Sie sind auf Seite 1von 22

International Journal of Heat and Mass Transfer 96 (2016) 482503

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Review

Combustion and heat transfer characteristics of nanofluid fuel


droplets: A short review
Saptarshi Basu ,1, Ankur Miglani 1
Department of Mechanical Engineering, Indian Institute of Science, Bangalore 560012, India

a r t i c l e

i n f o

Article history:
Received 9 October 2015
Received in revised form 4 December 2015
Accepted 21 January 2016

Keywords:
Nanofuels
Nanofluids
Energetic nanomaterials
Nanoparticle agglomeration
Droplet combustion
DarrieusLandau instability
Microexplosion
Atomization
Boiling
Flameacoustic interaction
Flame topology
Multiphase multicomponent droplets
Atomization control

a b s t r a c t
With the pressing need to meet an ever-increasing energy demand, the combustion systems utilizing fossil fuels have been the major contributors to carbon footprint. As the combustion of conventional energy
resources continue to produce significant Green House gas (GHG) emissions, there is a strong emphasis to
either upgrade or find an energy-efficient eco-friendly alternative to the traditional hydrocarbon fuels.
With recent developments in nanotechnology, the ability to manufacture materials with custom tailored
properties at nanoscale has led to the discovery of a new class of high energy density fuels containing
reactive metallic nanoparticles (NPs). Due to the high reactive interfacial area and enhanced thermal
and mass transport properties of nanomaterials, the high heat of formation of these metallic fuels can
now be released rapidly, thereby saving on specific fuel consumption and hence reducing GHG emissions.
In order to examine the efficacy of nanofuels in energetic formulations, it is imperative to first study their
combustion characteristics at the droplet scale that form the fundamental building block for any combustion system utilizing liquid fuel spray. During combustion of such multiphase, multicomponent droplets,
the phenomenon of diffusional entrapment of high volatility species leads to its explosive boiling (at the
superheat limit) thereby leading to an intense internal pressure build-up. This pressure upsurge causes
droplet fragmentation either in form of a microexplosion or droplet puffing followed by atomization
(with formation of daughter droplets) featuring disruptive burning. Both these atomization modes represent primary mechanisms for extracting the high oxidation energies of metal NP additives by exposing
them to the droplet flame (with daughter droplets acting as carriers of NPs). Atomization also serves as a
natural mechanism for uniform distribution and mixing of the base fuel and enhancing burning rates
(due to increase in specific surface area through formation of smaller daughter droplets). However, the
efficiency of atomization depends on the thermo-physical properties of the base fuel, NP concentration
and type. For instance, at dense loading NP agglomeration may lead to shell formation which would sustain the pressure upsurge and hence suppress atomization thereby reducing droplet gasification rate.
Contrarily, the NPs may act as nucleation sites and aid boiling and the radiation absorption by NPs (from
the flame) may lead to enhanced burning rates. Thus, nanoadditives may have opposing effects on the
burning rate depending on the relative dominance of processes occurring at the droplet scale. The fundamental idea in this study is to: First, review different thermo-physical processes that occur globally at the
droplet and sub-droplet scale such as surface regression, shell formation due to NP agglomeration, internal boiling, atomization/NP transport to flame zone and flame acoustic interaction that occur at the droplet scale and second, understand how their interaction changes as a function of droplet size, NP type, NP
concentration and the type of base fuel. This understanding is crucial for obtaining phenomenological
insights on the combustion behavior of novel nanofluid fuels that show great promise for becoming
the next-generation fuels.
2016 Elsevier Ltd. All rights reserved.

Abbreviations: NP, nanoparticle; CNT, carbon nanotubes; FGS, functionalized graphene sheets; PLR, particle loading rate; Lel, liquid phase Lewis number; R, initial droplet
radius; Rf, flame stand-off distance; a, ejection impact parameter; DL-instability, DarrieusLandau instability; SEM, scanning electron microscope; EW, ethanolwater; N,
bubble population; N/V, bubble population density; Af, projected flame area; /, droplet void fraction; f MR , most flame-responsive excitation mode.
Corresponding author.
E-mail address: sbasu@mecheng.iisc.ernet.in (S. Basu).
1
Equal contribution by both authors.
http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.01.053
0017-9310/ 2016 Elsevier Ltd. All rights reserved.

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

483

Contents
1.

2.

3.
4.
5.

6.

Introduction and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


1.1.
Multi-component droplet combustion with and without nanoadditives. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.
Studies on combustion characteristics of nanofluid fuel droplets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Thermo-physical mechanisms governing combustion behavior of nanofluid fuel droplets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Ebullition process induced volumetric shape oscillations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Transient evaporative instability (DarrieusLandau instability). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Regimes in ebullition activity and effects of nanophase additive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Regimes in Internal liquid-phase circulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Secondary break-up and shape deformation characteristics of burning nanofuel droplets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Effect of nanoparticle concentration on localized secondary break-up modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Control of instabilities in burning droplets: role of external stimuli (preferential acoustic excitation) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.
Effect of longitudinal acoustic pulsing on the droplet deformation behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Mechanism of flame response under external harmonic forcing: The effect of longitudinal traveling pressure wave . . . . . . . . . . . . . . .
5.3.
Effect of flame response on internal ebullition dynamics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.4.
Effect of external harmonic pulsing on the structural morphology of combustion residue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction and motivation


Recent developments in the interdisciplinary field of nanofluid
technology and the pressing need for an energy-efficient, environmental friendly alternative to conventional hydrocarbon fuels has
triggered recent research in the area of novel nanofluid fuels.
Nanophase fuel suspensions or popularly known as nanofuels
are stabilized colloids containing energetic nanoscale additives
(150 nm) typically in dilute concentrations (61 vol.%) [15]. The
key idea is to harness the energy potential at nanoscale by adding
trace amounts of energetic nanometer sized metallic particles to
traditional fuels such as gasoline and diesel. In this regards, suspended NPs play role as secondary energy carriers that can be custom engineered to act as combustion enhancers. These fuels
combine together the high specific surface area of NPs and the
well-known high oxidation energies of metals to produce high
energy density fuels that save on fuel economy and hence reduce
pollutant emissions. Numerous recent investigations [630] have
reported that conventional fuels and propellants laden with
nanoparticles (NP) improve combustion performance in terms of:
(1) increased volumetric energy density, (2) enhanced catalytic
activity, (3) low ignition delay (4) higher ignition probability, (5)
higher volumetric heat release rates, (6) faster burning rates, (7)
reduction in soot and pollutant emissions, (8) greater flexibility
in developing novel fuels/ propellants with custom tailored physical properties and (9) NPs can act as gelling agents to substitute the
existing low-energy inert gellants. For instance, Sabourin et al. [8]
reported that the addition of colloidal particles of functionalized
graphene sheets (FGS22) to monopropellant nitromethane
enhanced its burning rates significantly (>1.75 times) while its
ignition temperatures were reduced. The steady-state burning rate
of nitromethane (CH3NO2) increased with increase in NP concentration and decreased with the rise in pressure thus indicating that
burning rates of colloids are less pressure sensitive compared to
pure nitromethane. Sabourin et al. [8] proposed that the observed
burning rate enhancement is through two mechanisms: (a)
increased heat transfer resulting from enhanced radiative heat
transport properties of NPs and (b) increased thermal conductivity
due to NO and FGS oxidation reactions and the catalytic effect of
FGS functional groups. Jones et al. [31] reported that the volumetric heat of combustion can be increased by 15% with the addition
of 10 vol.% n-Al to biofuel (ethanol).

483
484
486
487
489
492
492
493
494
494
496
498
498
500
501
501
502

The potential advantages of nanostructured additives are attributed to their intrinsically high reaction cross sections, high specific
surface area and high density of surface functionalities. At
nanometer length scale a particle contains thousands of atoms
and as the particle size goes below 510 nm the ratio of surface
to bulk atoms increases rapidly. For instance, for a spherical iron
crystal the ratio is 25% at 5 nm and increases to as high as 90%
at 1 nm [32]. At such high density of surface atoms the bulk material tends to display the properties of surface atoms. As an example,
NPs have lower freezing and melting point, heat of fusion [3335]
and sintering temperatures [36] compared to their micron and larger sized counterparts. A higher specific surface area at nanoscale
also allows for increased reactivity [37,38] and catalytic activity
[39]. However, there is a lower limit to which the NP size can be
reduced since with decreasing particle size, the increasing thickness of inert oxide layer now constitutes a major mass fraction of
the particle which outweighs the benefits of high energy density.
Nanofluids display enhanced thermal conductivity (that increases
with temperature and decreases with particle size), mass diffusivity and radiative heat transfer and aids faster ion transfer [4045].
Optically enhanced radiative properties and high thermal conductivity accelerate the thermal exchange process between fuel droplets and the surrounding air. For instance, a 150% augmentation
in the thermal conductivity of engine oil has been reported with
the addition of 1 vol.% carbon nanotubes (CNTs) [46]. Since NPs
are easy to disperse and added only in dilute concentrations to
the host fluid their physical properties such as surface tension, viscosity and density are not altered substantially. For example, Tanvir and Qiao [47] reported that even with the addition of 3% by wt.
NPs, the variation in surface tension of ethanol and n-decane based
fuels is negligible (less than 5%; measured using the pendant drop
method).
From an application viewpoint, recent studies have considered
the direct usage of NPs as diesel fuel additives in compression ignition engines [4860]. They demonstrated a reduction in CO, HC and
NOx exhaust emissions and ignition delay time with a corresponding increase in the brake thermal efficiency and decrease in break
specific fuel consumption. Currently the usage of nanofluid fuels in
practical combustors as combustion promoters is however quiet
restricted. Limitations arise from: (1) inherent tendency of
nanoparticles to aggregate (through mechanisms like differential
settling, perikinetic or orthokinetic particle transport) and form

484

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Fig. 1. Spatio-temporal snapshot of distinct events in a liquid fuelled combustor.

agglomerates of different types (packed, porous etc). This also


complicates the key initial step of fuel characterization, (2) the
inevitable presence of aggregates in dry nanopowders affects the
pre-ignition reaction rates since it depends strongly on both
thermo-physical properties and size distribution of NPs [61], (3)
although nanoscale metal additives (particularly aluminum and
boron) enhance reaction rates and energy density of the base fuel
their performance benefits are curtailed due to the presence of
non-energetic native oxide layers [62] and (4) the solid metal oxides formed as reaction products may be carried along with the
exhaust as residual particulate matter thereby posing a health hazard. In addition, the efficacy of future air breathing propulsion systems utilizing energetic colloids depends strongly on the ease with
which a stable nanofuel suspension can be prepared and handled
(i.e. easy availability and low cost). Although NPs generally have
good dispersibility due to their small size and mass, they tend to
agglomerate rapidly as well. Thus, the current challenge is to identify nanostructured fuel additives that not only catalyze combustion reactions but also address these physio-chemical limitations.
In order to exploit the extraordinary characteristics of NPs, they
must be transported from the suspension to the gas-phase reaction
zone and distributed homogenously along with the fuel charge
[6364]. This highlights the significance of secondary atomization
in burning nanofluid fuel droplets. This is crucial in automotive
diesel fuels where colloidal nanocatalysts (increased density of
surface functionalities) can be dispersed homogenously through
atomization to reduce ignition delay and enhance combustion
characteristics by catalyzing gas-phase reactions. Moreover, such
effective use of nanocatalysts in monopropellant and bipropellant propulsion systems may allow the use of dispensable
catalysts thereby eliminating the need for elaborate structural catalysts [65]. This aids in increasing specific thrust through use of
high energy density propellants. In colloidal droplets containing
nanostructured ignition agents, secondary break-up can play a
key role in achieving distributed ignition which markedly
improves the combustion efficiency compared to single-point ignition [66]. Thus, the secondary atomization behavior and the ensuing droplet shape oscillations mark a crucial area of investigation
involving combustion of nanofuel droplets.

Fig. 2. Chronology of key processes in a liquid fuelled combustor.

1.1. Multi-component droplet combustion with and without


nanoadditives
Spray combustion utilized in liquid fuelled combustors is a
highly complex phenomenon which involves dynamics at multiple
spatio-temporal scales. Here, the processes occurring at the droplet
level (such as liquid phase circulation, surface regression, shape
oscillations, instabilities at vaporliquid interface and group combustion) collectively affect the large scale phenomena such as
flame extinction, blow-off and pollutant formation. Fig. 1 exhibits
a snapshot of hierarchical dynamics of spray combustion where
each droplet represents a sub-grid unit of the spray. In a spray
environment, the droplet level transport strongly affects the final
combustion efficiency through a complex interplay between multiple processes like atomization, vaporization, mixing, ignition
and flame stability. This interaction is governed by a non-linear
coupling among vortical flow structures (large and small scale),
acoustics and heat release fluctuations. Fig. 2 illustrates the
spatio-temporal scales involved in a typical liquid fuelled
combustor.

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

485

Fig. 3. Mode A: High speed images of a single catastrophic micro-explosion. Mode B: low intensity continuous atomization featuring bulk droplet shape oscillations. Scale
bars equal 2 mm.

