
and 60C.

HOH vs. Compressive Strength Development. The development of a mechanical property, like compressive strength, is
highly dependent on the rate and extent of cement hydration. In Fig. 2, plots of HOH and compressive strength measured
using an ultrasonic cement analyzer (UCA) acquired from the same slurry at 30 and 60C are shown. The same batch of
slurry was used for both calorimetry and UCA measurements. As can be observed from Fig. 2, compressive strength only
begins to develop once the slurry has begun to hydrate. This is expected because some extent of hydration is required to build
a percolation network to develop compressive strength. Fig. 3 illustrates that compressive strength develops at a similar rate
as cumulative heat evolution during the first 15 hr, both of which also appear to have a similar dependence on curing
temperature. To further demonstrate their correlations, in Fig. 4, compressive strength is plotted as a function of cumulative
heat evolution at two different temperatures (30 and 60C) for the same slurry. It is found that the correlation between
compressive strength and cumulative heat evolution is best described by two linear models with different slopes.
Interestingly, the dividing points between the two linear models appear to be similar for tests performed at different
temperatures (i.e., ~800 psi in UCA compressive strength).

SPE 156937

Fig. 2Plot of HOH and compressive strength measured by UCA acquired from the same slurry at 30C.

Fig. 3Progress of cumulative heat evolution and compressive strength as a function of elapsed time at different curing
temperatures.

SPE 156937

Fig. 4Correlations between cumulative heat evolution and compressive strength at 30 and 60C.

HOH and the Influence of a Traditional Accelerator. While designing slurries for low temperature wells, operators often
face a challenging need for accelerating the cement hydration (or the rate of compressive strength development) to gain
enough strength within a desired short time span, mainly to reduce WOC. Such acceleration is normally achieved by adding
several different traditional accelerators, CaCl2 for example. Fig. 5 shows how the addition of CaCl2 can cause acceleration in
the cement hydration.

Fig. 5HOH for a Class H (16.4-lbm/gal) mixed with water, suspending aid, and anti-air-entrainment agent, and measured at 60C
with varying amounts of CaCl2.

SPE 156937

HOH and the Influence of Nanomaterial. In this work thus far, the possibility of using nanomaterial as an additive for
acceleration is discussed. HOH plots for several nano- and micro-inorganic oxide materials are presented in Fig. 6. In
addition, 2% CaCl2 and the control slurries are studied for comparison.

Fig. 6HOH for a Class H (14.86-lbm/gal as described in Table 1) with various additives, measured at 60C.

As is very clear from Fig. 6, the addition of all the additives studied had an accelerating effect to a certain extent on
cement hydration, as evidenced by their shift in HOH peak position toward the lower hydration time compared to the control
slurry. The following observations can be made regarding the HOH peak width, height, and shape:
Peak width and height for control, nano-SiO2, nanoalumina, and microsand are essentially same.
Peak width and height for the silica fume containing slurry is very similar at early time, with some increased
hydration after the peak.
Peak width for the 2% CaCl2 slurry is relatively lower, but the height is significantly larger compared to the rest.
A cumulative evolution HOH for all the systems studied above has been plotted in Fig. 7. This is another way of looking
at the extent of total cement hydration at any given elapsed time; in addition, it provides an idea of how the compressive
strength development would look as a function of time. A slurry with 2% CaCl2 would have the quickest compressive
strength development, whereas a 0.9% nanosilica slurry is the next quickest. The overall order of rate of compressive strength
development (which is inverse of WOC) during early hours would be expected to be approximately 2% CaCl2 > 0.9% nanoSiO2 > 0.45% nano-SiO2 > 10% silica fume > microsand > alumina > control. The total HOH (area under each HOH curve)
in the first 48 hr for all the slurries is summarized in Table 4. With the exception of silica fume and CaCl2-containing
slurries, all other slurries had very similar total HOH. The observed higher value in the case of silica fume and a lower value
for the 2% CaCl2-containing slurries could be related to their changes in Ca/Si ratio and the Pozzolanic effects of silica fume;
whereas, the observed order of the rate of compressive strength development could be caused by the following factors:
For the SiO2-related slurries, the higher the solubility of SiO2, the faster the rate of strength development, which
would bring the order to 0.9% nano-SiO2 > 0.45% nano-SiO2 > 10% silica fume > microsand.
For the nanoalumina slurry, it appears that there might not be much chemical effect; whereas, a slight increase in the
rate of strength development could be caused by the adsorption of some of the water molecules on the high surface
area crystalline alumina particles, leading to an actual lowering of the amount of available water in the slurry. This
would, in principle, speed up the precipitation of Portlandite needed for breaking the dormant period. This effect
should be there for the silica materials, in addition to the relative availability of the soluble silica.
CaCl2 acts in a different way by enriching the concentration of Ca 2+ itself and thereby speeding up the precipitation
of Portlandite required for breaking the dormant period.

