Role of Primary and Secondary Surfactant Layers ... - ACS Publications

May 10, 2016 - (19) Sreeja, V.; Jayaprabha, K. N.; Joy, P. A. Water-Dispersible. Ascorbic-Acid-Coated Magnetite Nanoparticles for Contrast Enhance- me...
0 downloads 0 Views 1MB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article

Role of Primary and Secondary Surfactant Layers on the Thermal Conductivity of Lauric Acid Coated Magnetite Nanofluids Ramanujam Lenin, and Pattayil Alias Joy J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b12476 • Publication Date (Web): 10 May 2016 Downloaded from http://pubs.acs.org on May 14, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Role of Primary and Secondary Surfactant Layers on the Thermal Conductivity of Lauric Acid Coated Magnetite Nanofluids Ramanujam Lenin1, 2 and Pattayil Alias Joy1, 2,* 1

Physical and Materials Chemistry Division, CSIR- National Chemical Laboratory Pune 411008, India

2

Academy of Scientific and Innovative Research (AcSIR), CSIR- National Chemical Laboratory, Pune 411008, India

Abstract Lauric (dodecanoic) acid coated magnetite nanoparticles with different amounts of primary and secondary surfactant layers on the surface of the nanoparticles have been synthesized. Two sets of the surfactant coated nanoparticles are prepared; one with comparable amount of primary surfactant and the other with comparable amount of secondary surfactant. Nanofluids are prepared by dispersing the surfactant coated nanoparticles in toluene. Stability of the nanofluids is found to decrease with increasing amount of the secondary surfactant on the surface of the nanoparticles, due to the increased hydrophilic nature of the particles in the non-polar solvent. Thermal conductivity and viscosity of the nanofluids are found to increase with increasing the amount of the secondary surfactant layer on the surface of the nanoparticles. The enhanced thermal conductivity for fluids with particles having larger amount of the secondary surfactant is ascribed to the lower dispersibility of the particles in toluene due to the exposure of the acid group of the surfactant to the hydrophobic solvent, leading to aggregation of the particles. Only a small increase in the thermal conductivity is observed for fluids with larger amount of the primary surfactant on the surface of the nanoparticles due to the increased dispersibility owing to the large number of hydrophobic tail groups of the surfactant. Larger enhancement in the thermal conductivity is observed in the presence of a small magnetic field for the fluids containing particles with lower amount of the secondary surfactant. The overall results suggest that the thermal conductivity of the nanofluids depend on the amount and nature of the primary and secondary surfactants on the dispersed particles which in turn determine the interaction between the base fluid and the surfactant and therefore the dispersibility and stability of the nanofluids.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1.

Page 2 of 34

Introduction

Magnetic nanofluids are surfactant coated superparamagnetic nanoparticles dispersed in liquid carriers. Magnetic nanofluids are unique materials with tunable thermophysical properties in the presence of a magnetic field, and are important for their various applications.1-3 Tunable thermal conductivity of the magnetic nanofluids in the presence of a magnetic field is a recent research area for heat transport applications.4,

5

Experimental

studies have shown large enhancement in the thermal conductivity of magnetic nanofluids even in the presence of very small applied magnetic fields.6 Most of the applications using magnetic nanofluids require stable fluids without any sedimentation of the particles even in the presence of high magnetic fields. Although many experimental results are reported on the enhanced thermal conductivity of nanofluids, reproducibility is a major concern.7 Agglomeration and sedimentation are the most common problems encountered, due to the high surface energy of the dispersed nanoparticles. This is more complicated in the case of fluids containing magnetic nanoparticles, due to dipolar interaction between the particles in addition to van der Waals attraction.8 Stability of a fluid is the key factor in the thermal conductivity studies and applications of magnetic nanofluids, where the surfactant plays an important role in the dispersion and stability.9, 10 Use of various surfactants for stabilizing and dispersing magnetic nanoparticles, in suitable solvents, have been reported, such as long chain fatty acids,11, 12 sulphonic and phosponic acids,13 polymers,14-17 other organic acids,18, 19 etc. Oleic acid, containing a carboxylic acid head group and a hydrophobic tail end, is extensively studied, as a surfactant, for making magnetic nanofluids in non-polar solvents.2022

