You are on page 1of 9

Available online at www.sciencedirect.

com

ScienceDirect
Solar Energy 105 (2014) 82–90
www.elsevier.com/locate/solener

High temperature and long-term stability of carbon nanotube


nanofluids for direct absorption solar thermal collectors
Nathan Hordy ⇑, Delphine Rabilloud, Jean-Luc Meunier, Sylvain Coulombe
Plasma Processing Laboratory, Department of Chemical Engineering, McGill University, 3610 University St., Montréal, Québec H3A 0C5, Canada

Received 22 November 2013; received in revised form 24 January 2014; accepted 10 March 2014

Communicated by: Associate Editor H.-M. Henning

Abstract

Stable dispersions (nanofluids) are produced using plasma functionalized multi-walled carbon nanotubes (MWCNTs). To our knowl-
edge, this study presents a first quantitative demonstration of nanofluid stability over extended periods of time (currently tested up to
8 months) and after intense heating. No agglomeration is found to occur in the water and glycol-based nanofluids after heating at 85
and 170 °C, respectively. Significant agglomeration does occur for suspensions produced using the non-polar TherminolÒ VP-1 heat
transfer fluid. Optical characterization of the nanofluids demonstrates that the MWCNTs are highly absorbing over the majority of
the solar spectrum, allowing for close to 100% solar energy absorption, even at low concentrations and small collection volumes. These
absorption properties coupled with the stability of the nanofluids make them ideal candidates as direct solar thermal collectors.
Ó 2014 Elsevier Ltd. All rights reserved.

Keywords: Solar absorber; Carbon nanotube; Nanofluid; Solar thermal

1. Introduction of the working fluid. Unfortunately, India Ink, as well as


other organic and inorganic chromophores have been
Solar thermal collectors, which capture the Sun’s rays as shown to experience light- and temperature-induced degra-
heat, are currently used world-wide in applications ranging dation (Burke et al., 1982), which can lead to reduced
from hot water heating to large scale electricity production. absorption and fouling of surfaces over time, thus limiting
As demand for cleaner energy continues to grow, solar their use, especially at high temperatures. Rather than
thermal processes are expected to play an increasingly using dyes, the use of nanoparticles, as dispersed and solar
significant role in our global energy future (Philibert, ray-absorbing phase in the heat transfer fluid, represents an
2011). The concept of direct solar thermal collectors, in interesting alternative. Once dispersed in a host heat trans-
which solar radiation is absorbed and transported directly fer fluid, the nanofluid can pass through pumps and pipes
by the working fluid, was first introduced by Minardi and without adverse effects, such as clogging and fouling. Such
Chuang (1975). As most heat transfer fluids (water, glycols, unique properties make the nanofluids ideal candidates for
oils, etc.) are transparent over the majority of the solar direct solar thermal energy absorption applications (Taylor
spectrum (Otanicar et al., 2009), additives, such as India et al., 2011a).
Ink, can be added to improve the absorption characteristics Nanofluids, which can be defined more formally as sta-
ble suspensions of nanoparticles (particles with one dimen-
⇑ Corresponding author. Fax: +1 5143986678. sion <100 nm in size) in a host liquid, have gained a great
E-mail address: nathan.hordy@mail.mcgill.ca (N. Hordy). deal of attention over the past 15 years. Initially considered

http://dx.doi.org/10.1016/j.solener.2014.03.013
0038-092X/Ó 2014 Elsevier Ltd. All rights reserved.
N. Hordy et al. / Solar Energy 105 (2014) 82–90 83

