Next Article in Journal
Drawing a Path towards Circular Construction: An Approach to Engage Stakeholders
Previous Article in Journal
Environmental “Fee-to-Tax” and Heavy Pollution Enterprises to De-Capacity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of CI Engine Performance and Exhaust Air Quality Outfitted with Real-Time Emulsion Fuel Injection System

by
Krishnamoorthy Ramalingam
1,*,
Elumalai Perumal Venkatesan
2,3,
Abdul Aabid
4 and
Muneer Baig
4
1
Department of Mechanical Engineering, CK College of Engineering and Technology, Cuddalore 607003, India
2
Department of Mechanical Engineering, Aditya Engineering College, Surampalem 533437, India
3
Department of Mechanical Engineering, Jawaharlal Nehru Technological University Kakinada, East Godavari District, Kakinada 533003, India
4
Department of Engineering Management, College of Engineering, Prince Sultan University, P.O. Box 66833, Riyadh 11586, Saudi Arabia
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(9), 5313; https://doi.org/10.3390/su14095313
Submission received: 10 February 2022 / Revised: 31 March 2022 / Accepted: 8 April 2022 / Published: 28 April 2022
(This article belongs to the Section Energy Sustainability)

Abstract

:
The main target of the current research work is effectively eliminating fossil fuel dependency and improving the exhaust air quality of conventional Compression Ignition (CI) engines. This research paper demonstrates for the first time that a nanofluid (water without surfactant) stored in separate tanks can be quantified, collected, and immediately emulsified by a high shear mixer before transfer into the combustion chamber of a diesel engine. The experiment was carried out under different load states (25%, 50%, 75% and 100%) with a constant speed of 1500 rpm. Biofuel was extracted from citronella leaves using an energy-intensive process. The 5% water share was used for preparing the biofuel emulsion and nano-biofuel emulsion. A cobalt chromate nanoadditive was used to make the nanofluid. An experimental investigation was performed with prepared test fuels, namely, ultra-low sulphur diesel (ULSD), 100% Citronella (B100), surfactant-free Diesel emulsion (SDE), surfactant-free bioemulsion (SBE), and Surfactant free nano-bioemulsion (SNBE), in a test engine. The properties of the sample test fuels was ensured according to EN and ASTM standards. The observation performance results show that the SNBE blend exhibited lower BTE (by 0.5%) and higher SFC (by 3.4%) than ULSD at peak load. The emission results show that the SNBE blend exhibited lower HC, CO, NOx, and smoke emissions by 23.86%, 31.81%, 2.94%, and 24.63%, respectively, compared to USD at peak load. The CP and HRR results for SNBE were closer to ULSD fuel. Overall, the novel concept of an RTEFI (Real-time emulsion fuel injection) system was proved to be workable and to maintain its benefits of better fuel economy and greener emissions.

