You are on page 1of 12

Geosystem Engineering

ISSN: 1226-9328 (Print) 2166-3394 (Online) Journal homepage: https://www.tandfonline.com/loi/tges20

Stable SiO2–TiO2 composite-based nanofluid


of improved rheological behaviour for high-
temperature oilfield applications

Ravi Shankar Kumar & Tushar Sharma

To cite this article: Ravi Shankar Kumar & Tushar Sharma (2020): Stable SiO2–TiO2 composite-
based nanofluid of improved rheological behaviour for high-temperature oilfield applications,
Geosystem Engineering, DOI: 10.1080/12269328.2020.1713909

To link to this article: https://doi.org/10.1080/12269328.2020.1713909

Published online: 16 Jan 2020.

Submit your article to this journal

Article views: 8

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tges20
GEOSYSTEM ENGINEERING
https://doi.org/10.1080/12269328.2020.1713909

RESEARCH ARTICLE

Stable SiO2–TiO2 composite-based nanofluid of improved rheological behaviour


for high-temperature oilfield applications
Ravi Shankar Kumar and Tushar Sharma
Enhanced Oil Recovery Laboratory, Department of Petroleum Engineering, Rajiv Gandhi Institute of Petroleum Technology, Jais, India

ABSTRACT ARTICLE HISTORY


Nanofluid synthesis in pure water is associated with premature settlement resulting in least disper- Received 7 May 2019
sion stability. Therefore, in this study, polyacrylamide (PAM) is used as viscosity enhancer to improve Accepted 30 December 2019
dispersion stability of nanofluid stabilized by composites of silica and titania. Different techniques KEYWORDS
such as dynamic light scattering measurements, electrical conductivity, scanning electron micro- Nanoparticles; nanofluid;
scopy, and rheological studies are used to support the analysis. The use of silica and titania polymer; rheology; stability;
nanoparticles together with PAM has additional advantage over particle agglomeration, and thus, temperature
the dispersion stability improved. Further, nanofluid stabilized by composites of silica and titania was
tested for rheological measurements at 90°C to find nanotechnology applicability in high-tempera-
ture applications. The shear-thinning behaviour of nanofluids at high temperature (90°C) was least
affected by shear deformation and reduced to 0.48 mPa.s at higher shear rate (4200 s−1), while shear
thinning of PAM solution seriously varied with increasing shear deformation and takes the edge of
0.0005 mPa.s at higher shear rate (4200 s−1). In addition, the thermal stability of nanofluids was better
due to slight decrease in viscosity with increasing temperature, which makes them suitable to be
utilized at high-temperature applications in widespread industrial areas including oilfield where the
temperature becomes a major factor.

1. Introduction temperature (20–50°C) on viscosity values of PAM solu-


tions of concentration ranges from 0.2 to 0.5 wt%, and it
Polyacrylamide (PAM) is one of the cost-effective
was found that viscosity of solutions significantly
water-soluble polymers which is abundantly available
reduced with increasing temperature and finally reaches
in oilfield and largely used for water shut applications
to minimum level of 0.26 cp at 50°C. Therefore, it is
(Fatimah, Almohsin, Michael, Bataweel, & Alsharaeh,
important to improve the rheological behaviour of poly-
2019), permeability modifications (Marandi, Salehi, &
mer solutions using a better stabilizer that not only
Moghadam, 2018), oil recovery applications (Sharma,
works as rheology modifier but also shows least tem-
Iglauer, & Sangwai, 2016), and viscosity enhancement
perature dependency.
(Kumar & Sharma, 2018). PAM, being a high-molecu-
Nanofluid is an interdisciplinary nanomaterial which
lar-weight polymer, can provide important modifica-
is synthesized by small-size nanoparticles(NPs) and
tions in rheological behaviour of the base solution in
shows promising results in various applications such
which it is added. Additionally, PAM solutions are
as oil recovery (Zheng, Cheng, Wei, Li, & Zhang,
expected to exhibit the least effect of shear deformation,
2017), exploration (Liu, Jin, & Ding, 2016), fluid loss
and as a result, they can provide relatively better sweep
control (Medhi, Chowdhury, Gupta, & Mazumdar,
efficiency and mobility ratio in oilfield applications than
2019), and altering rock wettability (Rostami, Sharifi,
conventional methods (El-hoshoudy, Desouky, Al-
Aminshahidy, & Fahimpour, 2019). Moreover, the
Sabagh, Betiha, & Mahmoud, 2017). However, the use
enhanced use of NPs in several industrial applications
of PAM is not only limited to normal oilfield conditions
is supported by their superior thermal conductivity
but it is also used for high-temperature applications
(Keyvani, Afrand, Toghraie, & Reiszadeh, 2018) and
where most of the polymer methods fail to perform
heat transfer ability (Moradi, Toghraie, Isfahani, &
and show deformed rheological behaviour (Tapias,
Hosseinian, 2019), where temperature becomes an
Lizcano, & Lopes, 2018). The effect of temperature on
insignificant factor on properties of nanofluids. For
viscosity of PAM solutions is reported in several reports;
example, silica NPs are solid charged particles of size
for example, Yang (1999) studied the effect of increasing