The fatty acid surfactant separates the nanoparticles away from one another by steric

repulsion due to the long aliphatic hydrophobic chain on the surface of the nanoparticles. The steric repulsion suppresses the van der Waals force of attraction between the particles as well as the dipole-dipole magnetic interaction between the magnetic nanoparticles. Long chain saturated and unsaturated fatty acids (anionic stabilizers) such as capric, lauric, myristic, stearic and oleic acids are also used to stabilize the magnetic nanoparticles in water. The surface of the nanoparticles is coated with the primary fatty acid surfactant through chemical bond between the oxygen atom in the carboxylic group and the iron atom in the magnetic nanoparticles. Then a secondary fatty acid layer over the primary layer is formed through weak hydrophobic interactions between the carbon chains of the primary and secondary surfactants. The carboxylic group of the secondary surfactant makes the particles hydrophilic 2 ACS Paragon Plus Environment

Page 3 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

and allow to disperse in water.23-25 Wang et al26 studied bilayer surfactant coated magnetic nanofluid using sodium oleate as the primary surfactant and sodium dodezyl benzene sulphonate (SDBS) as the secondary surfactant, and it is found that the strength of the interaction between the inner (primary) and the outer (secondary) layers is weaker than the interaction between the inner layer and the surface. Surfactants on the surface of the nanoparticles not only provide stability to the nanofluids but also has great influence on the physical, chemical, and thermophysical properties of the magnetic nanofluids.9,27-31 Fu et al32 have reported decrease in the magnetization with increase in the number of layers of coating. Regmi et al33 found that the magnetohydrodynamic behavior of a magnetic fluid can be controlled by choosing the appropriate surfactants of different chain lengths. Nowak et al34 studied the effect of hydrodynamic diameter and core composition on magnetoviscous effect of magnetic fluids, using different surfactants. Vekas et al35 reported that the rheology of a magnetic fluid is highly sensitive to the type of stabilization (monolayer/bilayer or hydrophobic/hydrophilic) and quality of the surfactants. The thickness of the surfactant layer also has significant effect on the magnetic and magnetoviscous properties of magnetic fluids. The relative viscosity of short chain carboxylic acid (lauric acid, myristic acid) stabilized magnetic fluids is found to be less than that of the magnetic fluids stabilized using surfactants with larger chain length (oleic acid).11 Patel et al36 studied the effect of surfactant on heat transfer properties, and concluded that the effective heat transfer at the surface of the nanoparticles is decided by the type of surfactant coating. Most of the experimental work reported on magnetic nanofluids mainly focuses on the influence of the surfactants (ionic and nonionic), such as the chain length, on the physical, chemical and thermophysical properties. For the preparation of magnetic fluids in non-aqueous liquid carriers, for many applications, the as-synthesized surface coated magnetic nanoparticles in the aqueous phase are transferred to non-aqueous phase by washing with a highly polar solvent to remove the excess surfactant (secondary) on the surface of the particles. However, the complete removal of excess surfactant (to obtain completely monolayer coated particles) is not possible because it may remove the primary surfactant also which is directly attached to the surface of the magnetic nanoparticles.37 Sahoo et al38 reported that small amounts of secondary surfactant is always associated with the primary surfactants by weak van der Waals interaction. This excess surfactant (secondary layer/ bilayer) affects the dispersion of the particles in the nonaqueous carrier liquid. Eventually, the amount of the excess surfactant may affect the 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 34