SPE 156937

Fig. 7Cumulative heat evolution in first 48 hr of cement slurry hydration, measured at 60C.

TABLE 4TOTAL HEAT EVOLUTION FOR DIFFERENT SLURRIES CALCULATED AT 48 HR


Slurry No.

Cumulative heat (J/g cement)

371

370

378

413

367

378

344

Effects of Nanotube on Cement Mechanical Properties


CNTs are seamless, one-dimensional cylinders of graphene, where the carbon atoms are hexagonally arranged with respect to
one another, sp2 hybridized, and bridged through electronically delocalized -bonded molecular orbitals. The chemical and
physical properties of CNTs make them ideal candidates for a wide variety of applications, from catalysts and sensors to
composite materials (Bjrnstrm et al. 2004). They are widely known for their high tensile strength (10 to 63 GPa) and
ductility (12% elastic strain capacity) (Gong et al. 2009; Boul et al. 2009; Vigolo et al. 2005).
There are three different basic classifications of CNTs. These are single-wall CNTs (SWNTs), multiwall CNTs
(MWNTs), and carbon nanofibers (CNFs). SWNTs are comprised of a single seamless tube of grapheme, whereas MWNTs
are made up of multiple concentric tubes of graphene separated by 0.37 0.04 nm (Fig. 1). SWNTs are generally 1 nm in
diameter, whereas MWNTs range from 2 to 50 nm. CNFs are MWNTs, which have a diameter of 50 to 200 nm. The lengths
of SWNTs can range from tens of nanometers to several millimeters. MWNTs range in length from 100 nm to 100 um. CNFs
range from 50 to 100 m in length (Yu et al. 2009).
CNTs are not only valued for their superior mechanical properties, their electrical and chemical characteristics are also of
particular interest. The resistance of CNTs is known to increase by a factor of 100 when the strain changes from 0.0 to 3.2%
under tensile loading (Tombler et al. 2000). This has led to the development of CNT-cement composites for strain sensing
(Saafi et al. 2009).
There have recently been a number of demonstrations of the enhancement of cement mechanical properties through
cement reinforcement with CNTs (Han et al. 2009). It is known that the efficacy of fiber-reinforced cement or concrete
depends greatly on the mechanical properties of the fiber employed. Table 5 illustrates the mechanical properties of SWNTs
and MWNTs in comparison with ordinary Portland cement (OPC). Because CNTs are orders of magnitude stronger and
stiffer than OPC, they can share more load from the OPC matrix without fracturing.
CNTs and CNFs are some of the best nanomaterial available for the reinforcement of Portland cement, not only because
of their strength, but also because of their size (ranging from 1 nm in diameter up to tens of microns in length). CNTs and
CNFs offer the possibility of preventing the nucleation and growth of cracks to the microscale from the nanoscale. A lowerconcentration carbon nanomaterial is required to observe similar enhancements of cement mechanical properties, as observed
with micro- and millimeter-scale carbon fiber (Chen et al. 2005).

SPE 156937

TABLE 5PROPERTIES OF CNTS AND OPC (Shah et al. 2009)


Property

SWNT

MWNT

OPC

Elastic modulus (TPa)

~1.0

0.8 to 0.9

0.10 to 0.14

Tensile strength (GPa)

10 to 52

11 to 63

0.002 to 0.010

Elongation at break (%)

15

12

0.02

Density (kg/m3)