1. Introduction

Over the past twenty years, the rate of air pollution has grown rapidly and fossil fuels are becoming depleted due to the growth of industrialization and the drastic increase in the number of transport vehicles. The drain on global fossil energy sources is assessed as being likely to increase over the next ten years, and thought should be given to long-term utilization as well [1,2]. Among the many types of conversion equipment, the basic fuel engine has particular benefits, such as durability, reliability, power output, energy consumption, etc., Nevertheless, diesel fuel engines create high amounts of smoke emissions and nitrogen oxide [3,4]. At present, government emission regulations are very stringent in order to preserve human and environmental health. For these reasons, the scientific community seeks to discover renewable and emission-free alternate fuels for basic engines [5].
To satisfy energy demand, vegetable oil-based alternative fuels have received a great deal of consideration, as it is sustainable and non-toxic. Out of high viscous biofuel, today many researchers are attracted to the topic of low-viscosity biofuel because of its thermo-chemical properties [6,7]. Among low-viscosity fuels, citronella fuel has received the attention of several researchers due to its abundant availability and high yield [8,9]. This vegetable-based biodiesel can be utilized in conventional engines with only minor alterations. In addition, biodiesel has enhanced physical–chemical properties compared to diesel, namely, its Cetane number, oxygen content, and absence of aroma [10,11].
The high oxygen content present in biodiesel prompts total ignition, which may cause lower HC as well as CO formation. On other hand, NOx formation is increased due to elevated peak cylinder temperature, a negative obstacle for biodiesel use in diesel engines [12,13]. Nowadays, water in diesel or water in biodiesel as alternate energy is capable of significantly abating NOx and smoke emissions [14]. The water in emulsified fuel minimizes the peak combustion flame temperature because of the high LHV of water, which results in diminished NOx formation, and the positive effect of secondary atomization promotes better HC, CO, and smoke emission reduction in water–emulsion fuels. In addition, emulsion fuels show enhanced engine performance output response due to their improved air–fuel mixing rate, an effect of micro-explosion and atomization. Thus far, numerous kinds of water added to fuel have been introduced to base engines and their performance and emission output responses have been examined [15,16]. It has been shown that 20% water in diesel fuel emits lower PM and nitrogen oxide emissions by 40% and 50%, respectively, compared to base fuel [17]. In [18], the durability of a base engine fuelled with two different water proportions of emulsion fuel, namely E10 and E20, was evaluated. The test engine was operated for 200hrs energized with both water emulsion fuel, and the results showed a lower carbon store with 60% and half decrease for E10 and E20 fuel correspondingly compared to other emulsions.
Nevertheless, water in oil mixture fuels have one major downside, phase separation, which presents obstacles in the attempt to commercialize this technology. In order to stabilize the water in oil mixture, a surfactant is typically used [19]. Over the past decade, numerous researchers have conducted work to identify the optimal surfactant for a stable water–emulsion fuel. Several researchers have already managed to form stable water in oil emulsion fuels. However, it is costly, and moreover it requires an enormous volume of chemicals and additives, and the process is very tedious [20,21]. Water in oil emulsion fuels can probably be readied and supplied into the engine without of surfactant by persistently blending oil with water employing an in-line blending chamber in front of the delivered emulsion fuel provided to the engine chamber [22]. Ithnin first initiated the investigation into non-surfactant water emulsions with a base fuel in a base engine. The results show better improvement in efficiency and tailpipe emissions compared to diesel fuel, especially for NOx and smoke [23]. The water emulsion fuel had higher HC and CO concentrations when increasing the particular limit of water concentration, because higher ignition lag correlates with water emulsion fuel [24]. In this respect, various adaptations have been made to diminish the ignition delay, such as nanoadditives, metal-based additives, etc., among which nanoadditives have shown the better results [25]. Most commonly, nanoadditives have specific functions to perform, namely, diminishing engine tailpipe emissions, enhancing fuel stability over a long period, lessening flash point and ignition delay, and increasing oxygent content. In addition, nanoadditives have a low melting point, high surface area, better heat and mass transfer, good thermal conductivity, and good oxidation [26,27,28], and are used as performance-enhancing catalysts in diesel engines [29].
Of the numerous possible nanoadditives, cobalt chromate was chosen for the current research work owing to its low cost, good thermal properties, and non-hazardous nature. This research paper aims to evaluate the effect of emulsifier-free nanoemulsion fuels in a diesel engine. A real-time emulsion fuel injection system (RTEFI) was created and connected near the fuel supply arrangement of the engine. In the current work, a mono-cylinder four-stroke CI engine was used under five different load states (25%, 50%, 75%, and 100%) at a constant speed of 1500 rpm. As a comparison, surfactant-free nanoemulsion (SNE), surfactant-free diesel emulsion (SDE), and diesel (D) fuels were tested. The 5% water share was used for both the nanofluid and diesel. The cobalt chromate nanoadditive was used for making the nanofluid. In the test rig analysis, the BTE, CP, HRR and fuel utilization are discussed elaborately. The emission output responses, namely HC, CO, NOx and, smoke, are discussed in detail. A comparison of the performance and emission responses of the assorted nanofuel combinations and diesel is shown in Table 1