as revolutionary in the heat transfer community, the actual direct solar thermal energy collectors, agglomeration of
heat transfer enhancements (heat conductivity and heat nanoparticles translates directly into a decrease in absorp-
transfer coefficient) fell short of expectations. Recent works tion. Thus, it is essential that nanofluid stability (i.e. nano-
have demonstrated that more promising and scientifically- particles dispersion) be maintained over extended periods
sound applications may stem out of nanofluids, in particu- of time (years).
lar applications utilizing their highly-tunable optical Analytical techniques involving spectral absorbance
properties (Taylor et al., 2013). Indeed, the use of nanofl- (Hwang et al., 2007; Jiang et al., 2003; Yu et al., 2007),
uids as both a volumetric solar collector and heat transfer zeta-potential (Huang et al., 2009; Jiang et al., 2003), par-
fluid is now seen as a method of improving efficiencies and ticle-sizing (Fedele et al., 2011), and analytical centrifuga-
reducing costs in solar thermal devices (Mahian et al., tion (Harel et al., 2013; Krause et al., 2009) have been
2013). Given the high surface-to-volume ratio of nanopar- used as ways to quantify nanofluid stability. All of these
ticles, quantities on the scale of mg/L can be used to obtain techniques can provide initial indications if agglomeration
close to 100% absorption over relatively short distances has occurred, however testing must be conducted over long
(mm range). Furthermore, the concentration of nanoparti- time scales if stability quantification on the order of
cles can easily be controlled such that the incident radiation months or years is needed. Using an analytical centrifuge
is absorbed over the entire volume of nanofluid, instead of to examine the transmittance of the nanofluid over time
over a thin surface layer, thus limiting heat losses to the can provide accelerated stability measurements, as larger
surroundings (Otanicar et al., 2011). Several different types agglomerates will settle quicker than under normal gravita-
of nanoparticles in various base fluids have been modeled tional forces (Krause et al., 2009). However, the extent of
and tested experimentally for this purpose. These include agglomeration, unlike sedimentation, is not directly pro-
metallic nanoparticles (He et al., 2011; Kameya and portional to the applied gravitational force (Melik and
Hanamura, 2011; Lenert and Wang, 2011; Otanicar Fogler, 1984). Long-term testing may still be required. In
et al., 2010; Taylor et al., 2011a), carbon nanotubes addition, suggesting that low sedimentation velocities are
(CNTs) (He et al., 2011; Meng et al., 2012; Otanicar an indication of nanofluid stability, as has been recently
et al., 2010; Taylor et al., 2011a), graphite (Otanicar proposed by Lamas et al. (2012), is somewhat questionable.
et al., 2010; Taylor et al., 2011a, 2011b), carbon black A nanofluid that contains well-separated supernatant (par-
(Han et al., 2011; Sani et al., 2011), and carbon nanohorns ticle-free fluid) and a sediment phase prior to testing should
(CNHs) (Sani et al., 2011). All of these materials have been not be considered a stable and uniform system. If enough
shown to be effective solar absorbers, but graphitic parti- agglomeration has already occurred to create a heteroge-
cles, such as CNTs, are seen as the most promising for neous fluid, then most likely the suspension that remains
low to medium temperature applications (<400 °C) due no longer contains particles that are “nano” in size.
to their high spectral absorptivity over the entire solar Temperature stability of the nanofluid is another key
range (Taylor et al., 2011a). From an economic standpoint, aspect for all heat transfer applications. Typically, nano-
cost is not seen as a prohibitive factor, as the CNT additive particles in a host fluid are stabilized through the use of a
only adds around a dollar per liter of nanofluid produced. surfactant (Yu and Xie, 2012). Unfortunately, most surfac-
One of the main obstacles to the large-scale use of nano- tants decompose upon modest heating and can lose effec-
fluids, in all applications, is the stability of the nanoparti- tiveness at temperatures as low as 70 °C (Wen and Ding,
cles in suspension (Taylor et al., 2013; Yu and Xie, 2012). 2004). Direct solar thermal collectors will require operating
The unique properties of the suspension, attributed to temperatures from 60 °C for solar water heating, to values
small size and homogeneous dispersion of the particles, as high as 400 °C for concentrated solar power (Taylor
are lost if the nanoparticles agglomerate and settle out over et al., 2011b). To date, very few studies have tested the sta-
time. The stability of nanofluids has, up to now, been bility of nanofluids at high temperatures and those that
essentially described qualitatively with little consistency have made such tests reported significant agglomeration
for defining a “stable nanofluid”. More often than not, and irreversible nanofluid degradation (Tavares and
studies claim that stable nanofluids are produced without Coulombe, 2011; Taylor et al., 2011b). Sani et al. (2011)
any quantitative evidence being reported. Photographs of did report good stability of ethylene glycol-based nanofl-
the nanofluids are often provided as proof of stability. uids up to 150 °C, but no evidence was provided to support
However for concentrated nanofluids, particularly those their claim. To the best of our knowledge, no study has
containing CNTs, the suspension is often too dark to deter- been done to quantify the temperature stability of a
mine if settling has occurred (Yu and Xie, 2012). In addi- nanofluid.
tion, even if sedimentation is not present, agglomeration The purpose of this work is to quantitatively examine
of nanoparticles may still have taken place (Fedele et al., both the long-term and high-temperature stability of
2011). Agglomerated particles that remain in suspension CNT nanofluids for use in direct solar absorption. The
may provide the appearance of a stable suspension over a optical properties of three base fluids that are commonly
short period of time, but the available surface area of the used in solar thermal applications, namely ethylene glycol,
particles to achieve the specific purpose (absorb light, propylene glycol and TherminolÒ VP-1, are characterized
transfer heat, etc.), may have decreased significantly. For along with a range of concentrations of corresponding
84 N. Hordy et al. / Solar Energy 105 (2014) 82–90

nanofluids. Aqueous based nanofluids are also included for made prior to knowing the exact mass of MWCNTs pres-
comparison, given that nanoparticle–water dispersions are ent in suspension.
by far the most common in the literature.