physical, chemical and thermophysical properties of magnetic fluids. So far there are no studies reported in the literature on the effect of excess surfactant on the thermophysical properties of magnetic nanofluids. In the present work, we have studied the effect of the amount of the primary and secondary surfactants on the thermal conductivity of magnetite (Fe3O4) nanofluids. Studies are made on lauric acid (dodecanoic acid) coated magnetite nanoparticles where samples with different amounts of the surfactant on the surface of the nanoparticles are obtained by controlled washing of the nanoparticles in the reaction medium by changing the polarity of the solvent. 2. Experimental Section Materials For the synthesis of surfactant coated magnetite (Fe3O4) nanoparticles, FeCl3.6H2O (97%), FeCl2.4H2O (99%), and lauric acid (dodecanoic acid, 99%) were purchased from SigmaAldrich and the chemicals were used as-received without further purification. Analytical grade aqueous ammonia (25%), hydrochloric acid (37%), acetone (99.5%), hexane (95%) and toluene (99.5%) were purchased from Merck chemicals and used without any further purification. Double distilled water was used throughout the synthesis. Synthesis Lauric acid coated magnetite nanoparticles were synthesized by the coprecipitation of acidic solution of iron precursors (FeCl3.6H2O and FeCl2.4H2O) taken in the stoichiometric ratio (Fe3+ to Fe2+ ratio of 2:1) using ammonium hydroxide (NH4OH) as the base. The synthesis procedure is similar to that reported earlier for decanoic acid coated magnetite nanoparticles.39 Lauric acid, dissolved in acetone, was mixed with deaerated and preheated (50 oC) FeCl3.6H2O and FeCl2.4H2O (0.4M/0.2M) solution. NH4OH solution was added to the reaction mixture containing the iron precursor and the surfactant at a constant rate of 4 ml/s. A black precipitate was formed immediately after the addition of NH4OH. After complete addition of the base, the reaction temperature was increased to 80 °C and maintained at this temperature for 45 minutes to complete the coating process. The precipitate was completely dispersed in the reaction medium (water) indicating the completion of the coating process. The dispersion was then cooled to room temperature, and part of the dispersion was dried to obtain the nanoparticles in powder form for characterization (sample code: ML0). 4 ACS Paragon Plus Environment

Page 5 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The dispersion obtained was washed with acetone-hexane mixture, by changing the ratio of acetone to hexane which changes the polarity of the solvent, to remove the excess surfactant present on the surface of the nanoparticles as well as the free surfactant present in the reaction medium.40 The water-based fluid obtained in the reaction medium was taken in a 500 ml beaker and acetone-hexane mixture was added from a burette under constant stirring. The coated nanoparticles were transferred from the aqueous to the non-aqueous medium during the washing process. After the washing process, the organic layer containing the nanoparticles was separated using a separating funnel. The separated non-aqueous layer was washed three times with water to remove any unreacted metal ions. Different samples were obtained by repeating the synthesis procedure under identical conditions but washed under different conditions, as shown in Table 1. Finally the nanoparticles dispersed in the nonaqueous layer were dried at room temperature and the dried magnetite nanoparticles were used for further studies (sample codes ML1, ML2 and ML3 for acetone:hexane ratios of 1:1, 2:1 and 3:1, respectively). Another set of samples was obtained by washing with the less polar 1:1 and 2:1 solvent mixtures, and in these cases, the dispersion was stirred well for five minutes after complete addition of the solvent. These samples are labeled as ML1d and ML2d, respectively, where ‘d’ stands for delayed washing. Characterization Powder X-ray diffraction (XRD) studies on the dried powder samples were carried out on a PANalytical X’pert Pro X-ray diffractometer using Cu Kα (1.5418 Å) radiation. Microscopic images of the synthesized samples were taken using a transmission electron microscope (TEM) FEI model TECNAI G2 F30 operated at an accelerating voltage of 300 kV (C S = 0.6 mm, resolution = 1.7 Å). For TEM analysis, samples were prepared by dispersing the dried nanoparticles in a small amount of toluene by ultrasonication and dropped on to a carbon coated copper grid (200 mesh). Infrared spectra were recorded on a Perkin Elmer Spectrumone FT-IR spectrometer. The samples were prepared in the form of pellets using KBr as the blank. Thermogravimetric analysis (TGA) of the dried powder samples were carried out on a SDT Q600 TG-DTA analyser in the temperature range 25500 °C, at the heating rate 10 °C per minute, under nitrogen atmosphere. Magnetic measurements were performed on a SQUID-VSM (Quantum Design). Viscosity measurements on the fluid samples were carried out on a ARES strain controlled Rheometer using cup and bob geometry. Single point viscosity measurements were carried out at a shear rate of 50 S-1. Thermal conductivity of the magnetic fluids was measured using a homemade cell based on the Transient Hot Wire 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 34