2. Materials and Method

2.1. Configuration of RTEFI

Figure 1 exposes the arrangement of the real-time mixing system or RTEFI. Emulsifier-free emulsion fuel is prepared and supplied from this system to the engine, where fuel from the fuel tank is fed into a real-time mixing chamber through the fuel line. The filtered water or nanopowder-mixed water is injected into a real-time mixing system via a water injector from the tank. Initially, the immiscible liquids (water and oil) are homogenously mixed in a greater mixing chamber which responds to a high-power stator and rotator. The prepared emulsion fuel is then supplied to an ultrasonic transducer for introducing vigorous ultrasonic waves before the emulsifier-free emulsion fuel flows through into the diesel engine via the emulsion fuel supply line. The more prominent shear blenders and ultrasonic transducers are set by the regulator. The solenoid valve is used to control the supply of the water and oil fraction, which is shown in Figure 1 as well. The RTEFI system was placed closer to the fuel feed arrangement of the engine in order to eliminate separation of the prepared fuel phase.

2.2. Preparation and Characterization of Citronella Oil

Citronella (Cymbopogon nardus) oil was extracted through an energy-intensive process. The layout out of the energy-intensive setup is shown in Figure 2. Figure 3 and Figure 4 inform the chemical bonds, structure, and functional group present in Cymbopogon nardus oil. A complete study was previously made in our earlier publication (Krishnamoorthy et al. [30]) of oil extraction, FTIR, and GCMS.

2.3. Synthesis of Cobalt Chromite (COCr2O4) Nanoadditive

A redox solution at 250 °C, the commencement of solution combustion, and the end product are the three steps in this process. In a 1000 mL borosilicate container, cobalt (II) nitrate hexahydrate (Co(NO3)2·6H2O, Sigma-Aldrich 99% Purity) and chromium (III) nitrate nonahydrate (Cr(NO3)3·9H2O, Sigma-Aldrich 99% purity) were taken as the respective starting materials according to stoichiometric requirements, and were dissolved in the minimum amount of deionized water with continuous stirring. Then, a reducing agent of glycine fuel was added to the aforementioned solution and the solution was stirred at 200 rpm for 15 to 30 min. After that, the solution was heated with a hotplate and the temperature was elevated to 90 °C to allow for the gradual evaporation of water until a gel was produced. Furthermore, the gel was steadily heated to 200 °C until a flame and gas combustion reaction was accomplished in less than 60 s. The resulting powders were calcinated in air at 500 °C for 5 h and subsequently furnace cooled. After cooling, the resulting materials (CoCr2O4, as Co2+ and Cr3+ ions occupied the A and B sites, respectively) were subjected to grinding via ball milling.

2.4. Characterization of CoCr2O4

2.4.1. XRD Analysis

An X-ray diffraction (XRD) test was conducted for phase identification of the prepared nano-cobalt chromite. The test results in a picture output to conform and verify the crystal structure of the material. Thus, for the prepared nano-cobalt chromite compound the XRD pattern was recorded by X-ray diffractometer, with copper employed as the target material (with Cu Kα as a diffraction source) to explore the sample. It can be seen that the output of the prepared cobalt chromite compound matches the simulated output well. However, an impurity phase screening at near ~33o belongs to a Cr2O3 (@), which means that the single-step synthesis of the Solution Combustion technique does not offer a greater pure phase of CoCr2O4 when major amounts of Cr ion are quickly oxidized to form Cr2O3 precipitate. The crystallite size of the prepared cobalt chromite was found to be ~22 nm as calculated by the Scherrer formula
D p = 0.94 λ β ½   c o s θ
where Dp is the size of crystallite in nm, θ is the degree theta, and β is the radian.