2.2. UV–Vis characterization


2. Material and method
The various nanofluids were characterized using a
2.1. Nanofluid synthesis double-beam UV–Vis spectrophotometer (Cary 5000,
175–3300 nm). The samples were enclosed in a SuprasilÒ
The carbon nanotubes (CNTs) used in this study were quartz cuvette with a pathlength of 1 cm. All trials were
grown directly from stainless steel (SS) 316 mesh using a conducted at room temperature with an empty reference
thermal chemical vapor deposition (T-CVD) process. A path. To account for the difference between the refractive
detailed description of the growth process has been previ- index of quartz and air, the acquired spectra were corrected
ously described (Hordy et al., 2013b). The CNTs produced by treating the filled cuvette as a three-slab system, thus
using this procedure are multi-walled (hereafter referred as transmittance data presented represent only the nanofluid
MWCNTs) with an average diameter and length of and not the cuvette. The details of this procedure are
approximately 30 nm and 4 lm, respectively, and form a described in Otanicar et al. (2009 and references therein).
dense forest on the growth substrate. Following the synthe- Measurements were taken over the 200–1500 nm range
sis, the MWCNTs are surface-functionalized using a for the water, ethylene glycol and propylene glycol based
low-pressure capacitively-coupled radio frequency nanofluids, and from 200 to 2300 nm for the TherminolÒ
(13.56 MHz) glow discharge plasma sustained in an VP-1 based ones. Transmittance spectra of the pure base
argon/oxygen/ethane gas mixture (Hordy et al., 2013a). fluids (a) were also measured for reference. These two spec-
This plasma treatment has been shown to covalently graft tral ranges account for, respectively, 85% and 95% of the
oxygenated functionalities, such as carboxylic groups, onto total incident solar energy at air mass 1.5. Spectra were
the surface of the MWCNTs, thus rendering them polar not taken above 1500 nm for water and the glycols, as
and highly dispersible in water. The various nanofluids strong absorption bands in the IR mask any change due
are produced by placing the MWCNTs-covered SS sub- to the addition of the MWCNTs. However, for direct solar
strate into a 30 mL glass vial with approximately 20 mL collection purposes, one can assume that the base fluid
of the base fluid. The vial is then put into an ultrasonic would absorb the majority of the solar irradiation above
bath (Bransonic 2510) and sonicated for 15 min, causing 1500 nm. As this is not the case for TherminolÒ VP-1,
the MWCNTs to fracture at their roots and become sus- spectra were acquired up to 2300 nm. For all fluids the
pended within the base fluid. The mass of the sample is concentration of MWCNTs was limited to 52 mg/L as
measured using a Sartorius MC1 analytical balance (AC the transmitted signal for higher concentrations was below
210P, 0.1 mg precision) before and after in order to deter- the detection limit for the spectrophotometer and
mine the amount of MWCNTs removed. The four base flu- pathlength used. If a higher concentration is needed, one
ids used in this work are: water (treated via reverse could estimate the required concentration using the Beer–
osmosis), ethylene glycol (Fisher Scientific, >95%), propyl- Lambert law (Section 3.1).
ene glycol (Fisher Scientific, 99%), and TherminolÒ VP-1
heat transfer fluid (Solutia Inc., 73.5% diphenyl oxide
and 26.5% biphenyl). The initial samples generated, with 3. Results and discussion
concentrations up to 100 mg/L, were used to produce dilu-
tions corresponding to the concentrations shown in Table 1. 3.1. Optical characterization
Concentrations of two of the trials using TherminolÒ VP-1
(labeled b* and e*) are slightly higher than the correspond- MWCNT nanofluids of various concentrations were
ing samples for the other base fluids, as the dilutions were prepared using four base fluids commonly used in solar
thermal applications. Fig. 1 shows a series of MWCNTs-
ethylene glycol nanofluids with (a: pure ethylene glycol
Table 1 control, b)f: increasing MWCNT concentration). UV–
MWCNT concentration of various nanofluids produced.
Vis transmittance spectra were acquired at regular time
Label MWCNT concentration (mg/L) intervals following synthesis. The extinction coefficient,
a 0 a(k, c) (cm1), which is a function of MWCNT concentra-
b 5.6 ± 0.4 tion, c (mg/L), and wavelength, k (nm), was calculated
b* 6.5 ± 0.5
using Eq. (1), where the transmittance, T(k), is logarithmi-
c 11.0 ± 1.0
d 17.0 ± 1.4 cally proportional to the extinction coefficient for a given
e 27.0 ± 2.3 path length, l (cm) (Perrin, 1948).
e* 33.0 ± 2.4
f 53.0 ± 5.6 T ðkÞ ¼ eaðk;cÞl ð1Þ
N. Hordy et al. / Solar Energy 105 (2014) 82–90 85