(THW) method, where a platinum wire was used as the heat source.39 The measurements have been carried out (in the absence and presence of a magnetic field) on freshly prepared fluid samples. 3. Results and Discussion X-ray diffraction The dried powder samples obtained after washing with the solvent mixtures were characterized by powder X-ray diffraction for phase purity. XRD patterns of all five samples (Figure S1, Supporting Information) confirmed formation of magnetite phase (Fe3O4), without any other impurities. The average crystallite size was calculated for all the samples from the width of the most intense (311) peak in the XRD patterns using the Scherrer equation, D = 0.9λ/β cosθ, where D is the average crystallite size, λ is the wavelength of Xrays, β is full width at half maximum of the peak after correcting for the instrumental contribution to the peak broadening and θ is the angle of diffraction. The average crystallite size for all the samples is obtained as 6±1 nm. Thermogravimetric analysis Thermogravimetric analysis (TGA) is carried out to get information on the amount of surfactant coated on the nanoparticles. Figure 1 shows the TGA curves of free lauric acid (LA) and the different lauric acid coated samples. The TGA curves of neat LA and the assynthesized sample recovered by evaporation of the solvent after coating (ML0) are shown in the inset of Figure 1. TGA curve of LA shows a sharp single step weight loss at ~180 °C, due to its decomposition. For the unwashed coated sample (ML0), two major weight losses are observed below 400 °C and these correspond to desorption, decomposition and evaporation of the surfactant molecules from the surface of the nanoparticles.38 The first broad weight loss at the lower temperature for the coated nanoparticles starts at the decomposition temperature of lauric acid and is extended up to 260 oC, suggesting that the weight loss is not due to free lauric acid and hence corresponds to loosely bound or physically adsorbed surfactant molecules on the surface of the nanoparticles. The second weight loss at a higher temperature (>260 °C) is due to chemically bound surfactant on the surface of the nanoparticles. TGA curves of all the washed lauric acid coated samples also show two major weight losses below 400 °C (Figure 1), similar to that observed for the unwashed sample. The small weight loss below 150 °C, especially in the case of ML3, can be due to the evaporation 6 ACS Paragon Plus Environment

Page 7 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

of the solvent or moisture. Zhao et al41 observed similar kind of major weight losses in addition to solvent weight loss (at a lower temperature) for bilayer oleic acid coated magnetite nanoparticles. Thus, the TGA studies indicate the possibility of primary (chemically bonded) and secondary (physical adsorption) layers of lauric acid on the surface of the nanoparticles.

Figure 1. TGA curves of the lauric acid coated samples washed with different polar solvent mixtures (ML1, ML2, ML3), and samples washed after a delay of five minutes (ML1d & ML2d). Inset shows TGA curves of neat lauric acid (LA) and the unwashed lauric acid coated sample (ML0).

Table 1: Sample codes, amount of total, primary and secondary surfactants present in the different lauric acid coated samples and percentage of thermal conductivity enhancement for the corresponding nanofluids Sample set