< 1500

< 1500

1500

The effective reinforcement of cement with CNTs requires the consideration the following four factors: (1) the kind of
CNT used and the aspect ratio, (2) the technique of dispersion in the cement matrix, (3) the chemical functionality of the
outer CNT wall, and (4) the workability of the fresh cement composite paste. While addressing these four factors in the
following paragraphs, the authors have undertaken a study of the effect of functionalization of the outer wall of CNTs with
carboxylate groups.
There have been conflicting reports in the literature regarding the effect of carboxylate (-COOH) functionalization on
cement composite mechanical strength. Yi et al. (2005) suggested that the presence of defects (chemical functionality) on the
outer nanotube wall in CNTs can allow for greater load transfer than that available in pristine nanotubes, which have no such
additional chemical functionality. In particular, they maintain that while these defects can come at the expense of the
mechanical strength of the carbon fibers, -COOH functional groups can bond with the C-S-H phases in cement, enhancing
the reinforcement efficiency (Yi et al. 2005). In contrast, there have been studies demonstrating that functionalization with
carboxylates can decrease the reinforcing effects of CNTs in Portland cement (Kowald and Trettin 2009). Al-Rub et al.
(2012) found that the strength of Portland cement with pristine or annealed CNT material was higher than carboxy
functionalized CNTs (prepared through oxidizing acid treatment). They hypothesize that this may be because of excessive
formation of ettringite caused by the presence of sulfates. The sulfates are present because oxidizing acids (sulfuric and nitric
acids) were used to functionalize the nanotube material. Musso et al. (2009) also observed a decrease in mechanical strength
in cements with carboxylated nanotubes. They attributed this to that the functionalization caused the nanotube material to be
so hydrophilic that it would absorb the water in the cement mixture, hampering the proper hydration of the cement.
TABLE 6PLASMA ETCHED AND CARBOXYL FUNCTIONALIZED MWNTS
Diameter
13 to 18 nm
Length
3 to 30 m
Surface area
~200 m2/g
Maximum aspect ratio
2300

It has been reported in the literature that like any other nanomaterials, the workability of cement composite pastes is often
reduced by the addition of nanotube material because of their large surface areas. As an example, Kowald and Trettin (2009)
incorporated 0.5% MWNTs into Portland cement and observed a 12% increase in compressive strength. They found,
however, that the paste became too viscous to work with (Kowald and Trettin 2009). There have also been reports internally
of nanotubes changing the rheology enough to make cement unpumpable. A smaller amount of nanotube additive should
have less of an effect on the rheology of the liquid. Therefore, smaller amounts of material, as proposed by Shah et al. (2009)
and Tyson 2011, could yield cements with improved mechanical properties without seriously compromising the pumpability
of the paste.
As can be observed from Figs. 8 and 9, the addition of the carboxy functionalized MWNTs did not increase the
compressive strength of the cement. In fact, the inclusion of these functionalized materials lowers the compressive and tensile
strengths. These materials were studied at 0.2 and 0.1 % bwoc. The effect of MWNTs on compressive and tensile strengths is
shown in Fig. 9. The larger concentration of MWNTs was associated with a lower drop in mechanical strength than the 0.1
%bwoc MWNT material. Additionally, the addition of surfactant as a dispersant for the materials increased the compressive
and tensile strengths of the cements. This is likely because the surfactant enabled a better and more stable dispersion. The
quality of the dispersion is known to strongly influence the outcome of the mechanical properties of a MWNT cement
composite. Figs. 10 and 11 illustrate comparisons the Youngs modulus and Poissons ratio of all the samples.

10

SPE 156937

Fig. 8Mechanical testing of the cements prepared displayed a reduction in strength in the case of MWNT reinforcement.

Fig. 9Comparison of the MWNT additives and their effects on the mechanical strengths of cements.

SPE 156937

11

Without surfactant

With surfactant

1200000

Young's modulus (psi)

1000000
800000
600000
400000
200000
0
0

0.10%

0.20%

MWNT content
Fig. 10There is little significant variation between the samples with regard to the Youngs modulus.

Without surfactant

With surfactant

0.2
0.18

Poisson's ratio

0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0
0

0.10%

0.20%

MWNT content
Fig. 11Comparison of Poissons ratio of all samples.

Summary and Conclusions


In this paper, some recent results from our ongoing research regarding the influence of certain nanomaterials in oilwell
cement hydration and mechanical properties are presented. Following is a summary of the work presented in this paper:
Cumulative cement HOH was found to be proportional to the compressive strength development; therefore, it is
proposed that a cumulative HOH plot can be a predictive tool for following mechanical properties development at
early hours. A proper normalization and calibration with the experimentally measured mechanical properties can
make it a powerful tool.
Cumulative HOH versus compressive strength development revealed a possible two-fold mechanism for strength
development.
The order of acceleration observed was: 2% CaCl2 > 0.9% nano-SiO2 > 0.45% nano-SiO2 > 10% silica fume >
microsand > alumina > control. For the SiO2-related slurries, the higher the solubility, the greater the acceleration,
wherein solubility depends on particle size (higher is better) and crystallinity (amorphous is more soluble that