(NIR) and as a result, some noise can be seen above


1400 nm. This is due to the strong NIR absorption bands
of the various base fluids used and not to absorption from
the MWCNTs.
The extinction coefficient, also referred to as the attenu-
ation coefficient, includes both scattering and absorption
effects. However due to the small size of the MWCNTs,
it has been shown that scattering effects are very low, and
thus attenuation of the incident electromagnetic energy
a b c d e f
from the MWCNTs will be primarily caused by absorption
Fig. 1. Ethylene glycol-based nanofluids of various MWCNT concentra- (Gan and Qiao, 2012; Mercatelli et al., 2011). Given that
tions photographed eight months after synthesis. Labels a–f refer to the four base fluids are essentially transparent over the vis-
concentrations given in Table 1. ible range (400–700 nm), all of the observed absorption in
this range can be attributed to the MWCNTs. Moreover,
Fig. 2 shows the extinction coefficients (a) of as-pro- in this work optical saturation effects are neglected. As a
duced MWCNT nanofluids for the four base fluids (concen- result, the MWCNT concentration and optical absorption
tration levels are listed in Table 1). All produced nanofluids coefficient, a0 (k) (L mg1 cm1), can be decoupled to obtain
show an absorption peak centered around 260 nm followed the equation commonly referred to as Beer–Lambert law
by broadband absorption across the measured spectral (Eq. (2)).
range. This absorption behavior is characteristic of 0
T ðkÞ ¼ ea ðkÞcl ð2Þ
MWCNTs and can be attributed to 1D van Hove singular-
ities (Yu et al., 2007). Extinction coefficients larger than Eq. (2) predicts that the transmittance should be loga-
7 cm1, which correspond to transmittance below 0.1% at rithmically proportional to the MWCNT concentration,
a path length of 1 cm, led to transmitted intensity levels and thus for perfectly dispersed MWCNTs, the absorption
below the resolution of the detector in the near-infrared coefficient, a0 , should not vary as their concentration

Fig. 2. Extinction coefficient (a) of MWCNT nanofluids measured soon after synthesis. Base fluids are: (A) water, (B) ethylene glycol, (C) propylene glycol
and (D) Therminol VP-1.
86 N. Hordy et al. / Solar Energy 105 (2014) 82–90

Fig. 3. Stored energy fraction as a function of penetration distance for various concentrations of MWCNTs and for the four base fluids: (A) water, (B)
ethylene glycol, (C) propylene glycol, and (D) Therminol VP-1.