CONTACT Tushar Sharma tusharsharma.ism@gmail.com Enhanced Oil Recovery Laboratory, Department of Petroleum Engineering, Rajiv Gandhi
Institute of Petroleum Technology, Jais, UP, 229304, India
© 2020 The Korean Society of Mineral and Energy Resources Engineers (KSMER)
2 R. S. KUMAR AND T. SHARMA

ranges from 1 to 100 nm which are thermally conduc- of the study lies in the use of novel nanomaterials of
tive and possess the ability to improve the stability and SiO2–TiO2 those are thermally and rheologically stable
rheological behaviour of polymer solutions (Sharma, than pure PAM solution, a key factor for the use of
Kumar, & Sangwai, 2014). However, all NPs in nano- nanomaterials in oilfield schemes where temperature
fluid are expected to agglomerate to minimize the sur- becomes a limiting factor for PAM solution.
face energy, resulting in NPs to form large-size clusters Therefore, in this study, we have reported the synthesis
those intend to settle over the period of time (Yang, of a silica nanofluid using the varying concentration of
Kelkar, Corti, & Franses, 2016). The agglomeration SiO2 NPs (0.2–0.4 wt%, size ~15 nm) dispersed in base
reduces the overall population of NPs from the suspen- PAM solution of 1000 ppm. NP TiO2 (size ~20 nm), as a
sion, resulting in nanofluid to show premature settle- stabilizer, is included in silica nanofluid, considering its
ment (an instability issue) and deformed rheological favourable impact on NP agglomeration and nanofluid
properties (Kumar & Sharma, 2018). In nanofluid stability. The concentration of TiO2 NP was kept constant
synthesis, the temperature is of key importance as it and low, viz. 0.05 wt%, as suggested for TiO2 use in
can significantly affect the stability of dispersed NPs by oilfield schemes (Bayat, Junin, Samsuri, Piroozian, &
providing additional energy to Brownian motion of Hokmabadi, 2014). The viscosity vs. shear rate measure-
NPs. NPs, under enhanced Brownian motion, are sus- ments was conducted at high temperature (90°C).
ceptible to undergo random collisions with other NPs,
and as a result, NPs may agglomerate and settle. Thus, a
stable nanofluid may exhibit premature NP settlement 2. Experimental work
at high temperature and become rheologically unstable 2.1. Materials
(Zadeh & Toghraie, 2018). The occurrence of agglom-
eration between similar NPs can be controlled by intro- For this study, polymer PAM, with purity >90% and
ducing another NP (of similar charge) that will help to molecular weight 107 g/mol, was obtained from SNF
increase electrostatic repulsion and stabilize the nano- Floerger, India, and used as received. Hydrophilic SiO2
fluid. The inclusion of TiO2 (as co-stabilizer) in prime (amorphous, purity 99.50%, and size ~15 nm) and hydro-
nanofluid can render the effect of particle settlement via philic TiO2 (rutile, purity 92.00%, and size ~20 nm) were
stabilizing the main NPs as suggested by Yang and Hu used as received from Sisco Research Laboratories (SRL)
(2017). Moreover, TiO2 has mainly three morphologies Pvt. Ltd, India. Deionized water (DI) of electrical con-
which are rutile, anatase, and brookite (Abbasi, 2018). ductivity 0.0054 mS/cm was obtained by Millipore® Elix-
The rutile and anatase morphologies of TiO2 consist of 10 water purification apparatus for the preparation of
tetragonal structure, and it varied along the sharing of PAM solutions and nanofluids. A laboratory-grade mag-
corner edges or corner connections, while brookite has netic stirrer (speed ranges from 0 to 1200 rpm) was used
commonly orthorhombic structure (Abbasi, 2019). to prepare the PAM solutions, and industrial mixer
However, rutile morphology has shown a wide area of (Oster, speed from 6000 to 20,000 rpm) was used to
interest in various industrial applications including bio- synthesize the nanofluids.
medical and oil and gas industry. Since TiO2 is a better
heat conductive material and thermally stable, it may
2.2. Preparation of nanofluid
stabilize the nanofluid at high temperature by control-
ling the rate of agglomeration and provide long-term In this study, the synthesis of nanofluid using SiO2 and
stability to the nanofluid. Moreover, it is very clear that TiO2 NP (as co-stabilizer) is proposed for high-tempera-
NPs in nanofluids have the tendency to show agglom- ture applications. The scheme of the nanofluid formula-
erations and further sedimentation in the nanofluid tion is shown in Figure 1. The nanofluid is prepared in
which leads to instability and make nanofluid a rheolo- viscous base fluid of polymer PAM (1000 ppm) as nano-
gically deformed fluid. PAM solution (base fluid) fluid preparation in water was unsuccessful due to pre-
reduced the rate of sedimentation by increasing the mature sedimentation of NPs. To prepare nanofluid, first
viscosity of nanofluid. At high temperature, PAM fluid base fluid of 1000 ppm is prepared by dissolving 1 g of
possesses thermal degradation due to chain scission polymer PAM in DI water at a stirring speed of 400 rpm
(Król-Morkisz & Pielichowska, 2019); it leads to the for 8 h (Kumar & Sharma, 2018). Next, NP SiO2 of
decomposition of the solution. Further, it destabilizes varying concentration (0.2, 0.3, and 0.4 wt%) was added
the rheological behavior and limits their applicability in to viscous PAM solution, and the resulting milky colloi-
oilfield applications. Thus, the inclusion of thermally dal solution is regarded as silica nanofluid. PAM was
stable NP, viz. TiO2, in silica nanofluid of PAM added as a thickening agent to stabilize the dispersed
increases the thermal stability of base fluid. The novelty silica NPs in a nanofluid. Further, NP TiO2 of 0.05 wt%
GEOSYSTEM ENGINEERING 3

Figure 1. Schematic representation of the formulation methodology of stable nanofluid.