Sample code

Acetone : hexane ratio

Unwashed

ML0 ML1 ML2 ML3 ML3 ML2d ML1d

0 1:1 2:1 3:1 3:1 2:1 1:1

Set 1

Set 2

Total Primary surfactant surfactant (±1%) (±1%) 46 29 23 21 21 24 25

11 13 13 13 13 16 17

Secondary surfactant (±1%) 35 16 10 8 8 8 8

Thermal conductivity enhancement (±0.1%) < base fluid 5.9 5.4 2.1 2.1 3.5 3.9

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 34

Table 1 shows the total and the individual weight losses due to primary (chemically bound) and secondary (loosely bound) surfactant layers for all samples, obtained from the TGA curves. The washed samples are divided into two sets, based on the amount of the primary and secondary surfactant molecules on the surface of the nanoparticles, as shown in the Table. In the first set (Set 1 in Table 1), the amount of the primary surfactant is comparable whereas the amount of the secondary surfactant decreases with increasing the polarity of the solvent used for washing. In the second set (Set 2 in Table 1), the samples washed after a time delay (ML1d and ML2d) are compared with the sample obtained after a quick wash using highly polar solvent mixture (ML3). In the case of the samples obtained after delayed washing (ML1d and ML2d) and ML3, the amount of the secondary surfactant is comparable whereas the amount of the primary surfactant decreases with increasing polarity (Set 2 in Table 1). Thus, there are two sets of samples, one with comparable amount of the primary surfactant (ML1, ML2, ML3 in Set 1) and the other with comparable amount of the secondary surfactant (ML3, ML2d, ML1d in Set 2).

Figure 2. Schematic diagram showing the mode of attachment of primary and secondary surfactant layers in the different washed samples. (A) Nanoparticles dispersed in water (in the reaction medium, unwashed), (B) after washing with 1:1 solvent mixture (ML1), (C) after washing with 2:1 solvent mixture (ML2), (D) after washing with 3:1solvent mixture (ML3). 8 ACS Paragon Plus Environment

Page 9 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2 shows the schematic diagram of the mode of coating of the surfactant molecules on the nanoparticles’ surface after washing with different polarities of the solvent mixture by varying the acetone to hexane ratio. The amount of the secondary surfactant decreases with increasing the polarity of the solvent, by increasing the amount of acetone in the acetone-hexane mixture, used for washing. TEM studies TEM images and the corresponding particle size distribution histograms of the samples ML1 and ML3 are shown in Figure S2 (Supporting Information). The average particle sizes obtained from the histograms are 8.2 and 6.8 nm, for ML1 and ML3, respectively. The particle size obtained from the TEM image is slightly larger than the crystallite size calculated from the XRD pattern and the larger size in the TEM images can be attributed to the thickness of the amorphous or surfactant layer on the surface of the nanoparticles. The smaller size of ML3 indicates that the amount of surfactant on the surface of the nanoparticles decreases with increasing the polarity of the solvent used for washing, due to the removal of larger amount of the secondary surfactant when washed with solvent of larger polarity. The larger size of ML1 indicates larger amount of the surfactant on the surface of the nanoparticles due to washing the sample with the less polar solvent (1:1 acetone-hexane mixture). Although the samples for TEM analysis are prepared at the same concentrations, particles having less surfactant on their surface are assembled very closely (ML3 in Figure S2) and particles having larger amount of the surfactant on their surface are well separated (ML1 in Figure S2). Infrared spectroscopy The nature of bonding of the surfactant molecules with the surface of the nanoparticles is studied using IR spectroscopy. The IR spectra of neat lauric acid and the different lauric acid coated magnetite nanoparticles (ML0, ML1, ML2, ML3, ML2d, ML1d), in the 4004000 cm-1 region are compared in Figure S3 (supporting information). IR spectrum of neat lauric acid shows a strong band at 1711 cm-1 which corresponds to the carbonyl stretching (-C=O) vibration and the bands at 2851 cm-1 and 2927 cm-1 are due to symmetric and asymmetric stretching of the -CH2 group, respectively. The bands at 1407 cm1

and 1460 cm-1 in the spectra of the free acid and the coated particles correspond to -CH2

deformation. The medium intense band at 937 cm-1 in the spectra of free lauric acid corresponds to O-H out of plane bending and it is one of the characteristic vibrations for the 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 34