increases. If, on the other hand, the MWCNTs start to 2008)). The plasma functionalization treatment covalently
agglomerate as their concentration increases, the absorp- bonds polar functionalities such as carboxylic groups onto
tion coefficient will effectively vary with concentration. the MWCNT surface, which is why the MWCNTs are
DLVO theory, which describes the interaction of particles easily suspended in the three polar solvents (Hordy et al.,
within a liquid medium, predicts that an increase in 2013a). In addition, deprotonation of the carboxylic
agglomeration, caused by van de Waals attraction, will groups leads to a negative surface charge on the
occur at high concentrations as the inter-particle distance MWCNTs, thus limiting agglomeration. As this is not
decreases (Amrollahi et al., 2008; Missana and Adell, the case with TherminolÒ VP-1, agglomeration begins to
2000). Extracting the absorption coefficient (data not occur immediately after synthesis.
plotted) from the spectra collected in the visible range Using the extinction coefficient from each of the nanofl-
(400–700 nm), one can determine if agglomeration occurs. uids, a solar spectrum-weighted stored energy fraction (Am)
The standard deviation of the absorption coefficient for can be calculated (Drotning, 1978).
each of the three polar nanofluids (water, ethylene and pro- R
pylene glycol) was found to be less than 2%, suggesting that S m ðkÞ½1  eaðk;cÞx  dk
Am ðx; cÞ ¼ R ð3Þ
little to no agglomeration has occurred. The TherminolÒ S m ðkÞ dk
VP-1 nanofluids, on the other hand, had a mean standard
deviation of 17%, indicating that agglomeration had taken This quantity represents the fraction of solar energy that
place within hours of synthesis. It should be noted that is absorbed by the nanofluid over a given penetration
concentrations higher than 32 mg/L were not tested for distance, x (cm). Once again it should be noted that given
the TherminolÒ VP-1 based suspensions as severe agglom- the low amounts of scattering from MWCNTs, extinction
eration occurred. It was expected that the plasma function- coefficients are used as an approximation for absorption
alized MWCNTs may not disperse as well nor remain as coefficients. Ignoring the effects of scattering, even though
stable in TherminolÒ VP-1 as it is a non-polar solvent small, would not be ideal if one was trying to use the
(dielectric constant of diphenyl oxide is 3.9 (Knovel, absorption coefficient to precisely model a direct solar
N. Hordy et al. / Solar Energy 105 (2014) 82–90 87

Fig. 4. Transmission spectra for various concentrations of nanofluids, immediately after synthesis (dash) and after 3 months (solid). (A) Water, (B)
ethylene glycol, (C) propylene glycol, and (D) Therminol VP-1. Absorption pathlength was 1 cm.

0.8
Relative Concentration

0.6

0.4

0.2

0
0 30 60 90 120 150 180 210 240
Time (Days)
PG EG Water T VP-1
Fig. 6. Calculated effective MWCNT concentration in Therminol VP-1
Fig. 5. Relative concentration of the nanofluids over time, as determined based nanofluids over an 8 month period, as determined from optical
by absorption spectroscopy. Error bars represent ±1 standard deviation. measurements.

thermal collector. However, as the focus of this work is to


demonstrate the stability and general absorption ability of of integration were set from 200 to 1500 nm, so that the
the produced nanofluids, this assumption is seen as results could be compared to those found in literature
acceptable. (ASTM, 2012). The stored energy fraction at 1 cm fluid
The spectral distribution of the solar irradiation, Sm(k) thickness was within 1.2% of the values reported by
(W m2), was taken at an air mass of 1.5 and the bounds Otanicar et al. (2009) for each of the base fluids. The stored
88 N. Hordy et al. / Solar Energy 105 (2014) 82–90

Fig. 7. Transmittance spectra for various concentrations of nanofluids, after synthesis (dash) and after heating for 1 h at approximately 85% of the base
fluids’ boiling temperatures (solid). (A) Water (80 °C), (B) ethylene glycol (170 °C), (C) propylene glycol (170 °C), and (D) Therminol VP-1 (220 °C).
Absorption pathlength was 1 cm.

energy fractions, calculated using the extinction coefficients quantitative demonstration of a stable nanofluid to date.
reported in Fig. 2, are presented in Fig. 3 for a range of Fig. 4 displays the nanofluids transmission after synthesis
fluid thicknesses that is representative for solar thermal col- and after a 3 month storage period; a clear shift in trans-
lector geometries. It can be clearly seen by comparing mittance can be observed in cases where agglomeration
curves (b)–(f) with curve (a) that the addition of MWCNTs occurs. Using Eq. (2), this increase in transmittance can
has a significant effect on the energy storage ability of the be represented by a decrease in effective MWCNT concen-
nanofluid. At 10 cm of fluid thickness, even the least- tration. As can be seen from Fig. 5, only a small change
concentrated MWCNT nanofluids have a stored energy occurred for both ethylene and propylene glycol-based
fraction of 1. Using these results, one can easily tailor the nanofluids after 8 months (8.6% and 4.4% in effective
nanofluid concentration to obtain close to 100% absorp- concentration, respectively), and a gradual decrease
tion for any particular fluid thickness. occurred for the aqueous nanofluids (approx. 5% per
month). Error bars in Fig. 5, representing ±1 standard
deviation, were not included for the TherminolÒ VP-1
3.2. Long term stability based nanofluid due to the fact these suspensions were
Stability is one of the most critical issues preventing
nanofluids from making the transition from the laboratory Table 2
bench to large-scale use for which years of operation are Temperature to which each nanofluid was heated and boiling temperature
required for direct solar absorption applications. As a test of the corresponding base fluid.
of long-term stability, transmittance spectra of the different Base fluid Heating temp. (°C) Boiling temp. (°C)
nanofluids at various MWCNT concentrations were Water 85 ± 5 100
acquired over an 8-month period. The nanofluids were Ethylene glycol 170 ± 5 197
stored at room temperature and left without any agitation. Propylene glycol 170 ± 5 188
Therminol VP-1 220 ± 5 257
To the best of our knowledge, this is the longest
N. Hordy et al. / Solar Energy 105 (2014) 82–90 89