was added to silica nanofluid, and the resulting nanofluid ray spectroscopy (EDS/EDX) detector that was used to visua-
is regarded as nanofluid of SiO2–TiO2 composites. The lize the compositional details of NPs in the nanofluid system.
concentration of TiO2 was less, viz. 0.05 wt%, as recom- The rheological measurements for nanofluids were performed
mended (Bayat et al., 2014) to avoid oilfield problems, for viscosity vs. shear rate using HPHT rheometer (MCR-52,
e.g., pore blockage, formation damage, and reduced oil Anton Paar, Austria). The rheometer consists of a bob and
production. Finally, nanofluids were sonicated for 120– cup assembly, and the external shear force was applied
180 min using digital ultra-sonication bath at a frequency through a spindle attached to magnetic coupling. A varying
of 30 Hz to ensure that the nanofluids received homo- shear rate ranges from 0.1 to 1000 s−1 was applied. Digital
genization. A similar procedure is adopted to prepare ultrasonic cleaner (Labman India, frequency ~40 KHz) was
each nanofluid sample as given in Table 1. The compre- used to sonicate and homogenize NPs in formulated nano-
hensive details of the synthesized nanofluids (viz. PST-1, fluids. All chemicals, used to formulate the nanofluids, were
PST-2, and PST-3) and their nomenclature have been weighed using a digital weighing balance (Mettler Toledo,
provided in Table 1. ME204) with a repeatability of 0.1 mg.

4. Results and discussion


2.3. Nanofluid characterization
This section first discusses the stability results for nano-
Nanofluids were characterized for size distribution using
fluids using visual appearance followed by a discussion on
dynamic light scattering (DLS)-based instrument Zetasizer
their DLS and electrical conductivity measurements. SEM
(NanoZS, Malvern® France), which is based on 173° light-
analysis is presented to visualize the presence of titania
scattering angle to analyse colloidal suspensions. For scanning
particles in interstices of silica nanofluid. Finally, rheolo-
electron microscopy (SEM) analysis, a field-emission scanning
gical measurements for viscosity vs. shear rate as a func-
electron microscope (FE-SEM, Nova NanoSEM) was used.
tion of storage period and temperature were discussed.
Nova NanoSEM is configured with an energy-dispersive X-

4.1. Dispersion stability of nanofluids


Table 1. Compositional details of the formulated nanofluid and
corresponding nomenclature. These measurements were conducted to measure the rate
S. of sedimentation in nanofluids; nanofluid exhibiting least
No. Base fluid Nanoparticle concentration Nomenclature NP sedimentation with time is regarded as a stable nano-
1
2
1000 ppm PAM
solution
0.2 wt% SiO2 + 0.05 wt% TiO2
0.3 wt% SiO2 + 0.05 wt% TiO2
PST-1
PST-2
fluid (Kumar & Sharma, 2018). Therefore, the prepared
3 0.4 wt% SiO2 + 0.05 wt% TiO2 PST-3 nanofluids were transferred to a transparent vessel and
kept undisturbed to visualize the changes in colour with
4 R. S. KUMAR AND T. SHARMA

time. The formulated nanofluid samples with base fluid typical oilfield polymer) over a time period of 10 days,
(1000 ppm PAM solution), viz. PST-1 (0.2 wt% SiO2 + and they found that nanofluid exhibited better dispersion
0.05 wt% TiO2), PST-2 (0.3 wt% SiO2 + 0.05 wt% TiO2), stability with insignificant settlement. Therefore, in this
and PST-3 (0.4 wt% SiO2 + 0.05 wt% TiO2), were milky in study, the use of PAM as a base fluid reduced the effect of
appearance at the time of preparation. The nanofluids sedimentation by increasing the viscosity of solution.
progressively exhibited a transition from milky phase to Moreover, the main reason for enhancement in disper-
translucent due to settlement of dispersed NPs. The rate sion stability is attributed to the homogeneous distribu-
of sedimentation in nanofluids was measured gravimetri- tion of silica NPs in nanofluid. The presence of TiO2 may
cally; settled NPs from the vessels were collected and help to maintain uniform distribution of SiO2 NPs by
cleaned using toluene. It was followed by their drying in creating a surface charge or potential in nanofluid and
an oven to measure the actual weight of settled NPs out of reduced the effect of agglomeration. Thus, it can be
total NPs used at the time of preparation. The first sign of inferred that PST-3 nanofluid showed higher dispersion
colour change in PST nanofluids was observed after 3 stability with 0.4 wt% SiO2 and 0.05 wt% TiO2 NPs in the
weeks (~21 days) of preparation as shown in Figure 2. As system.
the gradual settlement occurs in PST nanofluids, samples
were carefully transferred to another transparent vessel
4.2. DLS measurements for nanofluids
for further analysis. The dispersion stability for PST-1
was observed 35 days, whereas it was 39 days for PST-2 Zeta potential and size distribution for nanofluids were
and more than 45 days for PST-3 as given in Table 2 measured at 25°C just after the sonication to avoid the
which is in line with our previous study (Kumar & possibility of NP aggregation, leading to a change in the
Sharma, 2018). For the nanofluids, it was observed that size of NPs. Zeta potential and size distribution for each
only 8–10% of NPs settled in first 21 days. For PST-1, nanofluid were measured on the first day of preparation,
more than 80% NPs settled after 35 days, while ~77% and the results are given in Table 1. SiO2 NPs were
and ~69% NPs settled for PST-2 and PST-3 after 35 days smaller in size (~15 nm); however, DLS measurements
of preparation, respectively. Saha, Uppaluri, and Tiwari showed an increase in NP size for all PST nanofluids,
(2018) also conducted the dispersion stability analysis for which indicates that NPs tend to agglomerate due to
3 wt% silica nanofluid in 5000 ppm Xanthan Gum (XG, a Vander Waals forces. On the first day of preparation,

Figure 2. Nanofluid stability analysis using visual appearance on (a) 1st day, (b) 21st day, and (c) 35th day for formulated nanofluid, viz.
PST-1, PST-2, and PST-3.