dimeric form of fatty acid molecules. The absence of this band in the spectra of all lauric acid coated samples indicates the absence of O-H out of plane bending vibration (Figure S3), suggesting that the fatty acid molecules exist as monomers.42 The broad intense band at ~600 cm-1 in the spectra of all the lauric acid coated samples corresponds to the stretching vibration of the Fe-O bond of Fe3O4.43 The unwashed sample ML0 shows a strong band at 1400 cm-1, due to the symmetric stretching frequency of the ammonium complexed carboxylate group (C11H23-COO- NH4+) of the secondary layer of lauric acid and/or free lauric acid present in the sample.44 The broad band near 3100 cm-1 in the unwashed sample corresponds to the OH stretching frequency due to the solvent (water). The intensity of the carbonyl band at 1711 cm-1 is reduced considerably after coating. All the coated samples show only a weak band at 1711 cm-1, indicating the presence of only small amount of free acid group. A close view of the spectra in the carboxylate band region, as shown in Figure 3, suggests decreasing intensity of this band (1711 cm-1) with increasing polarity of the solvent used for washing and the intensity of this band decreases for the Set 1 samples, in the order ML3 < ML2 < ML1 (Figure 3(a)). On the other hand, intensity of this band is comparable for the Set 2 samples ML3, ML2d and ML1d (Figure 3(b)). These observations are in line with the results obtained from the thermogravimetric analysis, where the amount of the secondary surfactant is found to decrease with increasing polarity of the solvent used for washing for the first set of samples (Set 1 in Table 1) whereas the amount of the secondary surfactant layer remains the same in the second set of samples (Set 2 in Table 1).

10 ACS Paragon Plus Environment

Page 11 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3. Close view of the IR spectra of the lauric acid coated samples in the carboxylate region. (a) samples in Set 1 and (b) samples in Set 2.

All the coated nanoparticles show new bands at 1526 cm-1, 1582 cm-1 and 1645 cm-1, (as shown in Figure 3) in the carboxylate region, which correspond to the vibration of metalcarboxylate bonds, suggesting the attachment of the COO- group of lauric acid on the surface of the nanoparticles. The positions and separation of the bands due to asymmetric carboxyl vibration (-COOasym) of the metal-carboxylate group from the symmetric carboxyl vibration (-COOsym) at 1433 cm-1, provide information about the type of co-ordination of the surfactant 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 34

molecules on the surface of the nanoparticles. The separation between the bands, Δν (= νCOOasym – νCOOsym), above 200 cm-1 indicates monodentate, Δν below 110 cm-1 is due to chelating bidentate and Δν between 140 cm-1 to 200 cm-1 is due to bridged bidentate ligand.45 Values for Δν for the three new asymmetric stretching vibrational bands at 1526 cm-1, 1582 cm-1 and 1645 cm-1 in the carboxylate region are obtained as 93 cm-1, 149 cm-1 and 212 cm-1, respectively. These values correspond to chelating bidentate, bridged bidentate and monodentate, respectively, suggesting different modes of attachment of lauric acid on the surface of the nanoparticles. Out of these new bands, the band at 1526 cm-1 is more intense than the other two, indicating that chelating bidentate coordination (Δν < 110 cm-1) with the surface of the nanoparticles is the prominent one. Also, the intensity of this band is comparable for all three samples in the set 1 (ML1, ML2 and ML3) whereas intensities of the other bands change with the polarity of the solvent used for washing. The intensity of the band due to bridged bidentate ligand (1582 cm-1) is highest for ML1 and lowest for ML3 and the intensity of the band varies in the reverse order for the monodentate ligand, indicating that washing with increasing polarity of the solvent converts the bridged bidentate ligand to monodentate ligand, apart from removing part of the secondary layer, as suggested by the corresponding decrease in the intensity of the carbonyl band at 1711 cm-1. On the other hand, for the second set of samples (ML3, ML2d and ML1d), the intensity of the band due to monodentate ligand is almost comparable and the intensity of the band due to bridging bidentate ligand increases in the order ML1d > ML2d > ML3, corresponding to the increase in the amount of the primary layer as observed from the TG analysis (Table 1). Ding et al46 reported the chelating bidentate and bridged bidentate bonding of oleate molecule in Feoleate complex. The intensity of the carbonyl band increases with decreasing the polarity of the solvent used, suggesting increasing amount of secondary surfactant on the nanoparticles.47 Magnetic measurements Magnetic measurements have been carried out to obtain more information on the nature of the particles (Figure S4, Supporting Information). Magnetization measurement as a function of magnetic field at room temperature showed absence of magnetic hysteresis loop (zero coercivity) and continuous increase in the magnetization at higher fields suggesting superparamagnetic nature of the nanoparticles. This is further confirmed by temperature dependent field cooled (FC) and zero field cooled (ZFC) magnetization measurements, which also gave additional information on the magnetic interaction between the nanoparticles and 12 ACS Paragon Plus Environment