temperatures going up to 170 °C for the glycols. In addi-


tion, the electrostatic repulsion caused by the deprotona-
tion of the carboxylic functionalities is sufficient to
overcome the increased number of high velocity collisions
between the MWCNTs that occur at high temperatures
(Amrollahi et al., 2008). This was not the case for the Ther-
minolÒ VP-1 based nanofluids, which experience significant
agglomeration after heating to 220 °C (Fig. 7D). As with
the long-term stability testing, the observed rate of agglom-
eration with the TherminolÒ VP-1 based nanofluids
increases with MWCNT concentration. Fig. 8 illustrates
the effect of heating cycles on a nanofluid which suffers
from agglomeration. All three TherminolÒ VP-1 trials
show significant agglomeration after 5 heating cycles. Once
again it should be noted that these are preliminary heating
Fig. 8. Effective MWCNT concentration in Therminol VP-1 nanofluids experiments conducted to give an indication of the stability
after 5 heating cycles at 220 °C (1 h each), as determined from optical of the nanofluids at high temperatures. A more rigorous
measurements.
study of the effect of temperature on the nanofluids and
the plasma-grafted functionalities is currently underway.
inherently instable. Fig. 6 illustrates the effect of agglomer-
ation on the effective MWCNT concentration in Thermi-
nolÒ VP-1 nanofluids for various times following 4. Conclusions
synthesis (up to 240 days). As can be seen the rate of
agglomeration is significant for all concentrations, however This study describes the synthesis and solar energy
more pronounced with the most concentrated sample. We absorption properties of nanofluids made up of a wide con-
calculated an 87% effective MWCNT concentration loss centration range of plasma-functionalized MWCNTs dis-
for the highest concentration (e*), as opposed 51% for the persed in water, ethylene glycol, propylene glycol, as well
least concentrated sample (b*). As discussed in Section 3.1, as TherminolÒ VP-1 heat transfer fluids. Most importantly,
this increased rate of agglomeration at higher concentra- this study reports for the first time a quantitative demon-
tions is predicted by the DLVO theory. It should be noted stration of the high temperature and long-term stability
that as the other nanofluids (EG, PG and H2O based) are of ethylene glycol and propylene glycol-based MWCNT
relatively stable, no concentration dependence on agglom- nanofluids intended for use in solar thermal energy collec-
eration was observed thus far for a period up to 8 months. tors. Long-term, room temperature stability (currently
tested up to 8 months) has been demonstrated for the gly-
col-based nanofluids, while a gradual MWCNT agglomer-
3.3. High temperature stability
ation was observed with the water-based nanofluids. Given
the polar nature of the oxygen functionalities bonded to the
The previous section examined nanofluid stability over
surface of the MWCNTs, high levels of agglomeration
extended periods of time at room temperature. An addi-
were found to occur for the non-polar TherminolÒ VP-1
tional constraint for a nanofluid to be used in a solar ther-
based nanofluids. In the high temperature tests, no agglom-
mal application, or any heat transfer application for that
eration was found to occur after heating to 85 and 170 °C
matter, is its stability at high temperatures. As a prelimin-
for the aqueous and glycol based nanofluids, respectively.
ary investigation of this aspect, a second batch of nanofl-
Optical characterization of various concentrations of the
uids was heated in a loosely capped vial sitting on a
nanofluids demonstrates that the MWCNTs are highly
temperature-controlled hot plate for 1 h to approximately
absorbing over the majority of the solar spectrum, allowing
85% of each base fluid’s respective boiling temperature
for close to 100% solar energy absorption, even at low con-
(Fig. 7). These temperatures were chosen to limit any pres-
centrations and small collection volumes. These absorption
sure build-up during heating. They are given in Table 2
properties, coupled with the long-term and high tempera-
along with the corresponding atmospheric boiling temper-
ture stability of the glycol-based nanofluids, make them
atures of each base fluid. Since the vials were capped, these
ideal candidates for direct solar thermal energy absorption.
experiments resulted in no net loss of the base fluid. The
fluid volume per sample was low (approx. 10 mL), thus
limiting the thermal stratification within the fluid. In total, Acknowledgements
each heating experiment took 2–3 h. As can be seen in
Fig. 7, little to no change occurs for the water- and The authors gratefully acknowledge the financial sup-
glycol-based nanofluids. This indicates that the plasma- port provided by the Natural Sciences and Engineering
introduced functionalities preventing the MWCNTs from Research Council of Canada, the Fonds de recherche du
agglomerating are not damaged at high operating Québec – Nature et technologies and McGill University.
90 N. Hordy et al. / Solar Energy 105 (2014) 82–90