Table 2. DLS measurements (zeta potential and average size distribution) of nanofluid as a function of sample storage time.
Zeta potential (mV) Size distribution (nm)
Dispersion stability
Nanofluid samples (days) 1st day (%) error 35th day (%) error 1st day (%) error 35th day (%) error
PST-1 35 −39.5 ±1.7 −29.5 ±0.5 302 ±1.5 127 ±0.4
PST-2 39 −42.7 ±0.9 −31.2 ±1.5 328 ±2.0 111 ±0.9
PST-3 49 −45.4 ±2.5 −34.1 ±1.2 399 ±1.5 107 ±0.5
GEOSYSTEM ENGINEERING 5

the average NP size of PST-1 and PST-2 was determined possess higher zeta potential as compared to silica NPs
as 302 ± 1.5% and 328 ± 2.0% nm, respectively, while NP (Kumar & Sharma, 2018). Generally, the presence of
size in PST-3 was determined as 399 ± 1.5% nm (see titania in silica nanofluid produces electrostatic poten-
Table 2). The occurrence of agglomeration leading to tial which leads to uniform distribution of NPs; thus, the
NP settlement in silica nanofluid is in accordance with nanofluid consisted of 0.4 wt% silica and 0.05 wt%
Kumar and Sharma (2018), where the average particle titania (PST-3) showed more stability than PST-2 and
size was observed 412 nm with 0.5 wt% silica and 0.05 PST-1. Moreover, a stable nanofluid represents the uni-
wt% titania in the 1000 ppm of PAM solution; however, form distribution of NPs in nanofluid which is never-
the average size of NPs reduced with time. In addition, theless superior for high-temperature applications. The
after 35 days of the preparation, the size of remaining more the NPs in suspension, the more they can partici-
NPs in PST-1 (127 ± 0.4% nm) and PST-2 (111 ± 0.9% pate for thermal stability.
nm) nanofluids was higher than the one with PST-3
(107 ± 0.5% nm) as given in Table 2. Typically, NPs
4.3. Electrical conductivity measurements of
with higher agglomeration settle faster with time; as a
nanofluids
result, the nanofluid body exhibits NPs of lower size as
shown in Table 2. The reduction in size of the NPs in all The electrical conductivity of nanofluid has a good depen-
PST nanofluids after the storage period of 35 days indi- dency on NP presence/settlement in suspension, and any
cates that nanofluid with higher size exhibited faster change in electrical conductivity directly indicates the
sedimentation. Hence, most of the NPs of higher sizes changes in NP population in bulk phase of nanofluid
settled and nanofluid left with NPs of lower size (see (Minea, 2019). Therefore, high electrical conductivity is
Table 2). The reduced average size of NPs showed the the representation of homogeneous nanofluid system
uniform distribution of NPs in the nanofluid due to the where NPs can easily communicate and show the better
presence of TiO2 and PAM solution. transportation of surface charges, while low conductivity
Zeta potential is inversely proportional to NP stabi- represents loosen structure that barely has NPs to commu-
lity in colloidal suspension; DLS measurements showing nicate conductivity as a large fraction of NPs settled with
insignificant change in zeta potential value (over a per- time. Thus, the increase in agglomeration significantly
iod of 35 days) of nanofluid illustrate uniform NP dis- reduces the population of dispersed NPs in nanofluid and
tribution without sedimentation/agglomeration, and can be easily identified by the electrical conductivity mea-
therefore, decreasing zeta potential is a clear indication surements (Minea, 2019). The electrical conductivity mea-
of increasing instability in nanofluids. In addition, any surements for formulated nanofluids were carried out at
colloidal suspension is called stable if zeta potential lies room temperature using electrical conductivity meter
in between ±30 mV limits (Kumar & Sharma, 2018). For (Hanna Instruments, USA, Model HI98129) with uncer-
nanofluid sample PST-1, zeta potential value was deter- tainty of ±0.5% to 3.0%. During each measurement, the
mined as −39.5 ± 1.7% mV on the first day of prepara- used electrode of conductivity meter was prudently cleaned
tion, and it decreases to −29.5 ± 0.5% mV after 35 days and dried at ambient temperature. The electrical conduc-
of storage period. It indicates that PST-1 became tivity measurements were conducted on fresh nanofluid
unstable after 35 days as zeta potential reaches closer samples and samples stored for 35 days to understand the
to the limit of stability. Similarly, for PST-2 and PST-3, phenomenon of NP settlement in nanofluids. The electrical
zeta potential values were −42.7 ± 0.9% and conductivity measurements, as a function of the storage
−45.4 ± 2.5% mV, respectively, on the first day of pre- period, for nanofluid samples, viz. PST-1, PST-2, and PST-
paration. These values reduced to −31.2 ± 1.5% and 3, are shown in Figure 3. From Figure 3, it was observed
−34.1 ± 1.2 mV after 35 days of storage time as given that the electrical conductivity of fresh nanofluids was 288,
in Table 2. Thus, PST-2 and PST-3 were still stable 292, and 297 µS/cm for PST-1, PST-2, and PST-3, respec-
colloidal suspensions than PST-1. The decrease in zeta tively. The electrical conductivity of PST-3 nanofluid was
potential indicates more settlements of NPs in the nano- found higher than that of PST-1 which is in accordance
fluid, resulting in lower dispersion stability in PST-1 as with zeta potential results of PST-3. It confirms that the
compared to PST-2 and PST-3. From the results, the high value of electrical conductivity is the sign of uniform
zeta potential of PST-1 is lower than PST-3 for both 1st NP distribution which helped to grant more surface
and 35th days (storage period). This indicates that tita- charges, resulting in achieving higher magnitude of elec-
nia composites with SiO2 in PST-3 are less agglomerated trical conductivity, consistent with the findings of Zawrah,
than the ones of PST-1 which is phenomenal and makes Khattab, Girgis, El Daidamony, and Abdel Aziz (2016). In
PST-3 a relatively more stable nanofluid than PST-1. addition, the high value of electrical conductivity for PST-3
From the literature, it was observed that titania NPs relatively shows uniform NP distribution which is helpful
6 R. S. KUMAR AND T. SHARMA