In practical spray systems, following primary atomization, the


liquid droplets are injected either with a high momentum or into
a strongly accelerating flow field. In either case, the resulting aerodynamic shear between the droplet and the flow field induces severe shape deformations which may eventually cause secondary
atomization of the droplet [6769]. This leads to an enhancement
in surface area, thereby resulting in effective vaporization and mixing. The usage of single component fuel, however, is quite rare in
practical burners and combustors. In most cases, multicomponent fuels with varying degrees of volatility differentials
among their constituent species are generally used (like Jet A).
During combustion of such multicomponent fuel droplets, the diffusion flame enveloping the droplet provides the heat required for
vaporization. However, differential vapor pressures leads to preferential entrapment of high volatile species within the droplet core
leading to its homogenous boiling and internal pressure surge
[70]. This can either lead to a single micro-explosion or force
non-periodic bulk shape oscillations on the droplet that sustain
throughout its lifetime. Such bimodal droplet disruption behavior
is characterized by continuous, multiple ejections of nucleated
micro-bubbles within the droplet as burning advances (Fig. 3 Mode
B). Note that the bubble ejection rate during atomization is a complex function of diffusional resistance (i.e. the liquid phase Lewis
number Lel), volatility differential and relative concentration of
component species [71]. Such fuel droplets (following primary
atomization) on ignition undergo violent shape oscillations leading
to frequent local fragmentations to generate a continuous stream
of micron sized satellite droplets. These small droplets provide
enhanced surface area for gasification, thereby aiding in homogenizing of the fuel charge and thus promoting clean and efficient
combustion (Fig. 3; Mode B). In some cases, when the degree of
volatility between the individual components is extremely high,
catastrophic breakup due to micro-explosion can occur (Fig. 3;
Mode A). Therefore, droplet level shape oscillations and mechanisms responsible for the same are crucial towards innovating
superior liquid fuel based combustion strategies.
Few earlier studies have reported severe droplet disruption in
both miscible binary mixtures and water-in-n-alkane emulsions
[7274]. Interfacial thermodynamic analyses revealed that the
bubbles primarily initiate at the water-hydrocarbon interface and
the severity of the disruptions strongly depends on the superheat
limit. Observations indicate that two states are possible. One state
involves swarms of micro-bubbles which coalesce, grow and eject

Fig. 4. Five distinct stages during combustion of stabilized n-C10H22/nano-Al


dispersion droplet (10 wt.% nano-Al and 2.5 wt.% surfactant) demarcated based
on the temporal variation of droplet size and temperature. Reproduced with
permission from Ref. [76].

randomly across the droplet surface. The other state involves the
formation of a single bubble which grows and exhibits (DarrieusLandau) instability at the evaporative liquidvapor interface.
Nanofuels exhibit distinct behavior with respect to bubble
inception, shape oscillations and generation of satellite droplets.
The droplet gasification effected by flame envelope and the connecting internal flow field lead to particle aggregation through a
combination of perikinetic, orthokinetic and differential settling
mechanisms. This agglomeration process alters the ebullition process in burning nanofluid fuel droplets. The micro-scale agglomerates act as preferential nucleation sites for bubble inception.
Additionally, the dispersed nanoparticles can also alter droplet
oscillation characteristics through changes in properties like viscosity and surface tension. Droplet oscillations have also been
observed in precursor droplets (with dissolved salts) due to chemical reactions but in a non-combusting environment. For example,
our recent study on vaporizing cerium nitrate precursor droplets
[75] showed that chemical transformation results in in-situ production of nanoceria and gaseous NOx. The production of NOx leads
to pressure rise which forces the droplet into a swell-shrink cycle.

486

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Fig. 5. Three stages in n-C10H22/micro-Al dispersion droplets. Reproduced with


permission from Ref. [76].

Subsequent section offers insights on bubbledroplet interactions and various types of self-excited instabilities in burning droplets. Particularly newer blends of fuels can be engineered to
undergo preferential instabilities leading to homogeneous combustion and higher efficiency. Understanding instabilities in a
burning droplet, the nature of internal two-phase circulation and
the multiscale mechanisms responsible for droplet shape oscillations forms a basis for engineering novel fuels. In this review article, we will focus exclusively on the droplet scale transport with
and without nanoparticles. This offers an exclusive option to
understand and design the fundamental building block of spray
combustion.
1.2. Studies on combustion characteristics of nanofluid fuel droplets
Although the potential benefits of energetic nanoscale additives
as combustion promoters is well established the combustion of
nanofuels at the droplet scale has received little attention. Literature search reveals that there are only handful studies that have
investigated the combustion behavior of NP laden fuel droplets.
For instance, Tyagi et al. [7] reported that the hot-plate ignition
probability of NP laden diesel droplets (in the range 688768 C)
is significantly higher than that of pure diesel droplets. Specifically,
the ignition probability is markedly higher for both n-Al and
n-Al2O3 droplets compared to pure diesel. However, the enhancement in ignition probability is not correlated with the variation
in particle size or type.
In a seminal study, Gan et al. [76] compared the combustion
behavior of millimeter sized droplets containing micron and
nanoscale Al additive. They observed that stabilized n-C10H22/
nano-Al dispersion droplets undergo five distinct combustion
stages: preheating, classical combustion, l-explosion, surfactant
flame and aluminum droplet flame (Fig. 4). In contrast, stabilized
n-C10H22/micro-Al dispersion droplets underwent only three distinct stages: preheating, classical combustion and l-explosion
(Fig. 5). For same initial particle loading rate (PLR) and surfactant
concentration a single l-explosion occurred later and with a much
stronger intensity in l-suspensions where most particles were burned.
However, nanosuspensions underwent a series of l-explosion events
releasing NPs continuously. This finally resulted in large agglomerates whose combustion took longer time and was less complete
due to oxide shell formation. This distinct burning behavior of
nano vs. micro suspensions was due to the distinct structure of
particle agglomerates formed during combustion. Gan et al. [76]
reported that in nanosuspensions particle aggregation was
dominated by Brownian diffusion (perikinetic transport) which

resulted in a more-uniformly distributed, porous spherical shell.


Contrarily, in microsuspensions, particle collision and aggregation
was dominated by orthokinetic transport thereby resulting in a
densely packed impermeable shell.
Next, Javed et al. [77] investigated the evaporation characteristics of stabilized n-Al/n-C7H16 at elevated temperatures (100
600 C). With addition of n-Al, the evaporation rate of stabilized
heptane droplets is reduced at low ambient temperatures
(<400 C) due to formation of a compact shell of large NP aggregates while the evaporation rate is enhanced at high temperatures
(>400 C) due to formation of porous shell of small NP aggregates.
Below 400 C, the degree of suppression in evaporation increases
with PLR and above 400 C the evaporation is enhanced up to
2.5 wt.% and decays beyond 2.5% with increase in PLR. As such,
2.5 wt.% represents some sort of critical concentration. Javed
et al. [77] also reported that in general addition of Al NPs suppresses bubble formation in stabilized heptane droplets. This is
however contradictory with the results of Miglani et al. [63] where
a marked rise in bubble population was observed with the addition
of NPs as will be shown later.
Javed et al. [78] studied the autoignition and combustion characteristics of n-Al/n-C7H16 droplets at elevated temperatures (600
850 C). Key findings of this work are: (1) Both pure and NP laden
heptane droplets show ignition time delay which follows Arrhenius temperature dependence, (2) Ignition delay time varies
depending on the NP concentration, (3) at low temperatures
(600700 C), the overall activation energy at dilute PLRs (0.5 wt.
%) is lower than that of pure droplets while at dense PLRs (2.5
and 5 wt.%) it is higher. As such, no ignition is observed at dense
PLRs. At higher temperatures (750850 C), the ignition delay time
is independent of PLR and is comparable to pure and stabilized
heptane droplet. However, droplet disruption characterized by
bubble formation and rupture is still dominant at dense PLRs, (4)
at higher temperatures (750850 C) the ignition delay time is
independent of PLR and is comparable to pure and stabilized heptane droplet, (5) irrespective of PLR, n-Al/heptane droplet combustion deviates from the classical D2-law as droplets undergo
frequent microexplosions (characterized by multiple time droplet
expansion and rupture). Onset of these microexplosions is earlier
and their intensity increases with increase in temperature and
(6) these microexplosions occurring at higher temperatures result
in lower burn time and total combustion time (i.e. faster gasification rate) for NP laden droplets compared to pure heptane droplets.
A complementary experimental study [79] focused on the
autoignition and combustion characteristics of n-Al/Kerosene droplets at dilute PLRs (0.1, 0.5 and 1 wt.%). It was observed that the
autoignition delay time of both n-Al/Kerosene droplets and pure
kerosene droplets follows Arrhenius temperature dependence i.e.
the ignition delay decreases exponentially with increase in ambient temperature. Even at dilute PLRs, NP addition lowered the minimum ignition temperature of kerosene droplets to 600 C at
which the pure kerosene droplets did not ignite. Compared to pure
and stabilized kerosene droplets n-Al/Kerosene droplets undergo
disruptive burning (frequent microexplosions) at all ambient temperatures and their combustion deviates from the classical D2-law.
However, the intensity of these fragmentations is low at dilute
PLRs. With initially dilute PLR and continuous disruptions all NPs
escape from the droplet and no residue is left on the suspension
fiber. In addition, due to this continuous rupturing the combustion
rate of NP laden droplets is substantially enhanced compared to
pure kerosene droplets.
Bello et al. [80] studied the effect of addition of dilute concentrations of MgO NPs on the droplet gasification rate, surface tension and heat of combustion of RP-2 propellant. They reported
that even at extremely dilute PLR (0.5 wt.%), the droplet surface
regression rate increases by two orders of magnitude. They further

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

illustrated that NPs augment the diffusive heat transport thereby


promoting bubble nucleation and subsequent droplet atomization.
Such droplet disruption sustains at all three stages of droplet combustion, namely droplet evaporation, combustion of RP-2 and
burning of stabilizer (Oleic acid). Besides, the dispersant (OA)
was observed to have no effect on the droplet burning rate
constant.
Javed et al. [77,81] also carried out detailed analyses of
microexplosions. They studied the microexplosion behavior of
n-Al/kerosene droplets with dense PLRs (2.5, 5 and 7 wt.%) and at
elevated temperatures (400800 C). They found that droplet
disruption characterized by bubble formation and rupture is
dominant at dense PLRs. They also investigated the evaporation
characteristics of stabilized n-Al/n-C7H16 at elevated temperatures
(100600 C). They found that irrespective of PLR, n-Al/heptane
droplet combustion deviates from the classical D2-law as droplets
undergo frequent microexplosions (characterized by multiple time
droplet expansion and rupture). Onset of these microexplosions is
earlier and their intensity increases with increase in temperature
and these microexplosions occurring at higher temperatures result
in lower burn time and total combustion time (i.e. faster gasification rate) for NP laden droplets compared to pure heptane droplets.
Based on the current literature it is noteworthy to consider
following points. First, although many studies [7679,80] have
reported the occurrence of boiling inside nanofluid fuel droplets
which leads to disruptive droplet burning, there is no clear evidence that such bubble formation occurs due to the addition of
NPs or does the support fiber serve as a source of bubble nucleation
sites. In fact, it is extremely difficult to image the bubble nucleation
process dynamically from a nanoscale nucleation site inside a millimeter sized droplet that burns within a short timescale (12 s).
Secondly, in most studies the burning rate constant is calculated
from the slope of experimental curve of (D/Do)2 vs. t/Do2 based on
a linear regression fit. This inherently assumes that the surface of
a burning nanofuel droplet retracts steadily in accordance with
the classical D2-Law. However, as the droplet undergoes aperiodic
inflation and deflation due to atomization and with agglomeration
of NPs impeding its gasification, the droplet may not follow the
D2-Law. Also, a disruptive event such as bubble ejection features
gas liberation and DD (daughter droplet) formation which essentially decouples the gas and liquid phase mass from the parent droplet. In cases of disruptive combustion, the burning rate constant
can calculated accurately based on the flame-termination method
described by [82].
Third, technically, microexplosion in reference to water/oil
emulsions and multicomponent liquids corresponds to a sudden
catastrophic break-up of the droplet due to explosive homogenous
boiling (which occurs internally due to the phenomenon of diffusional entrapment) and leads to disruptive burning. However, in
burning pendant droplets the suspension fiber acts as a nucleating
surface and promotes heterogeneous boiling. It is well-known that
heterogeneous boiling is associated with low degree of superheat
and aids continuous bubble formation thereby resulting in a large
bubble population (depending on the number of active nucleation
sites). This is corroborated from studies by Miglani et al. [63,64]
where a rapid rise in the bubble population was observed at early
stages (first 10%) of the droplet lifetime. In such a case where a
nucleating surface is present and homogenous boiling doesnt
occur the droplet will not undergo high intense break-up but
would rather exhibit sustained bulk shape oscillations due to
low-intensity atomization events. Consequently, it is apt to refer
to droplet disruption events as puffing followed by atomization
and not as microexplosions. Literature search reveals that there
is a dearth of information on the combustion behavior of
nanofluid fuel droplets and only a handful of research groups [7,
63,64,7680,8183,90,91] have attempted to explore the

487

underlying mechanisms at the droplet scale. Table 1 summarizes


the details of these studies. In fact only a few of these studies have
delved into the instabilities in burning nanofuel droplets in an
attempt to characterize the different mechanisms that govern
burning rate, secondary atomization and structure of the final precipitate. In the next few sections, we have attempted to review and
provide insights into the internal small scale dynamics of burning
droplets.