12

SPE 156937

crystalline). For the nanoalumina slurry, it appears that there might not be many chemical effects; whereas, a slight
increase in the rate of strength development could be caused by the adsorption of some of the water molecules on
the high surface area crystalline alumina particles, leading to an actual lowering of the amount of available water in
the slurry. This would, in principle, speed up the precipitation of Portlandite required for breaking the dormant
period. This effect should be present in the silica materials or any other high-surface-area hydrophilic materials.
CaCl2 acts in a different way by enriching the concentration of Ca 2+ itself, thereby speeding up the precipitation of
Portlandite required for breaking the dormant period.
Except silica fume and CaCl2-containing slurries, all of the other slurries studied have very similar total HOH. The
observed higher value in the case of silica fume and the lower value for the 2% CaCl2-containing slurries could be
correlated to their changes in Ca/Si ratio and the Pozzolanic effects of silica fume.
The rheologies of all CNT cement composites were flowable and appeared pumpable. It is believed that this occurs
because of the lower quantities of nanotube material used (0.1 to 0.2% bwoc). The MWNT material did not show an
enhancement of mechanical properties.
References
Al-Rub, R.K.A.; Tyson, B.M.; Yazdanbakhsh, A.; Grasley, Z. 2012. Mechnical Properties of Nanocomposite Cement Incorporating
Surface-Treated and Untreated Carbon Nanotubes and Carbon Nanofibers. Journal of Nanomechanics and Micromechanics 1-6.
API RP 10B-2, Recommended Practice for Testing Well Cements, first edition, 2010. Washington, DC: API.
Bjrnstrm J., Martinelli, A., Matic, A., Brjesson, L., and Panas, I. 2004. Accelerating Effects of Colloidal Nano-silica for Beneficial
CalciumSilicateHydrate Formation in Cement. Chemical Physics Letters 392: 242248.
Boul, P.J., Turner, K., Li, J., Pulikkathara, M.X., Dwivedi, R.C., Sosa, E., Lu, Y., Kuznetsov., O.V., Moloney, P., Wilkins, R., et al. 2009.
Single Wall Carbon Nanotube Response to Proton Radiation. J. Phys. Chem. 113 (32): 1446714473.
Chen, S.J., Collins, F.G., Macleod, A.J.N., Pan, Z., Duan, W.H., and Wang, C.M. 2011. Carbon Nanotube-Cement Composites: A
Retrospect. 2005. The IES Journal Part A: Civil & Structural Engineering 4 (4): 254265.
Cheng-yi, H. and Feldman, R.F. 1985. Hydration Reactions in Porltand Cement-Silica Fume Blends. Cement and Concrete Research 15
(4): 585592.
Gaitero, J.J., de Ibarra, Y.S., Erkizia, E., and Campillo, I. 2006. Silica Nano-Particle Addition to Control the Calcium-Leaching in CementBased Materials. Physica Status Solidi A 203: 13131318.
Gaitero, J.J., Campillo, I., and Guerrero, A. 2008. Reduction of the Calcium Leaching Rate of Cement Paste by Addition of Silica
Nanoparticles. Cement and Concrete Research 38: 11121118.
Gaitero, J.J., Campillo, I., Mondal, P., and Shah, S.P. 2010. Small Changes Can Make a Great Difference. Transportation Research Record
No. 2141, 15.
Gong, K., Du, F., Xia, Z., Durstock, M., and Dai, L. 2009. Nitrogen-Doped Carbon Nanotube Arrays with High Electrocatalytic Activity
for Oxygen Reduction. Science 323 (5915): 760764. doi: 10.1126/science.1168049.
Han, B., Yu. X., and Kwon, E. 2009. A Self-Sensing Carbon Nanotube/Cement Composite for Traffic Monitoring. Nanotechnology 20:
445501.
Hosseini, P., Booshehrian, A., and Farshchi, S. 2010. Influence of Nano-SiO2 Addition on Microstructure and Mechanical Properties of
Cement Mortars for Ferrocement. Transportation Research Record No. 2141, 1520.
Jalal, M., Mansouri, E., Sharifipour, M., and Pouladkdan, A.R. 2012. Mechanical, Rheological, Durability and Microstrurual Properties of
High Performance Self-Compacting Concrete Containing SiO2 Micro and Nanoparticles. Materials and Design 34: 389400.
Khaloo, A.R., Vayghan, A.G., and Bolhasani, M. 2011. Mechanical and Microstructural Properties of Cement Paste Incorporating Nano
Silica Particles with Various Specific Surface Areas. Key Engineering Materials 478: 1924.
Kowald, T. and Trettin, R. 2009. Improvement of Cementitious Binders by Multiwall Nanotubes. In Zdenk, B. et al., eds. Proceedings of
the 3rd International Symposium on Nanotechnology and Construction (NICOM 2009), Prague, Czech, Berlin, Heidelberg, 31 May2
June.
Land, G. and Stephan, D. 2012. The Influence of Nano-Silica on the Hydration of Ordinary Portland Cement. Journal of Material Science
47: 10111017.
Leemann, A. and Winnefeld, F. 2007.The Effect of Viscosity Modifying Agents on Mortar and Concrete. Cement and Concrete Research
29: 341349.
Ltifi, M., Guefrech, A., Mounanga, P., and Khelidj, A. 2011. Experimental Study of the Effect of Addition of Nano-Silica on the Behavior
of Cement Mortars. Procedia Engineering 10: 900905.
Mondal, P., Shah, S.P., Marks, L.D., and Gaitero, J.J. 2010. Comparative Study of the Effects of Microsilica and Nanosilica in Concrete.
Transportation Research Record No. 2141, 69.
Mostafa, N.Y. and Brown, P.W. 2005. Heat of Hydration of High Reactive Pozzolans in Blended Cements: Isothermal Conduction
Calorimetry. Thermochimica Acta 435 (2): 162167.
Musso, S., Tulliani, J.M., Ferro, G., and Tagliaferro, A. 2009. Influence of Carbon Nanotubes Structure on the Mechanical Behavior of
Cement Composites. Composites Science and Technology 69: 19851990.
Ozyildirim, C. and Zegetosky, C. 2010. Exploratory Investigation of Nanomaterials to Improve Strength and Permeability of Concrete.
Transportation Research Record No. 2142, 18.
Ravi, K., Bosma, M., and Gastebled, O. 2002. Safe and Economic Gas Wells through Cement Design for Life of the Well. Paper SPE
75700 presented at the SPE Gas Technology Symposium, Calgary, Alberta, Canada, 30 April2 May. doi: 10.2118/75700-MS.
Saafi, M. 2009. Wireless and Embedded Carbon nanotube Networks for Damage Detection in Concrete Structures. Nanotechnology 20:
395-502.