Figure 3. Electrical conductivity of formulated nanofluids as a function of time at ambient temperature.

to access more surface charges, resulting in higher magni- electrical conductivity of nanofluid reduced (Minea, 2019).
tude of electrical conductivity in the system (Zawrah et al., Thus, this study on the electrical conductivity of nanofluid
2016). After the 25 days of storage, there was an increase in indicates that nanofluid systems are metastable where NPs
electrical conductivity, and its value increased to 305, 319, are subjected to agglomerate and show changes in electrical
and 321 µS/cm for PST-1, PST-2, and PST-3, respectively conductivity with time. This study then summarized that
(Figure 3). The increase in electrical conductivity shows the NPs formed clusters of large agglomeration within the
aggregation period of NPs in the nanofluid, which indicates nanofluid, and this was regarded as the main reason for
that NPs formed clusters of larger size. Thus, NPs came reduction in electrical conductivity of nanofluid. Moreover,
closer and communicate easily to participate in electro- the reduction of the surface potential of NPs within the
phoresis which is also consistent with the previous study nanofluid showed irregular pathways for electrically con-
on electrical conductivity measurements for graphene ductive NPs, and thus, the electrical conductivity of nano-
oxide NPs (Hadadian, Goharshadi, & Youssefi, 2014). fluid reduced.
However, after 35 days of storage, a significant reduction
in electrical conductivity values was observed; electrical
4.4. SEM/EDS analysis of nanofluids
conductivity reduces to 235, 266, and 281 µS/cm for PST-
1, PST-2, and PST-3, respectively. The reduction in elec- For nanofluids, SEM/EDS analysis was performed to
trical conductivity indicates nanofluid consists of a reduced visualize the morphological structure of formed clusters
population of NPs to communicate in the system which and the presence of NPs in nanofluid, respectively.
can be easily supported by lower zeta potential value of From Figure 4(a,b), it was observed that the size of
PST-1 (Table 2). The electrical conductivity of PST-3 nano- NPs in the PST samples was greater than the individual
fluid was measured higher than PST-1 which is in accor- particle size used at the time of preparation of nanofluid.
dance with zeta potential results of PST-3, which is in Therefore, it is clear that the NPs in nanofluid agglom-
agreement with the reported electrical conductivity results erated in the form of clusters of larger size which was
of Minea (2019) where less agglomerated NPs proposed also observed in the size distribution study using DLS
higher electrical conductivity within nanofluid. However, method. Furthermore, the presence of TiO2 NP in the
the premature agglomeration significantly reduced the sur- interstices of SiO2 nanofluid was confirmed by EDS
face charge of NPs resulting in surface potential of NPs and mapping results for PST-1 in Figure 5. From Figure 5,
GEOSYSTEM ENGINEERING 7

Figure 4. FE-SEM image showing morphological structure of agglomerated NPs in the nanofluid (a) PST-1 and (b) PST-3.

it can be seen that TiO2 NPs uniformly distributed and through a spindle attached to magnetic coupling. A vary-
helped to stabilize SiO2 NPs in a nanofluid. Nanofluids ing shear rate ranges from 0.1 to 1000 s−1 was applied
were synthesized to find their usage in oilfield applica- (Moldoveanu, Ibanescu, Danu, & Minea, 2018). For the
tions where pore size of hydrocarbon formations is an effect of high temperature, viscosity measurements were
important factor. The smaller size and reduced sedi- conducted at 90°C and compared with the results at 25°C.
mentation rate allow nanofluids to easily invade the The viscosity of 1000 ppm PAM solution and formulated
pore space in the formation and helped to enhance oil nanofluids was measured thrice, in order to produce
recovery (Sharma et al., 2016). Thus, nanofluid exhibit- repeatability and accuracy in the results as presented in
ing smaller NP size will exhibit easy and unblocked Table 3. Viscosity as a part of rheology is one of the
transportation through pore space of formation. important parameters to describe the fluid behaviour of
viscous samples such as polymer solutions (Lu et al.,
2015), emulsions (Narukulla, Ojha, & Sharma, 2017;
4.5. Rheological analysis of nanofluids
Sharma et al., 2014), and nanofluids (Chen, Xie, Li, &
The viscosity measurements for fluid systems were per- Chen, 2018). These measurements were conducted on
formed using HPHT rheometer (MCR-52, Anton Paar, freshly prepared nanofluids and nanofluids after 35 days
Austria) to understand the effect of NP settlement on flow of preparation to understand the effect of NP settlement
behaviour. The rheometer consists of a bob and cup on flow behaviour. For the effect of high temperature,
assembly, and the external shear force was applied viscosity measurements were conducted at 90°C and
8 R. S. KUMAR AND T. SHARMA

Figure 5. FE-SEM image showing elemental mapping for SiO2 and TiO2 in PST-1 nanofluid after 35 days of storage period.