2. Thermo-physical mechanisms governing combustion


behavior of nanofluid fuel droplets
To this point it is well-established that boiling inside a burning
nanofuel droplet represents an unstable state where the droplet is
prone to atomization due to internal pressure build-up. This pressure up rise is a resultant of two complementary processes that aid
an increase in the vapor phase volume. First, the active mode of
bubble growth, which occurs directly due to continuous heat input
from the droplet diffusion flame and indirectly through bubble
coalescence events. Second, the passive mode features droplet surface regression due to the depletion of liquid solvent (through its
gasification). Note that the latter is a slow process that occurs over
a large timescale 12 s while the former mode is a short (submillisecond) timescale process and hence is the key mechanism
governing pressure build-up. However, with the onset of atomization events, the formation of daughter droplets (DD) and the coupled surface corrugations enhance the burning rate thereby
accelerating the latter process. In addition to faster surface regression, the bubble collapse during an atomization event internally
agitates the surrounding liquid. With several such events occurring
simultaneously there is strong churning in the liquid phase. Since
the bubbles are advected along with liquid phase circulation, the
chances of bubble coalescence increase substantially. On the other
hand, externally, the incipient DDs travel to the flame front and
distort the primary droplet flame thereby altering the heat input
to the droplet. This disruption in the flame topology serves as a
feedback for the parent droplet since it affects both the bubble
growth and droplet gasification rate. In this regards, ebullition
activity inside the droplet domain, atomization across the droplet
surface and diffusion flame enveloping the droplet are closely coupled in a feedback loop. Hence, it is interesting to review how the
addition of NPs at different concentrations would modulate this
coupling and affect the fundamental parameters of interest,
namely, the droplet burning rate and the flame stand-off distance.
Note that although the flame stands-off at a spatially distinct location it interacts with the first two processes via the DDs. These DDs
are transported to the flame front either by the buoyant convective
field or an inherent momentum imparted during the atomization
process or a combination of the two.
This section focuses on three potential modes of bubbledroplet
interaction in burning bi-component droplets (with and without
nanoparticles) following the work by Miglani et al. [63]. The first
mode features bubble ejections at the vaporliquid interface
thereby resulting in significant surface corrugations (termed as
minor ejections). Second intermittent mode exhibits rapid collapse
of large bubble volume (major ejection) that induces large scale
surface deformations and bulk shift in the centroid. The third mode
is mainly due to an evaporative DarrieusLandau instability at the
vaporliquid interface that corrugates the droplet surface at a
specific wavenumber. Due to these bubble ejections the droplet
internal recirculation is significantly altered thereby leading to
an intense mixing of chemical species which in-turn affects the
droplet gasification rate. Particularly, the oscillatory characteristics
of burning functional droplets are discussed for two test fuels: (1)
Nanoceria suspension (10 nm; 0.115 vol.% or 1 wt.% i.e. at dilute

488

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Table 1
Nanofluid fuel droplet studies reported in current literature.
References

Nanoparticle type, nominal size (nm) and Base fluid


particle loading rate (PLR)
Chemical
Surfactant (S)

Combustion characteristic studied


Ambient conditions (Temperature and Pressure)

Tyagi et al. [7]

Al (50 nm + 1.5 nm passive oxide layer)


Al2O3 (15 and 50 nm)

Diesel
No Surfactant

PLR: 0, 0.1 and 0.5 vol.%


Al

C2H5OH

Ignition probability (IP)


T: Hot-plate temperature range 688768 C
(raised in increments of 20 C)
P: 0.1 MPa
Comparison of particle agglomeration and microexplosion
behavior of micro and nanosuspensions

Gan et al. [76]

No surfactant
n-C10H22
S: Sorbitan Trioleate
Gan et al. [90]

Javed et al. [77]

B (80 nm)

C2H5OH

Fe (25 nm)

No surfactant
n-C10H22
S: Sorbitan Oleate
n-C7H16

Al (70 nm with 23 nm passivated oxide


layer)
PLR: 0.5, 2.5 and 5 wt.%

Effect of particle type, PLR and base fluid on overall


combustion behavior and microexplosion

Evaporation at elevated temperatures and microexplosion behavior

S: Sorbitan Trioleate T: 100600 C


(Span 85: C60H108O8)
P: 0.1 MPa

Javed et al. [91]

Al- Ligand protected (70 nm)


PLR: 0.1, 0.5 and 1 wt.%

Kerosene
S: Oleic acid

Evaporation at elevated temperatures and microexplosion behavior


T: 400800 C
P: 0.1 MPa

Javed et al. [81]

Ligand protected Al (2.5, 5 and 7.5 wt.%)

Kerosene
S: Oleic acid

Effect of PLR and ambient temperature on microexplosion behavior

Javed et al. [78]

Ligand protected Al (70 nm)

n-C7H16

PLR: 0.5, 2.5 and 5 wt.%

S: Oleic acid

T: 400800 C
P: 0.1 MPa
Effect of particle loading rate on autoignition properties
at elevated temperatures
T: 600850 C
P: 0.1 MPa

Javed et al. [79]

Ligand protected Al (70 nm)


PLR: 0.1, 0.5 and 1 wt.%

Kerosene
S: Oleic acid

Autoignition properties at elevated temperatures


T: 400800 C
P: 0.1 MPa

Miglani et al. [63]

CeO2 (10 nm)


PLR: 1 wt.%

C2H5OH/H2O

Internal ebullition dynamics


Mechanism driving secondary atomization in nanofluid fuel droplets
Regimes in internal flow circulation

Miglani et al. [64]

Al2O3 (20 nm)


PLR: 10 wt.%

C2H5OH/H2O

Secondary break-up control using preferential acoustic excitation

Miglani et al. [83]

Anatase TiO2 (48 nm)


PLR: 0.5, 1, 2.5, 7.5 wt.%
Anatase TiO2 (48 nm)

C2H5OH/H2O

Effect of PLR on the secondary break-up modes

C2H5OH/H2O

Feedback coupling between two key mechanisms occurring


globally at the droplet scale:
(a) Atomization frequency
(b) Shell formation by NP agglomeration

RP-2 (Rocket
propellant grade 2)

The effect of NP loading on droplets surface regression rate,


surface tension and calorific value

Miglani et.al [100]

PLR: 0.05, 0.5, 1, 2.5, 5, 7.5 and 10 wt.%


Bello et al. [80]

MgO (20 nm)


Surface-treated with long chain carboxylic
acid

S: Oleic acid (OA)


PLR: 0.05, 0.25, 0.5, 0.75 and 1 wt.%
Tanvir and Qiao [99] Al

C2H5OH
No surfactant

Effect of droplet size on the burning rate constant

Toluenediethyl
ether fuel mixture

Effect of soluble molecular [AlBrNEt3]4 cluster additive


compared to its pure nano-Al counterpart on the burning rate constant and
microexplosive characteristics of sub-millimeter sized droplets (650 lm)
Room temperature oxygen environment in a drop tower

PLR: 0.15 wt.%


Guerieri et al. [82]

[AlBrN(C2H5)3]4 Al (I) Tetrameric cluster


with 01.6 wt.% active aluminum content

PLR) with ethanolwater blend as the base fuel and (2) a pure ethanolwater blend. However, the conclusions are generic and applicable to most droplet systems containing trace amounts of
nanoparticles.
Both ethanolwater and ethanolwater with nanoceria droplets
(pendant mode) burn in natural convection mode with a teardrop

shaped flame envelope. The droplet is surrounded by a quasisteady diffusion flame showing a maximum fluctuation of 6% in
the flame stand-off distance Rf. In nanoceria laden droplets the
satellite droplets formed during atomization events serve an additional purpose of transporting NPs/particle aggregates to the flame
front where they burn with a characteristic reddish-yellow glow

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

489

Fig. 6. Phenomenology of single droplet combustion (with and without nanoparticles) (a) Time frozen snapshots of nanofuel droplet flame. Dotted circles indicate streaking
of micro-scale particle aggregates. These are transported to the flame envelope via fragmented daughter droplets (DD). DDs are formed during localized, secondary break-up
of the parent droplet following bubble ejections and act as parcels or carriers for NP transport to the flame. (b) Time frozen images of bi-component (ethanolwater) droplet
flame without nanoadditives. (c) Instantaneous images of deformed droplet: acquired at 15,000 fps at spatial resolution of 3.9 lm/pixel. (d) Temporal variation of the flame
stand-off distance (Rf) for droplet flame envelope is shown alongside [63]. Reprinted with permission from [Ankur Miglani, Saptarshi Basu and Ranganathan Kumar, Insight
into instabilities in burning droplets, Physics of fluids 26, 032101 (2014)]. Copyright [2014], AIP Publishing LLC.

(Fig. 6). Also, the droplet gasification rate reduces with increase in
nanoparticle concentration due to the formation of a gelatinous
shell (towards the end of droplet lifecycle) that inhibits liquid
phase diffusion. The surface regression timescale for a 2 mm fuel
droplet usually varies from 12 s. However, the timescale of droplet disruption/atomization events (shown later) is much shorter
compared to the droplet gasification rate. Fig. 6 illustrates the phenomenology of single droplet combustion.
2.1. Ebullition process induced volumetric shape oscillations
Miglani et al. [63] first analyzed in details the ebullition process
inside millimetre sized nanofluid fuel droplets burning in pendant
mode. They found that concurrent, multiple ejections of microbubbles are triggered shortly after their incipience causing
localized fragmentation of the primary droplet. These localized
fragmentations produce a cloud of daughter droplets surrounding
the parent droplet similar to a sputtering effect. Such bubble
expulsions are persistent throughout the droplet lifecycle and
are characterized by a timescale of about O (2 ms) and discharge
velocities of the order of 5 m/s. These are termed as minor ejections. The frequency and intensity of such ejections however varies
as the droplet lifetime progresses. Moreover, as the atomization
process is three dimensional and random in nature the experimental runs are stochastic. Consequently, the droplet shape oscillations
may not be solvable by standard modal analysis methods in order
to resolve different deformation modes. In this light, droplet deformation may be understood based on certain physical parameters.
For instance, Miglani et al. [63] proposed that apart from the extent
of accumulated superheat the relative ratio of evacuated vapor
space and the instantaneous droplet volume during ejection is
primarily responsible for volumetric droplet-shape deformation.
This was quantified by an ejection impact parameter [63,64,83]

aglobal t

n
X
p
j1

3
dj;eject

!,
p
6


D3eject ;

0 6 aglobal t < 1

where numerator represents the total vapor phase volume that collapses in the droplet interior during n simultaneous expulsion
events at a given instant and dj, D denote projected area diameters
of jth bubble and parent droplet respectively at the pre-ejection
instant.
The ejection parameter a is akin to a deformation index which
quantifies the extent of distortion in the overall droplet geometry
following bubble expulsions. aglobal represents a cumulative effect
of several expulsion events at any instant of the droplet lifetime.
Similarly alocal t represents a relevant parameter for localized single expulsion event that has maximum deformation inducing
potential at any given instant. Further it was shown that aglobal
and alocal vary directly with the bubble ejection frequency (n per
unit time), statistical bubble size distribution and inversely with
the droplet surface regression rate.
Miglani et al. [63] defined another parameter termed as droplet
void fraction (/) that measures the fraction of instantaneous droplet volume, V, that is occupied by the bubbles. This represents
the combined contribution of the bubble crowd density (i.e. n/V)
and the bubble size distribution towards instantaneous vapor
space volume and is given by:

/t

N
X
3
di =D3

i1

where, D and di are the projected area diameters of primary droplet


and individual bubbles respectively while N is the bubble population. Physically, / it is an indirect measure of the extent of pressure
accumulation within the droplet and thus it determines the droplets susceptibility to undergo an ejection at a given time instant.
Large / ? 1 is an indirect indicator of high pressure build-up within
the droplet and hence signifies an unstable droplet state with aggravated probability of undergoing a high intensity ejection. Therefore,
/ represents a kind of secondary, deformation susceptibility index
[63,64,83]. The ejection impact parameter (a) being the primary
deformation index. Clearly, as the bubble growth is supported by a
continuous heat input from the flame while the droplet volume

490

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Fig. 7. Dynamic shape deformation during a major bubble ejection. (a) Post coalescence bubble growth and droplet swelling, (b) collapse of vapor bubble space and
volumetric droplet shrinkage, (c) formation of standing liquid jet and droplet shape transition to lenticular geometry, (d) post-ejection first axis-switching and (e) postejection second axis-switching. Symbols tc and ts denote the time periods of constant and stagnated bubble growth respectively. The scale bar equals 1 mm [63]. Reprinted
with permission from [Ankur Miglani, Saptarshi Basu and Ranganathan Kumar, Insight into instabilities in burning droplets, Physics of fluids 26, 032101 (2014)]. Copyright
[2014], AIP Publishing LLC.