Page 13 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

hence the effectiveness of the surfactant coating (Figure S4). The superparamagnetic blocking temperatures (TB) of ML1, ML2, and ML3 (Set 1 in Table 1) are obtained as 75 K, 96 K, and 100 K, respectively. Even though the particle sizes of the magnetite core is comparable for all the three samples, TB decreases with increasing amount of the total surfactant. The decreasing blocking temperature with increasing amount of the surfactant indicates reduced anisotropy contribution originating from interparticle interactions, due to the increasing thickness of the non-magnetic layer (surfactant) on the surface of the magnetic nanoparticles.48,49 Detailed description of the magnetization data is provided in the supporting information (with respect to Figure S4). Thermal conductivity For thermal conductivity measurements, nanofluids are prepared by dispersing the different lauric acid coated magnetite nanoparticles in toluene as the base fluid. Fluids of ML1 and ML2 are found to form sedimentation after a week whereas ML3 gave a stable dispersion in toluene, without any sedimentation even after a month. Similarly, stable dispersions are also obtained for the delayed washed samples (ML1d and ML2d) similar to the dispersion of ML3. Thus, the long-time stability of the fluids depends on the amount of the secondary surfactant present on the particles and the nanoparticles with lower amounts of the secondary surfactant (samples ML3, ML2d, ML1d in Set 2) form the most stable fluids. Figure S5 (Supporting Information) shows the photographs of the freshly prepared fluids and a week after the preparation of the fluid samples. The freshly prepared fluids are well-dispersed without any sedimentation, whereas the fluids ML1 and ML2 form sedimentation after a week. The fluids of ML3, ML2d and ML1d showed stable dispersion even after a week. The lower stability of the fluids of ML1 and ML2 is due to the relatively larger number of carboxylic acid groups from the secondary surfactant layer exposed to the non-polar solvent, toluene. The large number of carboxylic groups on the surface makes the particles more hydrophilic and reduces dispersibility in the non-polar solvent. Figure S6 (Supporting Information) shows the photographs of the fluid samples ML1, ML3 and ML1d when away, brought closer to and removed from a laboratory magnet. Photographs of ML1d and ML3 show that the fluids are stable after moving away from the field whereas photograph of ML1 shows that a part of the fluid is not flown back immediately after moving away from the magnet, due to the less dispersibility and/or less solvent surface interaction.

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 34

Figure 4. Variation of the relative thermal conductivity of the fluid of ML1d in toluene with volume percentage of Fe3O4 (filled squares), compared with theoretical models (solid, dashed, and dotted lines). Inset shows variation of the relative viscosity of the fluid with volume percentage of Fe3O4 (open squares) compared with different theoretical models (solid lines) Figure 4 shows the variation of the relative thermal conductivity (keff/kf, where keff is the thermal conductivity of the nanofluid and kf is the thermal conductivity of the solvent) of the nanofluid of ML1d as a function of volume percentage of the Fe3O4 nanoparticles. No appreciable change in the thermal conductivity is observed below 1.6 volume% of Fe3O4 and above which almost linear increase in the thermal conductivity is observed up to the studied concentration of 5.6% of Fe3O4 (filled squares in Figure 4). Similar trend in the variation of the thermal conductivity with volume% has been reported for oleic acid coated magnetite nanoparticles (6.7 nm) dispersed in kerosene.50 Thermal conductivity of a stable dilute suspension of a solid in a liquid carrier can be explained by the classical Maxwell’s model,51 𝑘𝑒𝑓𝑓 3(𝑘𝑝 ⁄𝑘𝑓 − 1)𝜙 =1+ 𝑘𝑓 (𝑘𝑝 ⁄𝑘𝑓 + 2) − (𝑘𝑝 ⁄𝑘𝑓 − 1)𝜙 where keff, kp and kf are the effective thermal conductivity, the thermal conductivities of the particle (Fe3O4) and the base fluid, respectively, and ϕ is the volume fraction of the particles in the fluid. The experimental thermal conductivity ratio (keff/kf) matches with the Maxwell’s model only at higher concentrations (>4 vol%) showing larger deviation at lower 14 ACS Paragon Plus Environment