compared with the results at 25°C. The viscosity result for on shear deformation which makes them suitable to be
PAM solution is shown in Figure 6(a); the zero-shear utilized for varying shear flow conditions. Thus, as com-
viscosity (at 26.5 s−1) of PAM solution was determined pared to PAM solution, freshly prepared nanofluids were
as 1.42 mPa.s which is in agreement with the literature rheologically stable fluid without showing the sign of
value (Kumar & Sharma, 2018). The inclusion of SiO2 shear-thinning non-Newtonian. Figure 6(b) shows the
and TiO2 in PAM fluid significantly increased the viscos- effect of storage period (35 days) on viscosity of nano-
ity to 2.32 ± 4.5%, 2.41 ± 2.41%, and 2.54 ± 3.0% mPa.s fluids; the viscosity decreased to 1.61 ± 4.0%, 1.99 ± 1.5%,
for PST-1, PST-2, and PST-3, respectively, as shown in and 2.43 ± 2.5% mPa.s for PST-1, PST-2, and PST-3,
Figure 6(a) which is in accordance with our previous respectively. In addition, the flow behaviour of all nano-
study on viscosity determination for silica nanofluid fluids is dominated by shear-thinning non-Newtonian.
(Kumar & Sharma, 2018). The enhancement in the visc- The rate of shear thinning was higher in PST-1 than PST-
osity of polymer solution due to the inclusion of NPs is in 3 nanofluid; as a result, its viscosity reached a minimum
agreement with the reported study (Mahbubul, Saidur, level of 0.2 mPa.s at high shear rates. The rheological
Amalina, Elcioglu, & Okutucu-Ozyurt, 2015). In addi- measurements confirm that PST-3 was rheologically
tion, it is to be noted here that the PAM solution exhib- stable fluid than PST-1 as PST-1 exhibited higher viscos-
ited shear-thinning behaviour; viscosity decreases with ity reduction with increasing shear rate than PST-3
increasing shear rate. However, the flow behaviour of (Figure 6(b)). This indicates that rheological property
nanofluid seems Newtonian as shown by almost plateau (viscosity) of PST-1 is highly dependent on shear condi-
in viscosity variation with the shear rate in Figure 6(a). tions. This might be attributed to the fact that PST-1 after
Therefore, nanofluid flow behaviour is least dependent 35 days hardly left with NPs to absorb energy during
shear deformation, resulting in PST-1 exhibiting higher
shear thinning than PST-3. This can be easily correlated
Table 3. Viscosity results with accuracy at ambient temperature
with shear thinning of PAM solution which is sharp as it
and at high temperature (90°C).
Viscosity (mPa.s) at 25°C Viscosity (mPa.s)
exhibited no NPs to absorb energy during shear deforma-
Samples Day-1 (%) error Day-35 (%) error @ 90° C (%) error
tion, and therefore, PST-1 rheology is slightly stable than
PAM fluid 1.42 ±5.0 0.85 ±3.5 0.65 ±6.0 PAM as it contains some NPs in suspension. This again
PST-1 2.32 ±4.5 1.61 ±4.0 1.25 ±4.0 shows that PST-3 exhibited less agglomeration with time
PST-2 2.41 ±2.5 1.99 ±1.5 2.23 ±2.0
PST-3 2.54 ±3.0 2.43 ±2.5 2.39 ±1.5
which is in line with its zeta potential and electrical
conductivity results. Thus, TiO2 composites with SiO2
GEOSYSTEM ENGINEERING 9

Figure 6. Viscosity with shear rate results of formulated nanofluid as a function of storage time on (a) 1st day and (b) 35th day, and (c)
variation of viscosity with shear rate of PAM fluid and nanofluids at 90°C.

provided improved rheological properties in PST-3 than clear that the viscosity of fluid systems decreased with
PST-1 which is of key importance for various applications increasing temperature. However, the effect of temperature
where conventional colloidal suspensions face challenges. was relatively least on nanofluid flow behaviour which might
The knowledge of rheological properties at high tempera- be due to the uniform dispersion of NPs in nanofluid body,
ture is of key importance for oil recovery applications. consistent with the findings of Vallejo, Pérez-Tavernier,
Therefore, viscosity measurements were performed at 90°C Cabaleiro, Fernández-Seara, and Lugo (2018). Moreover,
for all fluid systems and compared to visualize the least effect the viscosity reduction was less in PST-3 as compared to
of temperature on flow properties of fluid. The effect of high PST-2 and PST-1, which is a clear indication of thermal
temperature (90°C) on the viscosity of PAM solution and stability due to the presence of titania NPs within nanofluid
nanofluids was investigated, and the results are provided in (Raud, Hosterman, Diana, Steinberg, & Will, 2017). The fact
Figure 6(c). It was observed that the viscosity of PAM solu- of enhanced thermal stability in NP-based colloidal suspen-
tion and nanofluids reduced; however, PAM solution exhib- sions is in line with the reported findings in the literature
ited a higher reduction than nanofluids due to thermal (Amiri et al., 2015; Cherecheş, Prado, Cherecheş, Minea, &
degradation of polymer chains by chain scission. The visc- Lugo, 2019). A similar trend of viscosity reduction was
osity of PAM solution reduced to 0.65 ± 6.0% mPa.s (at 26.5 observed by Zadeh and Toghraie (2018) where viscosity of
s−1) at 90°C, while the viscosity of nanofluids reduced to 1.25 silver/ethylene glycol nanofluid was determined for different
± 4.0%, 2.23 ± 2.0%, and 2.39 ± 1.5% mPa.s for PST-1, PST- volume fraction and temperature (up to 55°C). It was found
2, and PST-3, respectively, as shown in Figure 6(c). Thus, it is that viscosity reduced from 21 mPa.s (25°C) to ~12 mPa.s
10 R. S. KUMAR AND T. SHARMA