Fig. 8. Temporal variation of relative circumference enhancement C  t pC  1


2

pAp

during a major bubble ejection with alocal, max = 0.5 (From Fig. 7) showing that the
droplet undergoes a continuous swell-shrink cycle (i.e. puffing, de-puffing) (a): preejection small scale surface undulations, (b): pre-ejection droplet swelling, (c):
bubble collapse and ejection, (d):post-ejection shrinkage and (e): recovery period
characterized by continuous swelling and shrinking as droplet restores its
geometry.

decreases with time / is expected to increase continuously during


droplet lifetime.
High values of a (i.e. a ? 1) are well correlated with major bubble ejection events in which the discharged vapor volume is 0.5
0.6 times the instantaneous droplet volume. The volumetric shape
oscillations in droplets following such large ejections are mainly
due to the recoil thrust. These concurrent ejections lead to large
amplitude, low wavenumber surface perturbations that are comparable to the instantaneous droplet diameter. During such major
ejection the droplet loses its sphericity momentarily (O(2 ms)) as
its operational circularity drops to about 0.7. The directional nature
of the bubble ejections manifests as rapid axis-switching and reorientation of the droplet within a very short time as shown in Fig. 8.
On the other hand, a  1 signifies minor ejection events (low/
intermediate intensity) which initiate small-scale (mild) perturbations on the droplet surface. High-speed images of major and
minor bubble ejections (alocal = aglobal = 0.5, Fig. 7 and alocal = 0.057,
Fig. 10) illustrate an interesting snapshot of the pre-ejection bubble coalescence-growth, subsequent bubble collapse time-scales
and the resulting droplet shape deformation. Miglani et al. [63]
observed that, while the pre-ejection bubble residence time is
two orders of magnitude lower for the minor bubble ejection, the
time-scale and average velocities are of the same order i.e. O
(1 ms) and O(5 m/s) respectively for both types of ejections

Fig. 9. Critical length scales for estimating volumetric droplet shape oscillations.

491

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Fig. 10. Dynamic surface deformation during a minor bubble ejection. (a) bubble coalescence, (b) lamella dome formation, (c) liquid sheet rupture and gas escape from within
the bubble, (d) formation of surface crater, (e) depression fill-up and inception of standing liquid jet (ligament), (f) ligament growth and upward transport due to surface
disturbance from adjacent minor ejection event, (g) onset of ligament break-up and (h) ligament necking and daughter droplet pinch-off. For the image sequence snapshots
are at time instants (a) 0, 0.2 and 0.47 ms, (b) 0.53 ms, (c) 0.6 and 0.67 ms, (d) 0.73 ms, (e) 0.87 ms, (f) 1.33 ms, (g) 1.47 ms and (h) 1.6, 1.67, 1.8 and 2 ms. The local ejection
impact parameter for this event is 0.057. The Scale bar equals 1 mm [49]. Reprinted with permission from [Ankur Miglani, Saptarshi Basu and Ranganathan Kumar, Insight
into instabilities in burning droplets, Physics of fluids 26, 032101 (2014)]. Copyright [2014], AIP Publishing LLC.

Fig. 11. Regimes in ebullition activity based on normalized bubble population and
droplet void fraction (shown inset). Dotted lines with blue markers: nanoceria
dispersion droplet. Dotted lines with black markers: for ethanolwater droplet
(without nanoadditives). Symbols I, II, III and IV denote the regimes is droplet
deformation cycle based on the maximum, local ejection impact parameter. Time is
normalized with the total droplet lifetime (tl) [63]. Reprinted with permission
from [Ankur Miglani, Saptarshi Basu and Ranganathan Kumar, Insight into
instabilities in burning droplets, Physics of fluids 26, 032101 (2014)]. Copyright
[2014], AIP Publishing LLC. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

(a  1 and a ? 1). aglobal and alocal magnitudes are also expected to


increase with time (since the droplet surface regresses and the
mean bubble diameters grow continuously with time). This implies
that volumetric oscillations will also grow as the droplet proceeds
towards extinction.
Critical geometrical parameters that may be used to quantify
the degree of deformation in the droplet shape are marked in
Fig. 9. Based on Fig. 9, the relative increase in circumference and
augmentation in droplet surface area (due to surface wrinkling)
caused by bubble ejections are calculated as:

C
Relativ e circumference increase C  t p  1 C C C 
2 p Ap
3

Fig. 12. Two-part dynamic growth behavior of a major bubble during a single
isolated bubble regime, prior to ejection [63]. Reprinted with permission from
[Ankur Miglani, Saptarshi Basu and Ranganathan Kumar, Insight into instabilities
in burning droplets, Physics of fluids 26, 032101 (2014)]. Copyright [2014], AIP
Publishing LLC.

Surface area enlargement factor A t


2

Ac
Ap

where, Ac 4Cp is the developed (actual) surface area or circumference (C) equivalent projected area and Ap is the actual projected
area. Fluctuations in circumference C and C  are marked in Fig. 9.
Low values of alocal (Fig. 10) cause droplet surface wrinkling as
described earlier and is similar to the mechanism of bursting gas
bubbles at the liquid free surface as reported by Newitt et al.
[84]. Miglani et al. [63] demarcated four distinct regimes in the
droplet deformation cycle based on the intensity of bubble ejections that correspond to the regimes in ebullition activity. As illustrated in Fig. 11 these regimes are (I) bubble nucleation, (II) onset
of minor ejections (alocal, max  0.1); rapid rise in bubble population
but with slow bubble growth rate, (III) intermediate intensity ejections (0.1 < alocal,max < 0.3); decay in bubble population and (IV)
high intensity major ejections (alocal, max > 0.3); rapid growth in
bubble sizes with multiple coalescence events [6364,83].

492

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

2.2. Transient evaporative instability (DarrieusLandau instability)


In any burning droplet, several bubbles can coalesce into a single large bubble. Such events occur randomly and require statistically significant number of experimental observations. Miglani
et al. [63] found that in the pre-ejection phase a single isolated
bubble exhibits a distinct two-part growth behavior (Fig. 12). In
the first part, the evaporative front (bubbleliquid interface) grows
rapidly leading (89 mm/s) to swelling of the droplet (1520%).
In the second part, the bubble growth decays and ultimately stagnates. In the first part of the growth process they deciphered the
existence of evaporative instability at the dropletbubble interface.
Measurement of surface fluctuations at various angular locations
across the bubbledroplet surfaces illustrated an in-phase synchronized oscillation with a dominant frequency of around
30 Hz. The wrinkling of the bubble surface due to evaporative
instability waves is transferred to the droplet surface instantaneously. The dropletbubble hence acts as a coupled oscillator.
This self-excited instability is primarily due to Landau mechanism
described originally in reference to laminar flames [85] and subsequently modified by several investigators for unstable explosive
boiling of droplets at the superheat limit [8689]. The maximum
wavenumber and growth rate of interfacial evaporative instability
can be calculated from LandauDarrieus theory [85].

kmax 2ql v 2 =3er

xmax 2ql v 3 =3r3e3=2

where ql, m, r denote the liquid density, bubble growth velocity and
interfacial tension at the vaporliquid interface respectively and
e = qv/ql where qv is the vapor bubble density. The calculations corroborated well with the experimentally determined values of smallest surface deformation length scales and growth rate.
2.3. Regimes in ebullition activity and effects of nanophase additive
To this point it is clear that the ejection impact parameter (a) and
droplet void fraction (/) are the key factors governing droplet
deformation and oscillations in volume. For bi-component droplets, the first and second global modes (described previously) play
dominant roles in determining droplet distortion dynamics. However in burning nanofluid fuel droplets, aggregation due to perikinetics and orthokinetics lead to the formation of a semi-solid
porous network of weakly bonded agglomerates (microns in length
scale). The porous crust significantly inhibits internal recirculation
and liquid phase diffusion thereby decreasing the gasification rate.
This leads to a reduction in a as can be seen by a consistently lower
value of alocal, max in case of a nanofuel droplet. Furthermore, the
gelatinous cocoon structure behaves like a low-stiffness elastic
membrane that can resist the pressure build-up due to the formation and growth of bubbles inside the liquid core. Thus, the formation of gelatinous shell promotes bubble entrapment which causes
a drastic reduction in bubble ejection frequency thereby leading to
damped oscillations. Miglani et al. [63] showed (Fig. 11) that bubble population in burning nanoceria laden droplets is consistently
higher than conventional ethanolwater droplets. During regime II,
the rate of growth of bubble population in nanoceria droplets is
significantly higher compared to EW droplet and there is a divergence in the bubble population curve; leading to a large off-set
at the peak (Fig. 11). This suggests that nanophase additives do
act as potential nucleation sites even at very dilute concentrations
(<1 vol.%). This is a key finding from their work indicates that the
dispersed NPs act as heterogeneous nucleation sites under volumetric heating by the surrounding flame envelope. In a scenario
where homogenous boiling could be achieved in burning

Fig. 13. Bubbles entrapped in nanoparticle laden droplets. Reprinted with


permission from [Ankur Miglani, Saptarshi Basu and Ranganathan Kumar, Insight
into instabilities in burning droplets, Physics of fluids 26, 032101 (2014)].
Copyright [2014], AIP Publishing LLC.

multicomponent droplets in a contact free environment (levitation)


it would be worthwhile to investigate if nanoadditives can act as
nucleation sites and initiate a transition from homogenous to
heterogeneous boiling. However increase in bubble population
does not translate into high intensity bubble ejections. This is
because NP consolidation (gelatinous shell) at the receding droplet
surface renders it high surface tension which must now be
overcome to initiate an ejection.
Fig. 11 illustrates that nucleated bubbles show a much higher
propensity to coalesce as bubble population rises rapidly in regime
II which leads to increased ejection events. The bubble population
peaks around 20% of the droplet lifetime before exhibiting a sharp
decline (Regime III) mainly due to increased coalescence. Regime
III therefore shows increased droplet void fraction but reduced
number of bubbles. Regime II on the contrary showcases increase
in both bubble population and droplet void fraction. Miglani
et al. [63] reported that for traditional ethanolwater (EW) droplets, these inverse trends continue to diverge for rest of the droplet lifecycle with final regime IV being characterized by
multiple cascades of rapidly occurring bubble merging events. Formation of large bubbles acts as precursors to intense internal pressure upsurge and subsequent major ejections. The wavenumber of
the high intensity ejections scales roughly as the droplet size
(1 mm) with spatial shifting of the droplet centroid. Contrarily,
in nanoparticle laden droplets, the viscous gel type crust aids bubble entrapment (Fig. 13) thereby leading to stagnation in bubble
population (as seen by flattening in the curve; Fig. 11).
It has been hypothesized by Miglani et al. [63] that consistent
increase in / coupled with droplet diameter regression increases
the probability of high a events particularly near the extinction
of the droplet. This also corroborates the observations that functional droplets show increased susceptibility to self-excited bulk
shape oscillations at later stages of their lifecycle. Fig. 14 summarizes the dynamics of different global modes and stages of a burning bi-component droplet along with the ebullition dynamics. It is
noteworthy to mention here that the three modes (minor and
major bubble ejections and DL instability) discussed here represent
only a broad classification containing a whole range of bubble ejection events that occur during the droplet lifecycle. Specifically, the
minor and major ejections cover an entire range with a ranging
from a = 0 (pure droplet vaporization without internal boiling) to
a ? 1 (a single microexplosion event).