Page 15 of 34

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

concentrations (dashed line in Figure 4). In the Maxwell’s model, the particles are assumed to be motionless and the model does not consider any interaction between the particles. The Maxwell-Garnett effective medium approximation considers the thermal conductivity of the host material and that of the particles along with the interfacial thermal resistance (Rb) for two-phase solid-solid composite materials containing spherical particles (details in the Supporting Information).52-54 In the case of nanofluids, thermal conductivity of the host material (base fluid) is much lower than the thermal conductivity of the solid inclusions (nanoparticles) and therefore a modified form of the Maxwell-Garnett equation is used for nanofluids,55

𝑘𝑒𝑓𝑓 (1 + 2𝛼) + 2𝜙(1 − 𝛼) = (1 + 2𝛼) − 𝜙(1 − 𝛼) 𝑘𝑓 where, 𝛼 = 2𝑅𝑏 𝑘𝑓 ⁄𝑑, Rb is the interfacial thermal resistance and d is the diameter of the particles. The calculated effective thermal conductivity by including the interfacial resistance (Rb > 0) for the nanoparticles dispersed in toluene is shown in Figure 4 (dotted line). The interfacial resistance (Rb) is considered as 2.16 × 10-8W-1m2K for the nanoparticles dispersed in toluene39 and the particle size is taken as 8.5 nm. The calculated values using the model with interfacial resistance shows much lower values than the experimental results. The experimental thermal conductivity ratio is also compared with the Brownian motion induced microconvection model,55

(1 + 2𝛼) + 2𝜙(1 − 𝛼) 𝑘𝑒𝑓𝑓 = (1 + 𝐴 𝑅𝑒 𝑚 𝑃𝑟 0.333 𝜙) [ ] (1 + 2𝛼) − 𝜙(1 − 𝛼) 𝑘𝑓 where A is constant, Re is the Reynold’s number, Pr is the Prandtl number, m is a constant which depends on the base fluid. The calculated relative thermal conductivity using the microconvection model by taking the constant m=3.375, as reported earlier for magnetite nanofluids,39 is shown in (solid line in Figure 4, details in the Supporting Information). The Brownian motion induced microconvection model shows poor agreement with the experimental results at both higher and lower concentrations. Thus, microconvection cannot be the possible mechanism for the observed enhancement in the thermal conductivity. In the case of superparamagnetic nanoparticles, the magnetic moment fluctuates randomly due to the thermal energy. The magnetic interaction between the nanoparticles is 15 ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 34

negligible for the fluctuating magnetic moments. Moreover, the surfactant stabilized magnetic nanoparticles are separated from one another which reduces the interparticle magnetic interactions between the particles in the fluid. In the case of smaller particles, the van der Waals interaction contributes more to the self-assembly of nanoparticles than the magnetic interactions.56,57 At higher concentrations, the magnetic nanoparticles are closer to each other, and there can be increased van der Waals interaction between the surfactant coated particles. Due to the closer arrangement of the nanoparticles, the interdigitation of the fatty acid alkyl chain on the surface of the nanoparticles with the surfactants in the nearby nanoparticles is possible (Figure S7, Supporting Information).58,59 Even though the magnetic moments of the nanoparticles are highly fluctuating at room temperature, the van der Waals interaction between the surfactant molecules could possibly form small clusters/aggregation-like selfassembly at higher concentrations (Figure S8, Supporting Information). Even in the absence of a magnetic field, there is a possibility for self-assembly of the magnetite nanoparticles due to the small magnetic interaction between the particles.56 The formation of the clusters/aggregates due to the van der Waals interaction in addition to the weak magnetic interaction could be the reason for the observed enhancement in the thermal conductivity at higher concentration (>1.6 vol %). Philip et al50 have reported that the formation of small clusters (dimers and trimers) in the nanofluid containing oleic acid coated nanoparticles is responsible for the thermal conductivity enhancement at higher concentrations. The absence of any appreciable change in the thermal conductivity, at lower concentrations (