The authors also acknowledge Solutia Inc., for providing a Lenert, A., Wang, E.N., 2011. Optimization of nanofluid volumetric
free sample of TherminolÒ VP-1 heat transfer fluid. receivers for solar thermal energy conversion. Sol. Energy 86, 253–265.
Mahian, O., Kianifar, A., Kalogirou, S.A., Pop, I., 2013. A review of the
applications of nanofluids in solar energy. Int. J. Heat Mass Transfer
References 57, 582–594.
Melik, D.H., Fogler, H.S., 1984. Effect of gravity on Brownian floccu-
Amrollahi, A., Hamidi, A.A., Rashidi, A.M., 2008. The effects of lation. J. Colloid Interface Sci. 101, 84–97.
temperature, volume fraction and vibration time on the thermo- Meng, Z., Wu, D., Wang, L., Zhu, H., Li, Q., 2012. Carbon nanotube
physical properties of a carbon nanotube suspension (carbon nano- glycol nanofluids: photo-thermal properties, thermal conductivities
fluid). Nanotechnology 19, 315701. and rheological behavior. Particuology 10, 614–618.
ASTM G173-03, 2012. Standard Tables for Reference Solar Spectral Mercatelli, L., Sani, E., Zaccanti, G., Martelli, F., Ninni, P., Barison, S.,
Irradiances: Direct Normal and Hemispherical on 37° Tilted Surface. Pagura, C., Agresti, F., Jafrancesco, D., 2011. Absorption and
ASTM International, West Conshohocken. scattering properties of carbon nanohorn-based nanofluids for direct
Burke, A., Etter, D., Hudgens, C., Wiedenheft, C., Wittenberg, L., 1982. sunlight absorbers. Nanoscale Res. Lett. 6, 1–9.
Thermal and photochemical studies of solar-energy absorbers dis- Minardi, J., Chuang, H., 1975. Performance of a black liquid flat-plate
solved in heat-transfer fluids. Sol. Energy Mater. 6, 481–490. solar collector. Sol. Energy 17, 179–183.
Drotning, W., 1978. Optical-properties of solar-absorbing oxide particles Missana, T., Adell, A., 2000. On the applicability of DLVO theory to the
suspended in a molten-salt heat-transfer fluid. Sol. Energy 20, 313–319. prediction of clay colloids stability. J. Colloid Interface Sci. 230, 150–
Fedele, L., Colla, L., Bobbo, S., Barison, S., Agresti, F., 2011. Experi- 156.
mental stability analysis of different water-based nanofluids. Nanoscale Otanicar, T.P., Phelan, P.E., Golden, J.S., 2009. Optical properties of
Res. Lett. 6, 300. liquids for direct absorption solar thermal energy systems. Sol. Energy
Gan, Y., Qiao, L., 2012. Optical properties and radiation-enhanced 83, 969–977.
evaporation of nanofluid fuels containing carbon-based nanostruc- Otanicar, T.P., Phelan, P.E., Prasher, R.S., Rosengarten, G., Taylor,
tures. Energy Fuels 26, 4224–4230. R.A., 2010. Nanofluid-based direct absorption solar collector. J.
Han, D., Meng, Z., Wu, D., Zhang, C., Zhu, H., 2011. Thermal properties Renew. Sust. Energy 2, 033102.
of carbon black aqueous nanofluids for solar absorption. Nanoscale Otanicar, T.P., Phelan, P., Taylor, R.A., Tyagi, H., 2011. Spatially varying
Res. Lett. 6, 1–7. extinction coefficient for direct absorption solar thermal collector
Harel, Y., Azoubel, S., Magdassi, S., Lellouche, J.-P., 2013. A dispers- optimization. J. Sol. Energy Eng. 133, 024501–1–7.
ability study on poly(thiophen-3-yl-acetic acid) and PEDOT multi- Perrin, F., 1948. Whose absorption law? J. Opt. Soc. Am. 38, 72–74.