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

493

Fig. 14. Schematic showing the summary of different regimes and key findings. Blue bars indicate the timescales of various events [63]. (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article.) Reprinted with permission from [Ankur Miglani, Saptarshi Basu and Ranganathan
Kumar,Insight into instabilities in burning droplets, Physics of fluids 26, 032101 (2014)]. Copyright [2014], AIP Publishing LLC. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)

3. Regimes in Internal liquid-phase circulation


It is interesting to note that internal circulation pattern in burning functional droplets undergoes a continuous transformation as
droplet lifetime progresses. Typical internal flow regimes range
from a well behaved poloidal motion at initial stages to a completely disordered two-phase bubbly flow at later stages. In
nanofuels, the flow in addition is highly damped due to increased
liquid core viscosity as NP concentration increases (Fig. 14).
First occurrence of torroidal vortex is observed during the initial
 3


heat-up period of the droplet DD


< 0:05; Dt  0:35tl . The maxiD3
o

mum Reynolds number is estimated to be around 2030 during the

initial vaporization period (Dt  0.35tl). At later stages, Miglani


et al. [63] reported intermittent oscillatory behavior of the torroidal structure with small amplitude circumferential oscillations
and escalated velocity along the central axis of the droplet (90
150 mm/s). The initial vortical state is characterized by multiple
minor bubble ejections and slight displacement of the centroid in
the horizontal direction. The droplet sphericity is roughly within
2% as indicated by Miglani et al. [63]. The well behaved flow
structure completely collapses during later time instants
(t > 0.3tl) with increased intensity of the bubble ejections and recoil
of the droplet mass. The torroidal vortex transitions through a series of deformations and reorientations before settling into a solid
body rotation along an inclined axis. At late stages of the droplet

494

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

lifetime, with coalescence and significant ejections of bubbles, the


solid body rotation further disintegrates into a chaotic flow pattern. Miglani et al. [63] hypothesized that bubble ejections act as
micro-pumps which churn the liquid in various arbitrary directions leading to an oscillatory flow field. Bulk convection in the liquid is however controlled by large scale perturbations of the
droplet surface and cumulative effects of viscous shear and gas
expansion at the liquidgas interface. A thumbnail of the internal
recirculation patterns is shown in Fig. 14.
4. Secondary break-up and shape deformation characteristics of
burning nanofuel droplets
In the previous section, we described in details different global
modes, ebullition regimes and flow maps observed during combustion of bi-component EW droplet and nanoparticle laden functional droplets. However the results presented were only valid
for extremely dilute concentration and may not necessarily be
identical when the particle loading is increased to very high values.
Combustion behavior of nanofluid fuel droplets is controlled by
non-linear coupled transport processes at multiple spatiotemporal scales. Recent studies [7,63,64,7680,8183,90,91] have
clearly illustrated that particle aggregation and structural morphology control the bubble ejections, micro-explosion and combustion characteristics in nanofuels. These studies have
concluded that microexplosion and combustion dynamics of
nanofuels are significantly controlled by particle aggregation rate
and topology of the agglomerates. In addition, droplet deformation
mechanics is strongly influenced by particle loading concentration
and stability of the suspension. In the section, we review secondary
atomization pathways in burning nanofluid fuel droplets for varied
NP concentration (from dilute to dense limits) following the work
of Miglani et al. [83].
As discussed in the previous sections, self-induced boiling in
burning multicomponent droplets lead to bubble expulsion events
featuring a sequential processes of bubble collapse, ligament formation and ligament breakup into daughter droplets. These
pinched-off daughter droplets act as carriers or parcels of nanoparticles from the liquid core to the flame front. Bubble ejection is thus
the key mechanism for transporting nanoparticles to the droplet
flame in order to burn them and extract their high oxidation energies. At the flame front these daughter droplets are ignited and
burned thereby releasing the micro-agglomerates/NPs which
results in local flames that disrupt the primary droplet flame. Thus,
it is crucial to understand the cumulative effects of different
atomization modes (at varying NP loading) on the global flame
response.
4.1. Effect of nanoparticle concentration on localized secondary breakup modes
Figs. 15 and 16 showcase the break-up modes of nanotitania
droplet at dilute (0.5, 1 wt%) and dense (5 and 7.5 wt%) PLRs along
with the corresponding high speed images depicting flame
response. As can be seen, there are significant differences in the
mode of breakup based on the initial PLR.
Miglani et al. [83] demarcated five distinct disruptive modes
responsible for bulk oscillations in the droplet and consequent
flame perturbations. These modes in the increasing order of
severity were classified as (1) Needle-type ligament ejections
(alocal  0.01); /  0; high kinetic energy) (2) Needle ejection with
tip-base break-up (alocal  0.1), /  0 (3) low momentum needletype ligaments with only tip break-up (0.1 < alocal < 0.3); 0.1 < /
< 0.2, (4) low momentum, thick ligaments (alocal > 0.3); / > 0.3
and (5) localized catastrophic fragmentation with multiple

ligament formation (alocal ? 1); / ? 1. It was observed that modes


13 (low intensity modes) are generally present throughout the
droplet lifecycle for dilute PLRs while high intensity modes (4
and 5) control the combustion dynamics in droplets with dense
PLR.
Following Miglani et al. [83], it is interesting to characterize the
features of each mode. As an example mode 1 type atomization is a
very short timescale event O (0.5 ms) in which high aspect ratio
ligaments (1525) grow and fracture along the ligament length to
form multiple daughter droplets. The size of the daughter droplets
are of the order O (0.05 D0). These low momentum daughter droplets containing nanoparticles are transported to the wake primarily by buoyancy. Therefore, the diffusion flame envelope shows
brush like features at the wake tail while the remaining flame front
remains largely quasi-steady. This is validated independently by
measuring the maximum flame stand-off distance (Rf, max/Ro;
Rf, max is assumed to be the maximum stand-off distance of the droplet flame in the forward stagnation region at any given time
instant) which varies within 10% of the time-averaged magnitude


Rt
Rf ;av g 0l Rf tdt=tl for nanofuel droplet flames at dilute concentrations during modes 13 atomization. Mode 1 usually produces
small scale surface undulations of the droplet with virtually no
impact on the centricity. Mode 2 breakup is similar to mode 1 in
terms of short timescales O (0.5 ms). However the ligaments are
thicker (lower aspect ratio) and carry lower kinetic energy leading
to breakup into 23 daughter droplets. Both modes are crucial in
determining the small scale shape deformations in droplets across
all concentrations of nanoparticles. Mode 3 corresponds to medium
intensity atomization and similar to modes 1 and 2.
Burning droplets exhibit mode hops from low intensity (modes
13) to high intensity regimes (modes 4 and 5) as a function of
increased PLR from dilute (0.5% and 1% by wt.) to dense (5% and
7.5% by weight). This observation was noted by Miglani et al.
[83] based on a statistical analysis of the experimental runs. The
shift or transition is primarily due to two mechanisms: (1) the
increased particle loading (high concentration) leads to large number of heterogeneous nucleation sites resulting in higher bubble
density (i.e. number of bubbles per unit volume N/V). Consequently
bubble coalescence and growth probabilities increase. The second
mechanism is related to agglomeration. Higher particle loading
aids in quick formation of a semi-solid porous crust. This crust acts
as a shield between the outer natural convection flow field and
impedes the liquid phase diffusion to the droplet periphery. This
suppresses the droplet gasification rate thereby increasing the
temperature of the liquid core and the solid crust through sensible
heating. The enhanced temperature is conducive to rapid bubble
growth coupled with catastrophic breakup (modes 4 and 5). Both
these mechanisms lead to significant increase in droplet void fraction (/) and internal pressure and can be represented by secondary
deformation index (described in previous sections). Fig. 17 shows a
comparison of all the modes based on their deformation inducing
potential i.e. their ability to corrugate the droplet surface measured
in terms of relative circumference increase (C  ). This shows that
while mode 1 ejection barely induces surface corrugations
(C  < 0.1) the mode 5 catastrophic fragmentation offers significant
enhancement (120%). It was noted that low intensity modes (1, 2
and 3) dominate majority of the droplet lifecycle (0.8tl) for dilute
concentrations. On the other hand, high intensity modes (4 and 5)
control the combustion dynamics at dense PLR (0.5tl). However,
the frequency of these ejections and their ability to distort the droplet may be amplified or suppressed depending on NP agglomeration and subsequent shell formation (i.e. the shell strength).
Next, as shown in Fig. 18 an early transition to modes 4 and 5 at
high concentrations implies ejections of larger sized daughter
droplets. It is obvious that such fragmented droplets transport

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

495

Fig. 15. High resolution, high speed images of dominant secondary atomization modes for nanofuel droplets with dilute concentrations (0.5% and 1 wt.%) of dispersed
nanoparticles: (a) High kinetic energy needle-type ligament with multiple point fracture. Images are at time instants t = 0.25, 0.33 and 0.45 ms, (b) Pin ejection (0.16 and
0.25 ms), (c) Needle ejection with both tip-base break-up (0.75, 0.83, 0.91 and 1.08 ms) and (d) low momentum needle ligament with only tip break-up (0.66, 0.75, 0.83, 0.91
and 1.4 ms). Scale bar equals 1 mm. High speed images of the disrupted flame envelope corresponding to these events are shown on the right. Scale bar equals 2 mm [83].
Reprinted from ASME Journal of Heat Transfer, 137(10), Ankur Miglani, Saptarshi Basu, Effect of particle concentration on shape deformation and secondary atomization
characteristics of burning nanotitania dispersion droplet, 102001, Copyright (2015), with permission from ASME.

relatively larger fractions of suspended NPs to the flame zone compared to dilute counterpart. The effect is twofold. Firstly the number density of nanoparticles in the flame front is consistently
higher as evidenced by the high speed images of the flame front
at different stages of burning. Secondly, the ejections of large parcels of liquid act as projectiles which perforate and distort the
flame front at the global scale (high speed images of Fig. 16). These
physical mechanisms were compiled by Miglani et al. [83].
Miglani et al. [83] presented unique secondary atomization
regimes as shown in Fig. 19. It should be noted that the boundaries
among secondary breakup-regimes are fuzzy and represent gradual transitions. The temporal variation of alocal, max, however, is well
correlated with the transitions in the relative dominance of these
modes at all concentrations.
Effect of initial PLR on burning characteristics was also studied
by Gan et al. [90]. They reported that the mechanism of bringing
out NPs from the droplet domain (a pre-condition for them to
burn) also varies with the base fuel. For instance, in ethanol based
droplet, disruptive behavior features continuous disruptions of
mild intensity while for n-decane based droplet it is marked by
multiple-time disruptions of high intensity (Fig. 19). For latter case,
high intensity break-up is due to a large boiling point differential
between n-decane and surfactant while in former case (with no
added surfactant), the boiling point differential between ethanol
water is nominally small leading to low intensity disruptions
(water condenses continuously on droplet surface during combustion). At dilute PLRs combustion behavior features simultaneous

burning of both the base fuel and NP i.e. a single stage integrated
burning. Contrarily, at dense PLRs combustion proceeds in twostages: (1) some part of NPs undergo integrated burning and (2)
most of the NPs burn as a large agglomerate at later stages when
base fuel has vaporized completely.
In another study, Javed et al. [91] investigated the evaporation
behavior at high temperatures (400800 C), however with a different base fuel (i.e. kerosene) and at dilute concentrations. First,
they demonstrated that micro-explosions are observed only at relatively high temperatures (700800 C) in the kerosene-based
nanofluid fuel droplets while they are absent in pure and stabilized
kerosene droplets. This indicates that the microexplosion phenomenon was solely driven by the presence of Al NPs which provided multiple heterogeneous nucleation sites and promoted
internal pressure build-up through bubble formation. Secondly,
the evaporation rate of nanofuel droplets is significantly affected
by the ambient temperature. Particularly, at relatively low temperatures (400600 C), the augmentation in gasification rate with Al
NPs addition is marginal. However, at relatively high temperatures
(700800 C), the evaporation rate it is substantially enhanced due
to the melting of NPs at droplet surface. Thirdly, the evaporationrate enhancement increases with the NP concentration up to
0.5%, after which it decreases and becomes lower than that of
0.1% NP suspension. The maximum increase was obtained at
800 C at 0.5% where the evaporation rate increased by 56.7%.
In a separate work, Javed et al. [81] studied the microexplosion
behavior of n-Al/kerosene droplets with dense PLRs (2.5, 5 and

496

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Fig. 16. High resolution, high speed images of dominant secondary atomization modes in a burning nanofluid fuel droplet with dense initial PLRs (5% and 7.5 wt.%). (a), (b)
Variants of Mode 4 type ejection: very low momentum, thick ligament break-up, (c) localized catastrophic droplet fragmentation. Scale bars equal 1 mm. High speed images
of flame disruptions corresponding to these events are shown on right. Scale bars equal 2 mm. Typical flame envelope distortion sequence following Mode 5 type secondary
break-up [83]. Reprinted from ASME Journal of Heat Transfer, 137(10), Ankur Miglani, Saptarshi Basu, Effect of particle concentration on shape deformation and secondary
atomization characteristics of burning nanotitania dispersion droplet, 102001, Copyright (2015), with permission from ASME.

7 wt.%) and at elevated temperatures (400800 C). They reported


that NP addition initiates microexplosion at all temperatures
which is previously absent in pure and stabilized kerosene droplets
and hence suspended NPs possibly promote nucleation. The time
delay to onset of microexplosion decreases and its intensity
increases with both increase in PLR (at a particular temperature)
and with increase in temperature. Maximum enhancement in
evaporation rate (48.7%) occurs at 2.5 wt.% at 800 C. Thus
increase in gasification rate is observed till 2.5 wt.% beyond which
it decays. However, no enhancement in vaporization rate occurs at

lower temperatures (400500 C) due to delayed onset and lower


intensity of microexplosions.

5. Control of instabilities in burning droplets: role of external


stimuli (preferential acoustic excitation)
In previous sections, we reviewed the instabilities and atomization pathways in both bi-component as well as nanoparticle laden
fuel droplets. Most of the instabilities were self-excited i.e.

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

497

Fig. 17. Temporal variation of enhancement in circumference during different atomization modes. Different stages during Mode 5 type ejection are shown marked: (a) preejection period, (b) pre-ejection droplet swelling, (c) bubble collapse and ejection, (d) post-ejection droplet shrinkage, (e) onset of geometric shape recovery: post-ejection
first axis-switching and (f) full recovery to pre-ejection state.

Fig. 18. Regime plot showing the temporal distribution of dominant secondary break-up pathways at both dilute and dense concentrations [83]. Reprinted from ASME
Journal of Heat Transfer, 137(10), Ankur Miglani, Saptarshi Basu, Effect of particle concentration on shape deformation and secondary atomization characteristics of burning
nanotitania dispersion droplet, 102001, Copyright (2015), with permission from ASME.

Fig. 19. Different burning modes in n-decane/nano-Fe droplet with surfactants. Reproduced with permission from [90].