2.4.2. SEM Analysis

SEM microscopy was used for elemental analysis of the prepared cobalt chromite compound. Figure 5a is a pictorial representation of the SEM image of the prepared cobalt chromite compound from an SEM analyzer obtained using an X-ray spectrometer at magnification 20,000× to 167,000×. The experimental analysis was carried out at 8.00 kV and the acquired outcomes were a notable collection of particles with an important poriferous nature.

2.4.3. EDX Analysis

From the point of view of energy content, the content mass is quite natural, which allows the growth of crystallites. Figure 5b exemplifies the Energy Dispersive X-ray (EDX) spectrum of prepared the cobalt chromite compound, which shows that the cobalt, chromium, and oxygen elements are present in the prepared cobalt chromite compound. It is manifest from the collected results that no impurities are present in the prepared CoCr2O4 composition. Figure 5 and Figure 6 displays the recorded and simulated patterns of COCr2O4.

2.5. Test Fuel Preparation

In the current work, test fuel was prepared using two techniques. In the first technique, a nanoemulsion, fuel was prepared using water, an emulsifier, and nanopowder. In the second technique, the novel concept of RTEFI was introduced to prepare and supply the test fuel. The thermal property of test fuel is shown in Table 2.

3. Configuration of the Test Ring

For further investigation, a one-cylinder conventional Kirloskar-made naturally aspirated steady speed direct injection diesel engine with a power output of 5.2 kW was used. The fuel tank was 6.5 L, with the air drawn inside the cylinder through the inlet manifold during the suction stroke. The detailed specifications of the engine are provided in Table 3.
An inline mixing system, the RTEFI system, was attached close to the mechanical fuel injection structure. Pressure sensors were placed on the cylinder head to measure the combustion parameters.
A smoke analyzer and AVL DI gas analyzer were employed to quantify the tailpipe emissions. The emission analyzer specifications are sown in Table 4; Table 5 shows the Test Matrix.

Ambiguity Examination

Commonly, error arises in an instrument due to factors such as instrument condition, environmental conditions, the type of observations being made, and the test procedure. Dissimilarity among instrument makers, calibration techniques, and data collection procedures can lead to experimental uncertainty. Uncertainty analysis is used to avoid variations in the experimental results. The ambiguity among diverse measuring instruments and parameters is shown in Table 6.

4. Result and Discussion

4.1. Performance Output Response

Figure 7 shows the BTE for Diesel (D), Biofuel (B), surfactant-free diesel emulsion (SDE), surfactant-free bioemulsion (SBE), and surfactant-free nano-bioemulsion (SNBE) under diverse load states. From the graph, it can be seen that all the test fuels display increased BTE with increased engine loads. Both the SDE and SBE emulsions and the B fuel have lower BTE correlated with D. Nevertheless, SNBE fuel exhibits slightly higher BTE than SBE, which is due to the positive impact of the cobalt chromite additive [43,44]. There is significant improvement of BTE at high load conditions when fuelled with D, SDE, SBE, and SNBE, 2.22%, 1.26%, 1.64%, and 2.04%, respectively, compared to B100. Additionally, both emulsion fuels are combustible effectively because of micro-explosions and the secondary atomization of fuel. The BTE for D, SDE, SBNE, SBE, and B100 is 31.14%, 30.18%, 30.96%, 30.56%, and 28.92%, respectively, at peak load condition. This outcome might be a reason for the fast evaporation interaction of H2O, which is at first introduced in oil drops as exceptionally fine beads, which positively influences the air/fuel fixing and thus enhances brake thermal efficiency. Another important result is that the emulsion fuels had a higher ignition lag, which allows sufficient duration to physically prepare the test fuel. In addition, this result may be a positive influence on catalytic cracking of the hydrocarbon chain and the O2 buffer property of cobalt chromite nanoparticles.
The BSFC is a significant parameter; it is an indicator that measures engine effectiveness and the amount of fuel utilization inside the engine to develop a certain power output. Figure 8 shows that the BSFC output for D, SDE, SBNE, SBE, and B100 is 292 kJ/kW-h, 326 kJ/kW-h, 302 kJ/kW-h, 312 kJ/kW-h, and 345 kJ/kW-h, respectively, at top load state. The D fuel had better fuel consumption than other test fuels due to the high heat content present in D fuel. There is a significant improvement in BSFC when fuelled with D, SDE, SBE, and SNBE, respectively, 15.36%, 5.51%, 9.57%, and 12.46% compared to B100 under high load conditions. These explanations are consistent with Krishnamoorthy et al. [9], where BFO and BFO emulsion fuel had poor BSFC when correlated with diesel. On the whole, the SNBE fuel exhibits considerable improvement in performance output response as a result of greater dispersion of cobalt chromite nanopowder at the combustion phase.