(55°C) for 0.5% of volume fraction of nanofluid. Thus, it is Disclosure statement


clear that the effect of high temperature was least on nano-
No potential conflict of interest was reported by the authors.
fluid viscosity than PAM solution due to the presence of NPs
in the system. Since TiO2 is a thermally stable material, it
helped silica nanofluid to possess lesser shear deformation Funding
against temperature. However, PST-1 exhibited a higher
reduction in viscosity than PST-2 and PST-3. Since NP This work was supported by the Science and Engineering
Research Board, Department of Science and Technology,
concentration in PST-1 was lower, enough thermal bridging
Govt. of India [SB/S3/CE/057/2015].
between PAM and NPs could not be developed, and NPs
exhibited significant Brownian motions, resulting in higher
reduction in viscosity. The inclusion of TiO2 NPs in silica References
nanofluid provided hybrid composites of least dependency
Abbasi, S. (2018). Investigation of the enhancement and opti-
on temperature. Agromayor et al. (2016) tested graphene mization of the photocatalytic activity of modified TiO2
oxide NPs for thermal and heat transfer applications and nanoparticles with SnO2 nanoparticles using statistical
concluded that the use of NPs within nanofluid is a promis- method. Materials Research Express, 5(6), 066302.
ing solution for thermal applications where conventional Abbasi, S. (2019). Photocatalytic activity study of coated ana-
fluids face challenges. The designed nanofluid with the tase-rutile titania nanoparticles with nanocrystalline tin
dioxide based on the statistical analysis. Environmental
hybrid composites of silica and titania in this study is suitable Monitoring and Assessment, 191(4), 206.
for several industrial applications including oilfield where the Agromayor, R., Cabaleiro, D., Pardinas, A., Vallejo, J.,
temperature becomes a major challenge. Fernandez-Seara, J., & Lugo, L. (2016). Heat transfer per-
formance of functionalized graphene nanoplatelet aqueous
nanofluids. Materials, 9, 455.
Amiri, A., Sadri, R., Shanbedi, M., Ahmadi, G., Kazi, S. N.,
5. Conclusions Chew, B. T., & Zubir, M. N. M. (2015). Synthesis of ethy-
lene glycol-treated graphene nanoplatelets with one-pot,
The thermal stability of nanofluid, prepared by compo- microwave-assisted functionalization for use as a high per-
sites of SiO2–TiO2, is investigated for rheological proper- formance engine coolant. Energy Conversion and
ties at 90°C and compared with the ones of simple Management, 101, 767–777.
aqueous PAM solution (1000 ppm) to find the suitability Bayat, A. E., Junin, R., Samsuri, A., Piroozian, A., &
Hokmabadi, M. (2014). Impact of metal oxide nanoparti-
of nanofluid in oilfield applications. The study highlights
cles on enhanced oil recovery from limestone media at
important fundamental aspects of nanofluid synthesis, several temperatures. Energy & Fuels : an American
dispersion stability, NP agglomeration, and rheological Chemical Society Journal, 28(10), 6255–6266.
properties. SiO2 and TiO2 NPs were smaller in size (~15 Chen, X., Xie, X. T., Li, Y., & Chen, S. (2018). Investigation of
and 20 nm, respectively) and remained suspended for the synergistic effect of alumina nanofluids and surfactant
more than a month in the aqueous phase of PAM, while on oil recovery-interfacial tension, emulsion stability and
viscosity reduction of heavy oil. Petroleum Science and
in water, these NPs settled within a day. The enhanced Technology, 36(15), 1131–1136.
dispersion stability in nanofluid was induced by the inter- Cherecheş, E. I., Prado, J. I., Cherecheş, M., Minea, A. A., &
action between NPs and PAM entanglements. The Lugo, L. (2019). Experimental study on thermophysical
absorption of PAM chains on SiO2–TiO2 composites properties of alumina nanoparticle enhanced ionic liquids.
resulted in the formation of a steric monolayer around Journal of Molecular Liquids, 291, 111332.
El-hoshoudy, A. N., Desouky, S. E. M., Al-Sabagh, A. M., Betiha,
NPs that provided good strength to the network of dis-
M. A., & Mahmoud, S. (2017). Evaluation of solution and
persed NPs. Consequently, higher viscosity is obtained rheological properties for hydrophobically associated polyacry-
for nanofluids. In addition, nanofluids exhibited the least lamide copolymer as a promised enhanced oil recovery candi-
shear thinning than PAM solution; however, a slight date. Egyptian Journal of Petroleum, 26(3), 779–785.
change in shear thinning of nanofluids is attributed to Fatimah, A., Almohsin, A., Michael, F. M., Bataweel, M., &
time-dependent NP agglomeration and settlement which Alsharaeh, E. (2019). Polymer nanocomposites for water
shutoff application – A review. Materials Research Express,
had an inverse relationship with the concentration of 6, 032001.
SiO2 NPs. Furthermore, the thermal stability of nano- Hadadian, M., Goharshadi, E. K., & Youssefi, A. (2014).
fluids performed well than PAM solution, resulting in Electrical conductivity, thermal conductivity, and rheologi-
nanofluid viscosity and shear thinning marginally chan- cal properties of graphene oxide-based nanofluids. Journal
ged with increasing temperature (90°C). Thus, this study of Nanoparticle Research, 16(12), 2788.
Keyvani, M., Afrand, M., Toghraie, D., & Reiszadeh, M. (2018).
concludes that nanofluid can be an alternative solution
An experimental study on the thermal conductivity of cer-
for oilfield high-temperature applications where a con- ium oxide/ethylene glycol nanofluid: Developing a new cor-
ventional polymer fluid may show challenges. relation. Journal of Molecular Liquids, 266, 211–217.
GEOSYSTEM ENGINEERING 11