498

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

Fig. 20. Schematic of droplet deformation pathway at 100 Hz pulsing mode; showing timescales of key inter-connected mechanisms [64]. Reprinted from Combustion and
Flame, 161 (12), Ankur Miglani, Saptarshi Basu, Ranganathan Kumar, Suppression of instabilities in burning droplets using preferential acoustic perturbations, 31813190,
Copyright (2014), with permission from Elsevier.

threshold beyond which the flame completely blows-off/ extinguishes at the most flame-responsive forcing frequency f MR . The
most responsive mode (f MR ) was identified based on its effect on
the flame structure: (1) at any given forcing amplitude the flame
exhibits highest levels of Rf and Af fluctuations and (2) the blowoff amplitude is lowest at f MR .

phase coupling (in or out of phase) of heat release and pressure


oscillations. In the range of test frequencies (80200 Hz) the acoustic waves had no direct effect on the droplet topology but in 80
120 Hz band the external acoustics locks onto the natural flickering
modes of droplet flame (1316 Hz) with flame exhibiting large
amplitude fluctuations in Rf and Af. Specifically, maximum flame
response with Rf and Af oscillations at 14 Hz (sf = 72 ms) was
observed at 100 Hz excitation.
This flame response timescale (72 ms) is similar to the measured timescales of DL-instability (60 ms) and high intensity bubble ejections O (100 ms) [6364]. However, it is an order of
magnitude higher than that of low intensity expulsions O (2 ms).
Closeness of these timescales is suggestive of a potential interplay
between the droplet flame and internal boiling activity. Based on
these timescale estimates, Miglani el. al [64] devised a unique
pathway (as shown in Figs. 20 and 21) involving indirect bubble
acoustic interaction where the droplet flame controls droplet
deformation. Their findings illustrated that even without the usage
of ultrasound the susceptibility of droplet flame to acoustics can be
exploited to control the ebullition dynamics within the droplet
through indirect pathways. Next section describes the exact mechanism through which this indirect coupling occurs.

5.1. Effect of longitudinal acoustic pulsing on the droplet deformation


behavior

5.2. Mechanism of flame response under external harmonic forcing:


The effect of longitudinal traveling pressure wave

An in-depth understanding of the degree of coupling among


various sub-mechanisms like droplet surface regression, flame
response, major-minor bubble ejections and DL instability requires
a comparative estimate of their relevant timescales. Fig. 20 shows
the timescales of key inter-connected processes that govern droplet deformation behavior. For instance, the characteristic diffusive
timescale for a quasi-steady droplet flame is sdiff = Rf2/Di,j and is
estimated to be O (100 ms) (where Rf, the flame stand-off distance
and Di,j is gas-phase mass diffusivity) while the acoustic timescale
(sf = 1/fP) varies from 512.5 ms in the range of test frequencies.
The shorter acoustic timescale (sf < sdiff) along with forcing at near
extinction critical amplitude dictates that the flame response
parameters Rf and flame area (Af) should exhibit a significant phase
lag in relation to imposed pressure fluctuations [92,64]. However,
amplification/attenuation of fluctuations in Rf and Af depends on

Miglani et al. [64] undertook a comprehensive experimental


study to isolate the unique regimes encountered in the droplet
deformation cycle and ebullition mechanisms at various responsive modes (80 Hz 6 fP 6 120 Hz) as compared to the unresponsive
modes (un-excited and 120 Hz < fP 6 200 Hz) as shown in Fig. 21.
Here, f MR 100 Hz excitation was considered as a representative case for analyzing the dynamic acoustic-flame-bubble interaction in the responsive frequency band (80 Hz 6 fP 6 120 Hz).
Beyond 120 Hz, however, the variation in flame stand-off distance
Rf is within 6% thus indicating a quasi-steady heat source that
uniformly envelopes the droplet. At f MR 100 Hz excitation, the
stable droplet flame transforms to a oscillating heat source with
large amplitude fluctuations in Rf and Af at 14 Hz. At the dominant
14 Hz flickering mode, the flame undergoes a sequential transition
from local extinction and blow-off in the front stagnation region to

naturally occurring. It is prudent to investigate methods by which


such instabilities can be controlled so that the atomization behavior can be tailored according to the need of different applications.
In this section, a novel methodology on the control of instabilities
and secondary break-up using external stimuli is discussed based
on the study by Miglani et al. [64] It was reported that by using
longitudinal acoustic pulsing at selective frequencies and exploiting low pass filter behavior of the droplet flame envelope the secondary atomization frequency and the deformation inducing
potential of bubble ejections can be suppressed. In the range of
acoustic excitation frequencies comprising of both flameresponsive (80 Hz 6 fR 6 120 Hz) and non-responsive modes (unexcited and 120 Hz < fNR 6 200 Hz) the burning droplet was subjected to traveling pressure wave with RMS pressure amplitude
q
p02
f  78.2 Pa. This amplitude corresponds to the maximum

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

499

Fig. 21. Distinct regimes in the droplet deformation cycle with noticeably different characteristics in internal bubble dynamics at flame responsive modes
(80 Hz 6 fP 6 120 Hz) as opposed to non-responsive modes (un-excited and 120 Hz < fP 6 200 Hz) [64]. Reprinted from Combustion and Flame, 161 (12), Ankur Miglani,
Saptarshi Basu, Ranganathan Kumar, Suppression of instabilities in burning droplets using preferential acoustic perturbations, 31813190, Copyright (2014), with
permission from Elsevier.

Fig. 22. 2D flame image sequences showing dynamic flame behavior in a single flame oscillation cycle at fP = 100 Hz. Yellow bars indicate the flame stand-off distance Rf.
Flame oscillation frequency fFR at responsive pulsing mode is shown. Reprinted from Combustion and Flame, 161 (12), Ankur Miglani, Saptarshi Basu, Ranganathan Kumar,
Suppression of instabilities in burning droplets using preferential acoustic perturbations, 31813190, Copyright (2014), with permission from Elsevier. (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version of this article.)

gradual retraction towards leeward side of the droplet [64]. This


constitutes forward transition. This is followed by wake burning
and subsequent reverse propagation till its complete reattachment
in the front stagnation region. This constitutes the reverse transition (as shown in Fig. 22). It was observed that the oscillation time
period can be divided into three different flame states: (1) full
envelope flame (tFE  0.2tc), (2) flame partially surrounding the
droplet (tPE  0.3tc) and (3) wake flame state (tw  0.5tc). The maximum reduction in flame surface area and flame stand-off distance

is 1.85 and 1.7 times respectively compared to the nonresponsive modes.


Droplet flames exhibiting such low-pass filter behavior when
subjected to traveling pressure waves has also been reported previously for petrol and diesel droplets [93]. In this study disturbed
combustion in a narrow frequency band of 90110 Hz with a net
reduction in gasification rate compared to the unpulsed case
(0 Hz). This is very close to the critical range of 80120 Hz reported
by Miglani et al. [64]. In this range a decrease of 20% in the droplet

500

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

lifetime was observed at the non-responsive excitation modes


(0 Hz and 200 Hz) compared to that at critical 100 Hz excitation.
The flame-acoustic coupling can be physically deciphered based
on the wind effect [93] and the multi-state phenomena [9498]. The
wind effect refers to the sweeping-off of the gas molecules accumulated around the droplet surface and subsequently convecting
them away by the sinusoidal flow field set-up at the responsive
modes. However, in the multistate phenomenon, a burning droplet
under forced convection exhibits several flame states and undergoes sequential transitions from one state to another (fullenvelope ? side flame ? wake flame and reverse) with the change
in flow Reynolds number (Re) and temperature. Inaddition, the
transition in droplet flame configuration is distinct during the forward (i.e. an increase in free stream velocity or increasing Re) and
reverse velocity cycles (i.e. a decreasing Re). During forward transition and at low Reynolds number (0 < Re < 45) the droplet flame
is in a full-envelope state and is blown-off to the wake at a critical
Re  45. This is accompanied by 60% reduction in the droplet
vaporization rate [64]. As Re increases beyond a critical value
(Re  45), the wake flame is convected further downstream. The
flame undergoes complete extinction at Re  130 [50]. However,
during backward/reverse transition the wake flame first attains a
partially surrounded or side flame state at Re  25 and then reestablishes to full-envelope geometry (Re  20) by reattachment
in the front stagnation region. This is accompanied by a rapid
increase in the vaporization rate.
The reduction in the gasifcation rate during wake flame state (at
critical Re  45) is inline with the observations of Miglani et al. [64]
where 20% reduction was reported corresponding to a critical
Re  50 as the flame persisted in partial envelope or wake state
for 80% of the total cycle time (tPE + tw = 0.8tc = 57 ms). Based on
these findings, Miglani et al. [64] explained the flame-acoustic coupling phenomenologically as follows: First, the rms velocity fluctuations at the front stagnation point are significantly higher in the
responsive frequency band compared to non-responsive frequencies (being highest at most responsive 100 Hz forcing). At the
unpulsed mode (0 Hz) the natural bouyant velocity (u0 Hz) is measured to be O(1 m/s) corresponding to a maximum Re0 Hz  20 thus
the flame being in a full envelope state. Secondly, during a
single positive half-cycle at 100 Hz forcing (acoustic time scale
sf = 10 ms), the fuel vapor layer resulting from gasification across
the droplet surface is blown-off and deflected to the wake since
the bouyancy induced flow and imposed pulsing are additive (i.e.
u0 Hz + urms, f). This is coupled with an increase in Re to 50. Since
diffusion flame establishes at a location where fuel and oxidizer
(air) are present in stoichiometric concentrations, the depletion of
fuel vapors due to an increase in Reynolds number leads to a mixture fraction reduction thereby causing localized flame extinction.
Subsequently, the flame envelope retracts as fuel vapors are swept
around the droplet shoulders leading to a side flame condition. This
is followed by complete detachment and finally re-establishment in
the wake. This entire transition from full-envelope to wake flame
constitutes the positive half-cycle of 5 ms with Re  50.
Contrarily, during the negative half oscillation cycle, the
reversed acoustic field opposes the natural bouyant velocity (i.e.
u0 Hz  urms, f) which leads to a net flow reversal with a markedly
low Re < 10. Even at such low Re the flame is not restored to a
the full-envelope state. This occurs due to lack of accumulation
of sufficient fuel vapors in the front stagnation region which is a
pre-requisite condition for attaining stoichiometric fuelair ratio
and hence flame re-establishment. Miglani et al. [64] stated that
the Lack of fuel vapor build-up is hindered by three major factors:
(1) the droplet gasification is suppressed as the flame persists in
either wake or side flame state for majority of the lifetime
(80%), (2) the fuel vapor diffusion is quite slow since the charateristic diffusion time-scale Rf2/Di,j  O (100 ms) is an order higher

than the acoustic time scale (s = 10 ms) and (3) although the negative half oscillation cycle helps in gathering fuel vapors through
flow reversal, Re in this state is nominally low (Re < 10) and a following positive half-cycle with a much greater Re (50) nullifies
this effect.
Based on the aforementioned factors that delay fuel vapor
build-up, it was postulated that it takes about 7 oscillation cycles
of the droplet flame (i.e. 7s = 1/14 Hz  72 ms) for enough fuel
vapors to build-up and reach a stoichiometric concentration (i.e.
a critical mixture fraction) for restoring the full-envelope flame
configuration. This explains physically why the droplet flame exhibits a strong frequency response at 14 Hz.
Since flame surface area is one of the indicators of its global heat
release, it is expected that with the flame present in partial envelope or wake burning configuration for considerable time (80% of
tc) a noticeable reduction in Af would result in markedly reduced
heat input (Q) to the droplet at responsive modes compared to
its non-responsive counterparts [64] i.e.

Z
Q 100 Hz
0

t FE

q_ FE tdt

Z
0

t PE

q_ PE tdt

tw

q_ w tdt  Q 0Hz ; Q 200Hz

7
Here, q_ FE , q_ PE and q_ w indicate the oscillations in heat release during the full envelope flame, partial-envelope/side flame and wake
flame regimes respectively. Clearly, the reduction in heat input to
the droplet prolongs its lifespan. Moreover, as the droplet deformation indices a and / varies inversely with droplet gasification rate,
a reduction in droplet surface regression rate leads to altered
regimes in the internal boiling and relatively damped shape oscillations at responsive modes (this shown in Fig. 21).
5.3. Effect of flame response on internal ebullition dynamics
Miglani et al. [64] found that at the responsive 100 Hz pulsing
reduced heat input to the droplet resulted in a reduction in the
average droplet gasification rate by 20%. This caused delayed bubble nucleation (regime I) by initial 15% of tl and reduction in maximum bubble population. Consequently, the droplet deformation
indices were observed to have consistently low magnitudes
throughout the droplet life time i.e. / < 0.2 and alocal, max < 0.15
thus signifying suppressed atomization and a markedly low droplet shape deformation. This was substantiated by the complete
absence of regime IV and broadening of regimes I, II with regime
III extending for majority of the droplet lifetime (0.65tl).
However at non-responsive flame modes, the final regime IV
featuring high intensity bubble ejections with alocal, max > 0.3 is persistent for final 50% of the droplet lifecycle. Thus, in flame nonresponsive modes (0, 200 Hz) burning droplet suffers increasing
degree of deformation as time progresses; transitioning (regime
III) from low magnitude shape oscillations in regime I to intense
oscillations in final stages (regime IV).
Further, it has been reported that at non-responsive forcing
modes the DL instability occurs at millisecond timescale O
(100 ms) at the bubbledroplet interface during the pre-major
ejection transient regime [63]. This is because a continuous heat
input from the full-flame envelope present at these modes ensures
that the critical mass flux is sustained across the liquidvapor
interface for an active instability time period of ti  60 ms. However, at 100 Hz excitation, this DL-instability regime is absent. A
timescale estimate measured from experiments indicates that
flame residence time in in full envelope mode (tFE  15 ms) during
any cycle is  one-fourth of the instability timescale ti; at 100 Hz
[50]. Since triggering of evaporative instabilities requires a
threshold heat flux [89] (consistent with the Landau mechanism),

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

501

Fig. 23. SEM photographs of combustion residue of stabilized 10 wt.% nano-alumina dispersion droplet in unpulsed configuration (a), (b) and at 100 Hz sinusoidal forcing (c),
(d). Magnified images of representative entrapped bubble cavity are shown in (b) and (d) [64]. Reprinted from Combustion and Flame, 161 (12), Ankur Miglani, Saptarshi
Basu, Ranganathan Kumar, Suppression of instabilities in burning droplets using preferential acoustic perturbations, 31813190, Copyright (2014), with permission from
Elsevier.