4.2. Emission Output Response

The nitrogen oxide emission evaluation of the adopted test fuel and contrast with diesel under diverse load states are shown in Figure 9. Generally, a CI engine emits more NOx emission than the SI engine owing to the higher combustion temperature caused by the high oxygen presence inside the CC. The NOx output for D, SDE, SBNE, SB, E and B is 849 ppm, 847 ppm, 824 ppm, 839 ppm, and 852 ppm, respectively, at the top load state. From the graph it can be seen that the B fuel delivers more NOx emissions due to an enriched O2 proportion. It can be seen that B fuel exhibits 50% greater nitrogen oxide formation relative to diesel at peak load state. At 10%, the water concentration in SBE lowers NOx formation by 25% compared to pure B fuel at full load owing to LHV and higher ID of SBE. Furthermore, the addition of the nanoadditive in SBE promotes lower NOx formation by about 10%. The SBNE blend exhibits the maximum NOx reduction due to its positive property of a high surface area to volume ratio [45,46,47,48].
Figure 10 represents HC formation for Diesel (D), Biofuel (B), surfactant-free diesel emulsion (SDE), surfactant-free bioemulsion (SBE), and surfactant-free nano-bioemulsion (SNBE) under varied load states. The HC output for D, SDE, SBNE, SBE, and B is 76 ppm, 67 ppm, 58 ppm, 61 ppm, and 72 ppm, respectively, at the top load state. At low load, it can be seen that the B blend exhibits lower HC formation compared with other test fuels, as the B blend has higher O2 and lower carbon content, which can enhance the proper process of combustion. On the other hand, the SBE blend had greater HC formation than the B blend due to the LHV of water present in the SBE blend, resulting in incomplete combustion. At high load states the SBE blend emits lower HC emission from the CI engine tailpipe than the B blend, owing to the positive impact of micro-explosions and secondary atomization [49,50,51,52]. The cobalt chromite nanoadditive added to the SBE blend considerably reduce HC formation and positively enhance the ignition lag, heat transfer rate, and explosion rate, resulting in fully complete combustion. On the whole, at the maximum load condition the SBNE blend exhibits greater HC emission reduction than all other test fuels due to its positive property of a high surface area to volume ratio.
Generally, the magnitude of CO formation was minimal in the CI engine compared with the SI engine. Usually, the formation of CO emissions depends on the air/fuel ratio in the IC engine. Figure 11 shows CO formation for Diesel (D), Biofuel (B), surfactant-free diesel emulsion (SDE), surfactant-free bioemulsion (SBE) and surfactant-free nano-bioemulsion (SNBE) at varied load ranges. CO decreased with increased engine loads from 0% to 100% for all the test fuels, owing to peak in-cylinder temperature being associated with higher load conditions. At high load state, the SBE blend emits lower CO emission from the CI engine tailpipe than the B blend, owing to the positive impact of micro-explosions and secondary atomization. The cobalt chromite nanoadditive added to the SBE blend considerably reduce CO formation and positively enhance the ignition lag, heat transfer rate, and explosion rate, ad resulting in fully complete combustion [53,54,55,56]. On the whole, at the maximum load condition the SBNE blend exhibits greater CO emissions reduction than all other test fuels due to its positive property of a high surface area to volume ratio.
Figure 12 shows the smoke formation for Diesel (D), Biofuel (B), surfactant-free diesel emulsion (SDE), surfactant-free bioemulsion (SBE), and surfactant-free nano-bioemulsion (SNBE) under varied load range. In order to uphold the constant speed of the engine under the all load states the engine allows the incremental fuel to burn, which influences smoke formation. The graph shows that the biofuels had lower smoke formation than diesel at all load states, which is due to the availability of the oxygen proportion in B fuel. Furthermore, the additional of water content in biofuel diminishes the smoke output compared to diesel thanks to improved atomization [57,58]. At a high load state, the SBE blend emits less smoke from the CI engine tailpipe than the B blend, owing to the positive impact of micro-explosions and secondary atomization. The cobalt chromite nanoadditive added to the SBE blend considerably reduces smoke formation and positively enhances the ignition lag, heat transfer rate, and explosion rate, resulting in fully complete combustion [59,60,61]. On the whole, at the maximum load condition the SBNE blend exhibits greater smoke emission reduction than all other test fuels thanks to its positive property of a high surface area to volume ratio [62,63,64].