SPE 156937

13

Senff, L., Hotza, D., Repette, W.L., Ferreira, V.M., and Labrincha, J.A. 2009. Influence of Added Nanosilica and/or Silica Fume on Fresh
and Hardened Properties of Mortars and Cement pastes. Advances in Applied Ceramics 108 (7): 418428.
Senff, L., Hotza, D., Repette, W.L., Ferreira, V.M., and Labrincha, J.A. 2010. Effect of Nanosilica and Microsilica on Microstructure and
Hardened Properties of Cement Pastes and Mortars. Advances in Applied Ceramics 109 (2): 104110.
Shah, S.P., Konsta-Gdoutos, M.S., and Metaxa, Z.S. 2009. Advanced Cement Based Nanocomposites. In Recent Advances in Mechanics
Selected Papers from the Symposium on Recent Advances in Mechanics. Academy of Athens, Athens, Greece. Springer Press.
Shih, J.Y., Chang, T.P., and Hsiao, T.C. 2006. Effect of Nanosilica on Characterization of Portland Cement Composite. Materials Science
and Engineering A (424): 266274.
Thomas, J.J., Jennings, H.M., and Chen, J.J. 2009. Influence of Nucleation Seeding on the Hydration Mechanisms of Tricalcium Silicate
and Cement. Journal of Physical Chemistry C (113): 43274334.
Tombler, T.W. 2005. Coupled Electro-Mechanical Behavior of Carbon Nanotubes. Nature 2000 (323): 168170.
Tyson, B.M., Al-Rub, R.K.A., Yazdanbakhsh, A., and Grasley, Z. 2011. Carbon Nanotubes and Carbon Nanofibers for Enhancing the
Mechanical Properties of Nanocomposite Cementitious Materials. Journal of Materials in Civil Engineering 23: 1028-1035.
Vigolo, B., Coulon, C., Maugey, M., Zakri, C., and Poulin, P. 2005. An Experimental Approach to the Percolation of Sticky Nanotubes.
Science 309 (5636): 920923. doi: 10.1126/science.1112835.
Yi, G.Y., Wang, P.M., and Zhao, X. 2005. Mechanical Behavior and Microstructure of Cement Composites Incorporating Surface-Treated
Multi-Walled Carbon Nanotubes. Carbon 43: 12391245.
Yu, M.F., Lourie, O., Dyer, M.J., Moloni, K., Kelly, T.F., and Ruoff, R.S. 2009. Strength and Breaking Mechanism of Multiwalled Carbon
Nanotubes under Tensile Load. Science 287: 637640.

Вам также может понравиться