Miglani et al. [50] conjectured that the flame residence time per
cycle in full envelope state (tFE  15 ms) is not sufficient to supply
the required heat flux. As such DL instability appears to be
suppressed at 100 Hz excitation.
5.4. Effect of external harmonic pulsing on the structural morphology
of combustion residue
To provide further corroborating evidence that the regimes in
bubble statistics (population and number density) are altered at
the responsive modes Miglani et al. [64] conducted a similar set
of experiments with nanofluid fuel droplets. These experiments
offered two new insights. First, they enabled the investigation of
explosion blow holes in the final precipitate that resulted from
bubble ejection events (using ex-situ SEM analysis). Secondly,
nanoparticle laden droplets offer flexibility in testing the effect of
selective acoustic pulsing in controlling the morphology of final
combustion residue. This is pivotal for materials manufacturing
industry where flame synthesis is used for producing nanophase
powders from colloidal droplets.
It is evident from the SEM micrographs of the final precipitate
(Fig. 23; as reported by Miglani et al. [64]) that the size and population density of explosion holes and size of trapped vapor bubble
cavities (i.e. structural porosity measures) are noticeably higher at
the non-responsive frequencies as compared to their responsive
counterparts. Since porosity is expected to be higher in droplets
that favor high frequency and high intensity bubble ejections (i.e.
high a and /), the forcing at responsive 80120 Hz band clearly
leads to suppression of droplet deformation. Thus, this methodology provides a novel control strategy for extending the stability
range of droplet and tuning structural morphology of the final
combustion residue.

6. Concluding remarks
The principal problem affecting several industries (for example,
thermal spray, pharmaceutics, gas turbine combustors, airbreathing propulsion engines) is the lack of understanding of the
thermo-physical processes at the droplet scale particularly for
reacting systems. As a result, the task of creating new synthetic
fuels or generating high quality coatings or custom controlled
drugs become onerous with time consuming trial and error based
experimentation. In this research work, an attempt is made to put
forward mechanisms of bubbledroplet interactions that can
potentially lead to better understanding of combustion phenomenon in burning nanofuel droplets. The current review offers
a diagnostic insight into the physics of burning droplets subjected
to instabilities at different length and time scales following the
works of Miglani et al. [63,64,83,99]. Following are the major
observations:
1. Identification of the mechanisms that lead to secondary
atomization of burning functional droplets (both with and
without nanoscale additives). In addition, the DarrieusLandau
evaporative instability occurring at the bubbledroplet interface has been observed for the first time in context with burning
functional droplets.
2. Identification of distinct secondary atomization pathways in
burning nanofluid fuel droplets and understanding how they
vary temporally with the change in initial NP concentration.
3. Secondary break-up events (or ejections) represent localized
events while their additive effect is observed as the number
of ejections per unit time i.e. atomization frequency. Also, NP
agglomeration leads to crust formation at droplet sub-surface.
Both these global mechanisms occurring at the droplet scale

502

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

have been reviewed and how the extent of feedback coupling


between them affects the atomization process is identified as
a function of NP loading rate.
4. A strategy for controlling the secondary atomization using
external stimuli is proposed. Particularly, by subjecting the droplet flame to external acoustic pulsing in a selective frequency
band (80 6 fp 6 120 i.e. the flame responsive bandwidth) the
droplet deformation can be suppressed. This is due to
the altered regimes in internal boiling that in-turn modulate
the nature of atomization events.
Based on the results reported in the current literature,
nanoparticle laden fuels have shown great promise towards
improved combustion in terms of increasing fuel economy by as
large as 1520% while reducing the pollutant emissions by as high
as 4060%. Compared to other fuels such as hydrogen that are
proposed as future energy carriers metals such as Aluminum,
Boron, Magnesium and Iron are available abundantly in nature,
are easy to store and transport and their end product (metal oxide
formed after combustion) is recyclable. In addition, with recent
advancements in material processing and characterization it is
now possible to manufacture energetic NPs with custom tailored
thermo-physical properties which enable a controlled release of
energy. In this context, energetic nanoparticle laden fuels show
great promise for an efficient well to wheel cycle. Since most of
the modern day combustors/IC engines utilize spray based combustion for power generation, a fundamental understanding on the
combustion behavior of nanofuels at the droplet scale could help
in assessing their efficacy in energy formulations. If successful, this
new-class of fuels could well be the future energy carriers that will
transform the global energy landscape.
References
[1] Jeffrey A. Eastman, S.R. Phillpot, S.U.S. Choi, P. Keblinski, Thermal transport in
nanofluids 1, Annu. Rev. Mater. Res. 34 (2004) 219246.
[2] Sarit Kumar. Das, Stephen U.S. Choi, Hrishikesh E. Patel, Heat transfer in
nanofluidsa review, Heat Transfer Eng. 27 (10) (2006) 319.
[3] Stephen U.S. Choi, Nanofluids: from vision to reality through research, J. Heat
Transfer 131 (3) (2009) 033106.
[4] S.U.S. Choi, Enhancing Thermal Conductivity of Fluids with Nanoparticles,
231, ASME-Publications-Fed, 1995. 99106.
[5] Sarit K. Das, Stephen U. Choi, Wenhua Yu, T. Pradeep, Nanofluids: Science and
Technology, John Wiley & Sons, 2007.
[6] Richard A. Yetter, Grant A. Risha, Steven F. Son, Metal particle combustion and
nanotechnology, Proc. Combust. Inst. 32 (2) (2009) 18191838.
[7] Himanshu Tyagi, Patrick E. Phelan, Ravi Prasher, Robert Peck, Taewoo Lee,
Jose R. Pacheco, Paul Arentzen, Increased hot-plate ignition probability for
nanoparticle-laden diesel fuel, Nano Lett. 8 (5) (2008) 14101416.
[8] Justin L. Sabourin, Daniel M. Dabbs, Richard A. Yetter, Frederick L. Dryer, Ilhan
A. Aksay, Functionalized graphene sheet colloids for enhanced fuel/propellant
combustion, ACS Nano 3 (12) (2009) 39453954.
[9] T. Edwards, Liquid fuels and propellants for aerospace propulsion: 19032003,
J. Propul. Power 19 (2003) 10891107.
[10] L.Q. Maurice, H. Lander, T. Edwards, W.E. Harrison III, Advanced aviation
fuels: a look ahead via a historical perspective, Fuel 80 (2001) 747756.
[11] D.T. Wickham, R.L. Cook, S. De Voss, J.R. Engel, J. Nabity, Soluble nanocatalysts
for high performance fuels, J. Russ. Laser Res. 27 (2006) 552561.
[12] D.T. Wickham, R.L. Cook, J. Engel, M. Jones, J. Nabity, Soluble nanocatalysts for
high performance fuels, in: Presented at the 19th ONR Propulsion Meeting,
Los Angeles, CA, December 20, 2006.
[13] N. Ichinose, Y. Ozaki, S. Kashu, Superfine Particle Technology, Springer-Verlag,
New York, 1992.
[14] J. Song, J. Wang, A.L. Boehman, The role of fuel-borne catalyst in diesel
particulate oxidation behavior, Combust. Flame 146 (2006) 7384.
[15] J.M. Valentine, J.D. Peter-Hoblyn, G.K. Acres, Emissions reduction and
improved fuel economy performance from a bimetallic platinum/cerium
diesel fuel additive at ultra-low dose rates, in: Presented at the CEC/SAE
Spring Fuels & Lubricants Meeting & Exposition, Paris, France, June, 2000.
[16] B. Mellor, M. Ford, Investigation of ignition delay with DMAZ fuel and MON
oxidiser, in: Presented at the 42nd AIAA/ASME/SAE/ASEE Joint Propulsion
Conference and Exhibit, Sacramento, CA, July 912, 2006.
[17] D. Amariei, L. Courtheoux, S. Rossignol, Y. Batonneau, C. Kappenstein, M. Ford,
N. Pillet, Influence of fuel on thermal and catalytic decompositions of ionic
liquid monopropellants, in: Presented at the 41st AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, Tucson, AZ, July 1013, 2005.

[18] M.M. Mench, C.L. Yeh, K.K. Kuo, Propellant burning rate enhancement and
thermal behavior of ultra-fine aluminum powders (ALEX), in: Presented at
the 29th International Annual Conference of ICT, Karlsruhe, Germany, June
30July 3, 1998; pp 115.
[19] G.A. Risha, E. Boyer, R.B. Wehrman, K.K. Kuo, Performance comparison of
HTPB-based solid fuels containing nano-sized energetic powder in a
cylindrical hybrid rocket motor, in: Presented at the 38th AIAA/ASME/SAE/
ASEE Joint Propulsion Conference, Indianapolis, IN, July 710, 2002.
[20] G.A. Risha, E. Boyer, B. Evans, K.K. Kuo, R. Malek, Characterization of nanosized particles for propulsion applications, Mater. Res. Soc. Symp. Proc. 800
(2004). AA6.6.1.
[21] G.A. Risha, J.L. Sabourin, S.F. Son, B. Tappan, V. Yang, R.A. Yetter, Combustion
and conversion efficiency of nanoaluminum-water mixtures, Combust. Sci.
Technol. 180 (2008) 21272142.
[22] G.V. Ivanov, F. Tepper, Challenges in Propellants and Combustion 100 Years
after Nobel, in: K.K. Kuo et al. 1836 R.A. Yetter et al./Proceedings of the
Combustion Institute 32 (2009) 18191838 (Eds.), Begell House, New York,
1997, pp. 636645.
[23] R.W. Armstrong, B. Baschung, D.W. Booth, M. Samirant, Nano Lett. 3 (2003)
253255.
[24] V.N. Simonenko, V.E. Zarko, Comparative study of the combustion behavior of
composite propellants containing ultrafine aluminum, in: Proc. of the 30th
Annual Conference of ICT, 1999, 211.
[25] V. Weiser, N. Eisenreich, Kelzenberg, Influence of the metal particle size on
ignition and combustion of energetic materials, in: Proc. of the 32nd Annual
Conference of ICT, 2001, 341.
[26] P. Lessard, F. Beaupre, P. Brousseau, Burn rate studies of composite
propellants containing ultrafine metals, in: Proc. of the 32nd Annual
Conference of ICT, 2001.
[27] J.R. Luman, B. Wehrman, K.K. Kuo, et al., Proc. Combust. Inst. 31 (2007) 2089.
[28] L. Meda, G. Marra, L. Galfetti, S. Inchingalo, F. Severini, L. DeLuca, Compos. Sci.
Technol. 65 (2005) 769773.
[29] M.J. Chiaverini, K.K. Kuo, A. Peretz, G.C. Harting, AIAA 97-3080, 1997.
[30] M.J. Chiaverini, N. Serin, D.K. Johnson, Y.-C. Lu, K.K. Kuo, G.A. Risha, J. Propul.
Power 16 (2000) 125132.
[31] Matthew Jones, Calvin H. Li, Abdollah Afjeh, G.P. Peterson, Experimental
study of combustion characteristics of nanoscale metal and metal oxide
additives in biofuel (ethanol), Nanoscale Res. Lett. 6 (1) (2011) 112.
[32] K.J. Klabunde, J. Stark, O. Koper, et al., J. Phys. Chem. 100 (1996) 12142
12153.
[33] R.R. Vanfleet, J.M. Mochel, Surf. Sci. 341 (1995) 4050.
[34] F. Delogu, J. Phys. Chem. B 109 (2005) (1941) 2193821941.
[35] S.L. Lai, J.Y. Guo, V. Petrova, G. Ramanath, L.H. Allen, Phys. Rev. Lett. 77 (1996)
99102.
[36] O.J. Gregory, S.B. Lee, R.C. Flagan, Commun. Am. Ceram. Soc. 70 (1987) C-52.
[37] A.P. Ilyin, A.A. Gromov, V.I. Vereshchagin, E.M. Popenko, V.A. Surgin, H. Lehn,
Combust. Explosion Shock Waves 37 (2001) 664.
[38] Y.S. Kown, A.A. Gromov, A.P. Ilyin, E.M. Popenko, G.H. Rim, Combust. Flame
133 (2003) 385.
[39] D.D. Beck, R.W. Siegel, J. Mater. Res. 7 (1992) 2840.
[40] P. Battacharya, S.K. Saha, A. Yadav, P.E. Phelan, R.S. Prasher, J. Appl. Phys. 95
(2004) 64926494.
[41] R.S. Prasher, P. Battacharya, P.E. Phelan, J. Heat Transfer 128 (2006) 588595.
[42] R.S. Prasher, P.E. Phelan, P. Battacharya, Nano Lett. 6 (2006) 15291534.
[43] S. Krishnamurthy, P. Battacharya, P.E. Phelan, Nano Lett. 6 (2006) 419423.
[44] R.S. Prasher, P.E. Phelan, Modeling of radiative and optical behavior of
nanofluids based on multiple and dependent scattering theories, in: ASME
IMECE, Orlando, FL, 2005, Paper No. IMECE2005-80302.
[45] H. Tyagi, P.E. Phelan, R.S. Prasher, Predicted efficiency of a nano-fluid based
direct absorption solar receiver, in: ASME Energy Sustainability Conference,
Long Beach, CA, 2007, Paper No. ES2007-36139.
[46] S.U.S. Choi, Z.G. Zhang, W. Yu, F.E. Lockwood, E.A. Grulke, Appl. Phys. Lett. 79
(2001) 22522254.
[47] Saad. Tanvir, Li. Qiao, Surface tension of nanofluid-type fuels containing
suspended nanomaterials, Nanoscale Res. Lett. 7 (1) (2012) 110.
[48] M.A. Lenin, M.R. Swaminathan, G. Kumaresan, Performance and emission
characteristics of a DI diesel engine with a nanofuel additive, Fuel 109 (2013)
362365.
[49] W.M. Yang et al., Emulsion fuel with novel nano-organic additives for diesel
engine application, Fuel 104 (2013) 726731.
[50] W.M. Yang et al., Impact of emulsion fuel with nano-organic additives on the
performance of diesel engine, Appl. Energy 112 (2013) 12061212.
[51] G.R. Kannan, R. Karvembu, R. Anand, Effect of metal based additive on
performance emission and combustion characteristics of diesel engine
fuelled with biodiesel, Appl. Energy 88 (11) (2011) 36943703.
[52] V. Arul Mozhi Selvan, R.B. Anand, M. Udayakumar, Effect of cerium oxide
nanoparticles and carbon nanotubes as fuel-borne additives in diesterol
blends on the performance, combustion and emission characteristics of a
variable compression ratio engine, Fuel 130 (2014) 160167.
[53] Soner Gumus et al., Aluminum oxide and copper oxide nanodiesel fuel
properties and usage in a compression ignition engine, Fuel 163 (2016) 80.
[54] H. Soukht Saraee et al., Reduction of emissions and fuel consumption in a
compression ignition engine using nanoparticles, Int. J. Environ. Sci. Technol.
(2015) 18.
[55] T. Shaafi et al., Effect of dispersion of various nanoadditives on the
performance and emission characteristics of a CI engine fuelled with diesel,