walled carbon nanotube composites using an analytical centrifuge. J. Philibert, C., 2011. Solar Energy Perspectives. International Energy
Colloid Interface Sci. 390, 62–69. Agency, IEA, Paris.
He, Y., Wang, S., Ma, J., Tian, F., Ren, Y., 2011. Experimental study on Sani, E., Mercatelli, L., Barison, S., Pagura, C., Agresti, F., Colla, L.,
the light-heat conversion characteristics of nanofluids. Nanosci. Sansoni, P., 2011. Potential of carbon nanohorn-based suspensions for
Nanotechnol. Lett. 3, 494–496. solar thermal collectors. Sol. Energy Mater. Sol. Cells 95, 2994–3000.
Hordy, N., Coulombe, S., Meunier, J.-L., 2013a. Plasma functionalization Tavares, J., Coulombe, S., 2011. Dual plasma synthesis and characteriza-
of carbon nanotubes for the synthesis of stable aqueous nanofluids and tion of a stable copper–ethylene glycol nanofluid. Powder Technol.
poly(vinyl alcohol) nanocomposites. Plasma Process. Polym. 10, 110– 210, 132–142.
118. Taylor, R.A., Phelan, P., Otanicar, T., Adrian, R., Prasher, R., 2011a.
Hordy, N., Mendoza-Gonzalez, N.-Y., Coulombe, S., Meunier, J.-L., Nanofluid optical property characterization: towards efficient direct
2013b. The effect of carbon input on the morphology and attachment absorption solar collectors. Nanoscale Res. Lett. 6, 225.
of carbon nanotubes grown directly from stainless steel. Carbon 63, Taylor, R.A., Phelan, P.E., Otanicar, T.P., Walker, C.A., Nguyen, M.,
348–357. Trimble, S., Prasher, R., 2011b. Applicability of nanofluids in high flux
Huang, J., Wang, X., Long, Q., Wen, X., Zhou, Y., Li, L., 2009. Influence solar collectors. J. Renew. Sust. Energy 3, 023104–1–15.
of pH on the stability characteristics of nanofluids. In: Presented at the Taylor, R.A., Coulombe, S., Otanicar, T., Phelan, P., Gunawan, A., Lv,
Symposium on Photonics Optoelectronics, Wuhan, pp. 1–4. W., Rosengarten, G., Prasher, R., Tyagi, H., 2013. Small particles, big
Hwang, Y., Lee, J.K., Lee, C.H., Jung, Y.M., Cheong, S.I., Lee, C.G., Ku, impacts: a review of the diverse applications of nanofluids. J. Appl.
B.C., Jang, S.P., 2007. Stability and thermal conductivity character- Phys. 113, 011301.
istics of nanofluids. Thermochim. Acta 455, 70–74. Wen, D., Ding, Y., 2004. Effective thermal conductivity of aqueous
Jiang, L.Q., Gao, L., Sun, J., 2003. Production of aqueous colloidal suspensions of carbon nanotubes (carbon nanotubes nanofluids). J.
dispersions of carbon nanotubes. J. Colloid Interface Sci. 260, 89–94. Thermophys. Heat Transfer 18, 481–485.
Kameya, Y., Hanamura, K., 2011. Enhancement of solar radiation Yu, W., Xie, H., 2012. A review on nanofluids: preparation, stability
absorption using nanoparticle suspension. Sol. Energy 85, 299–307. mechanisms, and applications. J. Nanomater. 2012, 1–17.
Knovel Critical Tables, 2008. Knovel Critical Tables, second ed. Knovel. Yu, J., Grossiord, N., Koning, C.E., Loos, J., 2007. Controlling the
Krause, B., Petzold, G., Pegel, S., Pötschke, P., 2009. Correlation of dispersion of multi-wall carbon nanotubes in aqueous surfactant
carbon nanotube dispersability in aqueous surfactant solutions and solution. Carbon 45, 618–623.
polymers. Carbon 47, 602–612.
Lamas, B., Abreu, B., Fonseca, A., Martins, N., Oliveira, M., 2012.
Assessing colloidal stability of long term MWCNT based nanofluids. J
Colloid Interface Sci 381, 17–23.

You might also like