4.3. Combustion Output Response

In a diesel engine the cylinder pressure commonly depends on various factors, namely, the equivalence rate, A/F rate, evaporation rate, ignition delay and CD period, and fuel burned separately during the premixed burning stage. Figure 13 shows the variation in Cylinder Pressure (CP) with respect to the crank angle for Diesel (D), Biofuel (B), surfactant-free diesel emulsion (SDE), surfactant-free bioemulsion (SBE), and surfactant-free nano-bioemulsion (SNBE) under varied load ranges. The maximum CP output for D, SDE, SBNE, SBE, and B100 is 65 bar, 67 bar, 61 bar, 64 bar, and 63 bar, respectively, at the top load state. Both SDE and SBE emulsion and B fuel show lower CP correlated to D due to the CV of D fuel. Nevertheless, SNBE fuel exhibits slightly higher CP than SBE due to the positive impact of cobalt chromite additive. Additionally, emulsion fuel combusts effectively thanks to secondary atomization and micro-explosions. This result may be a reason for the fast process of H2O evaporation that is organically introduced in oil drop attacks as exceptionally fine drops, which influences air/fuel fixing and enhances brake thermal efficiency. Another important element of this result is that emulsion fuel has a higher ignition lag, which allows time to physically prepare the test fuel. This result may show a positive influence on catalytic cracking of the hydrocarbon chain and the O2 buffer property of cobalt chromite nanoparticles [62,63,64].
HRR is the measure of heat radiated by most of the burn throughout the combustion stage inside the base engine cylinder. In the present work, the HRR was calculated from the CP with respect to the Crank angle (CA) using following the expression equation
d Q n e t d θ = ( γ γ 1   P   d V d θ ) + ( 1 γ 1   V   d P d θ   )
where:
  • d Q n e t d θ = Net heat release rate in (kJ/m3 CA);
  • V = Instantaneous volume in (m3);
  • P = Instantaneous pressure in (bar);
  • γ = ratio of specific heat.
Figure 14 shows the variation in HRR with respect to the crank angle for Diesel (D), Biofuel (B), surfactant-free diesel emulsion (SDE), surfactant-free bioemulsion (SBE), and surfactant-free nano-bioemulsion (SNBE) under varied load ranges. Both SDE and SBE emulsion and B fuel show lower CP correlated to D due to the CV of D fuel. Nevertheless, SNBE fuel exhibits a slightly higher HRR than SBE due to the positive impact of the cobalt chromite additive.