Król-Morkisz, K., & Pielichowska, K. (2019). Thermal decom- the presence of salts for wettability alteration. Journal of
position of polymer nanocomposites with functionalized Dispersion Science and Technology, ISSN : 1532:2351,1–12.
nanoparticles: Polymer composites with functionalized Saha, R., Uppaluri, R. V. S., & Tiwari, P. (2018). Silica nano-
nanoparticles, in Polymer composites with functionalized particle assisted polymer flooding of heavy crude oil:
nanoparticles : synthesis, properties, and applications, Emulsification, rheology, and wettability alteration charac-
Elsevier Inc., cop. 2019. ISBN: 978-0-12-814064-2. 430–435 teristics. Industrial & Engineering Chemistry Research, 57
Kumar, R. S., & Sharma, T. (2018). Stability and rheological proper- (18), 6364–6376.
ties of nanofluids stabilized by sio2 nanoparticles and SiO2-TiO2 Sharma, T., Iglauer, S., & Sangwai, J. S. (2016). Silica nano-
nanocomposites for oilfield applications. Colloids and Surfaces fluids in an oilfield polymer polyacrylamide: Interfacial
A: Physicochemical and Engineering Aspects, 539, 171–183. properties, wettability alteration, and applications for che-
Liu, H., Jin, X., & Ding, B. (2016). Application of nanotech- mical enhanced oil recovery. Industrial and Engineering
nology in petroleum exploration and development. Chemistry Research, 55(48), 12387–12397.
Petroleum Exploration and Development, 43(6), 1107–1115. Sharma, T., Kumar, G. S., & Sangwai, J. S. (2014). Enhanced
Lu, X., Jiang, H., Li, J., Zhao, L., Pei, Y., Zhao, Y., . . . Fang, W. oil recovery using oil-in-water (o/w) emulsion stabilized by
(2015). Polymer thermal degradation in high-temperature nanoparticle, surfactant and polymer in the presence of
reservoirs. Petroleum Science and Technology, 33(17–18), NaCl. Geosystem Engineering, 17(3), 195–205.
1571–1579. Tapias, F. A., Lizcano, J. C., & Lopes, R. B. (2018). Effects of
Mahbubul, I. M., Saidur, R., Amalina, M. A., Elcioglu, E. B., & salts and temperature on rheological and viscoelastic beha-
Okutucu-Ozyurt, T. (2015). Effective ultrasonication pro- vior of low molecular weight HPAM solutions. Revista
cess for better colloidal dispersion of nanofluid. Ultrasonics Fuentes: El reventón energético, 16(1), 19–35.
Sonochemistry, 26, 361–369. Vallejo, J. P., Pérez-Tavernier, J., Cabaleiro, D., Fernández-
Marandi, S. Z., Salehi, M. B., & Moghadam, A. M. (2018). Sand Seara, J., & Lugo, L. (2018). Potential heat
control: Experimental performance of polyacrylamide hydro- transfer enhancement of functionalized graphene nano-
gels. Journal of Petroleum Science and Engineering, 170, 430–439. platelet dispersions in a propylene glycol-water mixture.
Medhi, S., Chowdhury, S., Gupta, D. K., & Mazumdar, A. Thermophysical profile. The Journal of Chemical
(2019). An investigation on the effects of silica and copper Thermodynamics, 123, 174–184.
oxide nanoparticles on rheological and fluid loss property Yang, L., & Hu, Y. (2017). Toward TiO2 nanofluids—Part 1:
of drilling fluids. Journal of Petroleum Exploration and Preparation and properties. Nanoscale Research Letters,
Production Technology, ISSN: 2190: 0566, 1–11. 12, 417.
Minea, A. A. (2019). A review on electrical conductivity of Yang, M. H. (1999). Rheological behavior of polyacrylamide
nanoparticle-enhanced fluids. Nanomaterials, 9, 1592. solution. Journal of Polymer Engineering, 19(5), 371–381.
Moldoveanu, G. M., Ibanescu, C., Danu, M., & Minea, A. A. Yang, Y., Kelkar, A. V., Corti, D. S., & Franses, E. I. (2016).
(2018). Viscosity estimation of Al2O3, SiO2 nanofluids and Effect of interparticle interactions on agglomeration and
their hybrid: An experimental study. Journal of Molecular sedimentation rates of colloidal silica microspheres.
Liquids, 253, 188–196. Langmuir, 32(20), 5111–5123.
Moradi, A., Toghraie, D., Isfahani, A. H. M., & Hosseinian, A. Zadeh, A. D., & Toghraie, D. (2018). Experimental investiga-
(2019). An experimental study on MWCNT–water nano- tion for developing a new model for the dynamic viscosity
fluids flow and transfer in double-pipe heat exchanger of silver/ethylene glycol nanofluid at different temperatures
using porous media. Journal of Thermal Analysis and and solid volume fractions. Journal of Thermal Analysis and
Calorimetry, 137(5), 1797–1807. Calorimetry, 131(2), 1449–1461.
Narukulla, R., Ojha, U., & Sharma, T. (2017). Stable & re- Zawrah, M. F., Khattab, R. M., Girgis, L. G., El Daidamony,
dispersible polyacryloyl hydrazide–Ag nanocomposite H., & Abdel Aziz, R. E. (2016). Stability and electrical
Pickering emulsions. Soft Matter, 13, 6118–6128. conductivity of water-base Al2O3 nanofluids for different
Raud, R., Hosterman, B., Diana, A., Steinberg, T. A., & Will, applications. HBRC Journal, 12(3), 227–234.
G. (2017). Experimental study of the interactivity, specific Zheng, C., Cheng, Y., Wei, Q., Li, X., & Zhang, Z. (2017).
heat, and latent heat of fusion of water based nanofluids. Suspension of surface-modified nano-SiO2 in partially
Applied Thermal Engineering, 117, 164–168. hydrolyzed aqueous solution of polyacrylamide for
Rostami, P., Sharifi, M., Aminshahidy, B., & Fahimpour, J. enhanced oil recovery. Colloids and Surfaces A:
(2019). Enhanced oil recovery using silica nanoparticles in Physicochemical and Engineering Aspects, 524, 169–177.

You might also like