S. Basu, A. Miglani / International Journal of Heat and Mass Transfer 96 (2016) 482503

[56]

[57]
[58]

[59]

[60]

[61]
[62]

[63]

[64]

[65]

[66]
[67]
[68]
[69]

[70]
[71]
[72]

[73]

[74]
[75]

biodiesel and blends-A review, Renewable Sustainable Energy Rev. 49 (2015)


563573.
T. Shaafi, R. Velraj, Influence of alumina nanoparticles, ethanol and
isopropanol blend as additive with dieselsoybean biodiesel blend fuel:
combustion, engine performance and emissions, Renewable Energy 80 (2015)
655663.
Rakhi N. Mehta, Mousumi Chakraborty, Parimal A. Parikh, Nanofuels:
combustion, engine performance and emissions, Fuel 120 (2014) 9197.
J. Sadhik Basha, R.B. Anand, The influence of nano additive blended biodiesel
fuels on the working characteristics of a diesel engine, J. Braz. Soc. Mech. Sci.
Eng. 35 (3) (2013) 257264.
J. Sadhik Basha, R.B. Anand, An experimental investigation in a diesel engine
using carbon nanotubes blended waterdiesel emulsion fuel, Proc. Inst. Mech.
Eng., Part A: J. Power Energy 225 (3) (2011) 279288.
J. Sadhik Basha, R.B. Anand, Role of nanoadditive blended biodiesel emulsion
fuel on the working characteristics of a diesel engine, J. Renewable
Sustainable Energy 3 (2) (2011) 023106.
Edward L. Dreizin, Metal-based reactive nanomaterials, Prog. Energy
Combust. Sci. 35 (2) (2009) 141167.
C.E. Johnson, S. Fallis, A.P. Chafin, T.J. Groshens, K.T. Higa, I.M.K. Ismail, T.W.
Hawkins, Characterization of nanometer- to micron-sized aluminum
powders: size distribution from thermogravimetric analysis, J. Propul.
Power 23 (2007) 669682.
Ankur Miglani, Saptarshi Basu, Ranganathan Kumar, Insight into
instabilities in burning droplets, Phys. Fluids (1994-present) 26 (3) (2014)
032101. Reprinted with permission from [Ankur Miglani, Saptarshi Basu and
Ranganathan Kumar, Insight into instabilities in burning droplets, Physics of
fluids 26, 032101 (2014)]. Copyright [2014], AIP Publishing LLC.
Ankur Miglani, Saptarshi Basu, Ranganathan Kumar, Suppression of
instabilities in burning droplets using preferential acoustic perturbations,
Combust. Flame 161 (12) (2014) 31813190. Reprinted from Combustion
and Flame, 161 (12), Ankur Miglani, Saptarshi Basu, Ranganathan Kumar,
Suppression of instabilities in burning droplets using preferential acoustic
perturbations, 3181-3190, Copyright (2014), with permission from
Elsevier.
D. Amariei, L. Courtheoux, S. Rossignol, Y. Batonneau, C. Kappenstein, M. Ford,
N. Pillet, Influence of fuel on thermal and catalytic decompositions of ionic
liquid monopropellants, in: Presented at the 41st AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit, Tucson, AZ, July 1013, 2005.
B. Chehroudi, G.L. Vaghjiani, A.D. Ketsdever, Method for Distributed Ignition
of Fuels by Light Sources. U.S. Patent 7,517,215 B1, April 14, 2009.
B.E. Gelfand, Droplet breakup phenomenan in flows with velocity lag, Prog.
Energy Combust. Sci. 22 (3) (1996) 201265.
L.P. Hsiang, G.M. Faeth, Near-limit drop deformation and secondary break-up,
Int. J. Multiphase Flow 18 (5) (1992) 635652.
D.R. Guildenbecher, C. Lopez-Rivera, P.E. Sojka, Droplet deformation and
breakup, in: N. Ashgriz (Ed.), Handbook of Atomization and Sprays-theory
and Applications, Springer, New York, USA, 2011.
C.K. Law, Internal boiling and superheating in vaporizing multicomponent
droplets, AIChE J. 24 (4) (1978) 626632.
C.K. Law, Combustion Physics, Cambridge University Press, New York, USA,
2006.
C.T. Avedisian, I. Glassman, High pressure homogeneous nucleation of
bubbles within superheated binary liquid mixtures, J. Heat Transfer 103 (2)
(1981) 272280.
C.T. Avedisian, I. Glassman, Superheating and boiling of water in
hydrocarbons at high pressures, Int. J. Heat Mass Transfer 24 (4) (1981)
695706.
C.T. Avedisian, R.P. Andres, Bubble nucleation in liquidliquid emulsions, J.
Colloid Interface Sci. 64 (3) (1978) 438.
A. Saha, S. Basu, C. Suryanarayana, R. Kumar, Experimental analysis of
thermo-physical processes in acoustically levitated heated droplets, Int. J.
Heat Mass Transfer 53 (2010) 56635674.

503

[76] Yanan Gan, Li Qiao, Combustion characteristics of fuel droplets with addition
of nano and micron-sized aluminum particles, Combust. Flame 158 (2) (2011)
354368.
[77] Irfan Javed, Seung Wook Baek, Khalid Waheed, Evaporation characteristics of
heptane droplets with the addition of aluminum nanoparticles at elevated
temperatures, Combust. Flame 160 (1) (2013) 170183.
[78] Irfan Javed, Seung Wook Baek, Khalid Waheed, Autoignition and combustion
characteristics of heptane droplets with the addition of aluminium
nanoparticles at elevated temperatures 162 (Combust. Flame) (2015) 191206.
[79] Irfan Javed, Seung Wook Baek, Khalid Waheed, Autoignition and combustion
characteristics of kerosene droplets with dilute concentrations of aluminum
nanoparticles at elevated temperatures, Combust. Flame 162 (3) (2015)
774787.
[80] Michael N. Bello et al., Reaction dynamics of rocket propellant with
magnesium oxide nanoparticles, Energy Fuels 29 (9) (2015) 61116117.
[81] Irfan Javed, Seung Wook Baek, Khalid Waheed, Effects of dense
concentrations of aluminum nanoparticles on the evaporation behavior of
kerosene droplet at elevated temperatures: The phenomenon of
microexplosion, Exp. Therm. Fluid Sci. 56 (2014) 3344.
[82] Philip M. Guerieri et al., Molecular aluminum additive for burn enhancement
of hydrocarbon fuels, J. Phys. Chem. A (2015).
[83] A. Miglani, S. Basu, Effect of particle concentration on shape deformation and
secondary atomization characteristics of burning nanotitania dispersion
droplet, ASME J. Heat Transfer (2015), http://dx.doi.org/10.1115/1.4030394.
Reprinted from ASME Journal of Heat Transfer, 137(10), Ankur Miglani,
Saptarshi Basu, Effect of particle concentration on shape deformation and
secondary atomization characteristics of burning nanotitania dispersion
droplet, 102001, Copyright (2015), with permission from ASME.
[84] D.M. Newitt, N. Dombrowski, F.H. Knelman, Liquid entrainment 1. The
mechanism of drop formation from gas or vapor bubbles, Trans. Inst. Chem.
Eng. 32 (1954) 244.
[85] L.D. Landau, E.M. Lifshitz, Fluid Mechanics, Pergamon Press, London, UK, 1959.
[86] J.E. Shepherd, Dynamics of vapor explosions: rapid evaporation and
instability of butane droplets exploding at the superheat limit (Ph.D.
Thesis), California Institute of Technology, 1981.
[87] D. Frost, B. Sturtevant, Effects of ambient pressure on the instability of a
liquid boiling explosively at the superheat limit, J. Heat Transfer 108 (1986)
418.
[88] J.E. Shepherd, B. Sturtevant, Rapid evaporation at superheat limit, J. Fluid
Mech. 121 (1982) 379.
[89] J.G. Xie, T.E. Ruekgauer, R.L. Armstrong, Evaporative instability in pulsed
laser-heated droplets, Phys. Rev. Lett. 66 (1991) 23.
[90] Yanan Gan, Yi Syuen Lim, Li Qiao, Combustion of nanofluid fuels with the
addition of boron and iron particles at dilute and dense concentrations,
Combust. Flame 159 (4) (2012) 17321740.
[91] Irfan Javed, Seung Wook Baek, Khalid Waheed, Ghafar Ali, Sung Oh Cho,
Evaporation characteristics of kerosene droplets with dilute concentrations of
ligand-protected aluminum nanoparticles at elevated temperatures,
Combust. Flame 160 (12) (2013) 29552963.
[92] H.J. Kim, C.H. Sohn, S.H. Chung, J.S. Kim, KSME Int. J. 15 (4) (2001) 510521.
[93] J. Blaszczyk, Fuel 70 (9) (1991) 10231025.
[94] H.H. Chiu, Prog. Energy Comb. Sci. 26 (2000) 381416.
[95] H.H. Chiu, J.S. Huang, Atomization Sprays 6 (1996) 126.
[96] H.H. Chiu, L.H. Hu, 27th Int. Symposium on Combustion, The Combustion
Institute, Pittsburgh, PA, 1998, pp. 18891896.
[97] G.A. Agoston, H. Wise, W.A. Rosser. Sixth Int. Symposium on Combustion. The
Combustion Institute: Pittsburgh, PA, (1957) 70817.
[98] D.B. Spalding, Fourth Int. Symposium on Combustion, The Combustion
Institute, Pittsburgh, PA, 1954, pp. 847864.
[99] Saad Tanvir, Li Qiao, Effect of addition of energetic nanoparticles on dropletburning rate of liquid fuels, J. Propul. Power 31 (1) (2014) 408415.
[100] Ankur Miglani, Saptarshi Basu, Coupled mechanisms of precipitation and
atomization in burning nanofluid fuel droplets, Sci. Rep. 5 (2015).

Das könnte Ihnen auch gefallen