5. Conclusions

The performance, emissions, and combustion of citronella biodiesel blended with water and a nanoadditive were investigated in a single-cylinder diesel engine with attached RTEFI. The obtained conclusions are listed below.
  • The BTE of the engine was enhanced with SBE fuel by 1.64% and with SNBE fuel by 2.04% compared to unadulterated Citronella fuel at the most extreme load range. In all cases, conventional fuel activity displayed the greatest BTE compared to all other test fuels
  • At the maximum load range, the BSFC of the engine was enhanced with SBE and SNBE fuel by 9.57% and 12.46%, respectively, compared to pure Citronella fuel. and was. The respective BSFC for the D, SDE, SBE, and SNBE fuels was 292 kJ/kW-h, 326 kJ/kW-h, 312 kJ/kW-h, and 302 kJ/kW-h.
  • The SDE, B100, SBE, and SNBE fuels had a significantly reduced HC in the tailpipe, by 13.16%, 5.27%, 19.75%, and 23.68%, respectively, and CO in the tailpipe. by 13.65%), (4.54%), (22.73%), and (31.82%), respectively, compared with conventional diesel fuel at the peak load condition. However, the SDE blend exhibited higher HC and CO emissions than Diesel fuel.
  • At all load ranges, the NOx emission trend was lower for SBE than for diesel and the SDE blend (by 1.18% and 2.95%, respectively) as well as for SNBE (by 0.94% and 2.72%, respectively). However, pure citronella fuel showed greater NOx formation than SBE fuel.
  • A significant improvement in the opacity of smoke was noted for both emulsion bio-fuels relative to B100 fuel, with a respective decrease of 10.77% and 20%, while Diesel fuel had increased smoke formation compared to both the nanoemulsion and emulsion fuels.
  • CP and HRR increased for both nanoemulsion and emulsion SBE and SBNE fuels compared to convention diesel fuel; however, B100 fuel had minimal CP and HRR compared to Diesel fuel.

Author Contributions

Conceptualization, K.R., A.A. and M.B.; Data curation, E.P.V.; Investigation, K.R.; Methodology, K.R. and E.P.V.; Software, E.P.V.; Supervision, E.P.V. and M.B.; Validation, A.A.; Visualization, A.A.; Writing—original draft, K.R.; Writing—review & editing, A.A. and M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the Structures and Materials (S&M) Research Lab of Prince Sultan University and the authors acknowledge the support of Prince Sultan University for paying the article processing charges (APC) of this publication.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to express heartful thanks to S. Saravanan, Principal, CK College of Engineering and Technology, Cuddalore. The authors thank K. Annamalai, Automobile Department, MIT, Anna University, Chromepet and Chennai 44 for allowing us to utilize the Engine Research Lab for this work.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CICompression Ignition
DIDirect Injection
COCarbon monoxide
NOxNitrogen oxides
CO2Carbon dioxide
HCHydrocarbon
BSECBrake Specific Energy Consumption
BTEBrake Thermal Efficiency
IDIgnition Delay
CNCetane Number
ASTMAmerican Society for Testing and Materials
BPBrake power
TFCTotal fuel consumption
HRRHeat Release Rate
UBHCUnburned Hydrocarbon
H2OWater
HSUHartridge Smoke Units
O2Oxygen
CACrank Angle
LGOLemongrass Oil
WCOWaste Cooking Oil
GC-MS Gas Chromatography-Mass Spectrometry
FT-IR Fourier Transform Infrared spectroscopy
USIAUnited State Energy Information Administration
RTEFIReal Time Emulsion Fuel Injection

References

  1. Zou, C.; Zhao, Q.; Zhang, G.; Xiong, B. Energy revolution: From a fossil energy era to a new energy era. Nat. Gas Ind. B 2016, 3, 1–11. [Google Scholar] [CrossRef] [Green Version]
  2. Ramalingam, K.; Balasubramanian, D.; Chellakumar, P.J.T.J.S.; Padmanaban, J.; Murugesan, P.; Xuan, T. An assessment on production and engine characterization of a novel environment-friendly fuel. Fuel 2020, 279, 118558. [Google Scholar] [CrossRef]
  3. Venu, H.; Madhavan, V. Influence of diethyl ether (DEE) addition in ethanol-biodiesel-diesel (EBD) and methanol-biodiesel-diesel (MBD) blends in a diesel engine. Fuel 2017, 189, 377–390. [Google Scholar] [CrossRef]
  4. Agarwal, A.K. Biofuels (alcohols and biodiesel) applications as fuels for internal combustion engines. Prog. Energy Combust. Sci. 2007, 33, 233–271. [Google Scholar] [CrossRef]
  5. Bezergianni, S.; Dimitriadis, A. Comparison between different types of renewable diesel. Renew. Sustain. Energy Rev. 2013, 21, 110–116. [Google Scholar] [CrossRef]
  6. Ashok, B.; Nanthagopal, K.; Chaturvedi, B.; Sharma, S.; Raj, R.T.K. A comparative assessment on Common Rail Direct Injection (CRDI) engine characteristics using low viscous biofuel blends. Appl. Therm. Eng. 2018, 145, 494–506. [Google Scholar] [CrossRef]
  7. Ramalingam, K.; Kandasamy, A.; Balasubramanian, D.; Palani, M.; Subramanian, T.; Varuvel, E.G.; Viswanathan, K. Forcasting of an ANN model for predicting behaviour of diesel engine energised by a combination of two low viscous biofuels. Environ. Sci. Pollut. Res. 2019, 27, 24702–24722. [Google Scholar] [CrossRef] [PubMed]
  8. Kumar, J.R.; Vikram, C.J.; Naveenchandran, P. Investigation for the performance and emission test using citronella oil in single cylinder diesel engine. Middle East J. Sci. Res. 2012, 12, 1754–1757. [Google Scholar]
  9. Hoekman, S.K.; Broch, A.; Robbins, C.; Ceniceros, E.; Natarajan, M. Review of biodiesel composition, properties, and specifications. Renew. Sustain. Energy Rev. 2012, 16, 143–169. [Google Scholar] [CrossRef]
  10. Atabani, A.; Silitonga, A.; Badruddin, I.A.; Mahlia, T.M.I.; Masjuki, H.; Mekhilef, S. A comprehensive review on biodiesel as an alternative energy resource and its characteristics. Renew. Sustain. Energy Rev. 2012, 16, 2070–2093. [Google Scholar] [CrossRef]
  11. Ramalingam, K.; Kandasamy, A.; Subramani, L.; Balasubramanian, D.; Thadhani, J.P.J. An assessment of combustion, performance characteristics and emission control strategy by adding anti-oxidant additive in emulsified fuel. Atmos. Pollut. Res. 2018, 9, 959–967. [Google Scholar] [CrossRef]
  12. Hoekman, S.K.; Robbins, C. Review of the effects of biodiesel on NOx emissions. Fuel Process. Technol. 2012, 96, 237–249. [Google Scholar] [CrossRef]
  13. Al-Dawody, M.; Bhatti, S. Optimization strategies to reduce the biodiesel NOx effect in diesel engine with experimental verification. Energy Convers. Manag. 2013, 68, 96–104. [Google Scholar] [CrossRef]
  14. Ramalingam, K.; Kandasamy, A.; Chellakumar, P.J.T.J.S. Production of eco-friendly fuel with the help of steam distillation from new plant source and the investigation of its influence of fuel injection strategy in diesel engine. Environ. Sci. Pollut. Res. 2019, 26, 15467–15480. [Google Scholar] [CrossRef] [PubMed]
  15. Elsanusi, O.A.; Roy, M.M.; Sidhu, M.S. Experimental Investigation on a Diesel Engine Fueled by Diesel-Biodiesel Blends and their Emulsions at Various Engine Operating Conditions. Appl. Energy 2017, 203, 582–593. [Google Scholar] [CrossRef]
  16. Vigneswaran, R.; Annamalai, K.; Dhinesh, B.; Krishnamoorthy, R. Experimental investigation of unmodified diesel engine performance, combustion and emission with multipurpose additive along with water-in-diesel emulsion fuel. Energy Convers. Manag. 2018, 172, 370–380. [Google Scholar] [CrossRef]
  17. Leng, L.; Chen, J.; Leng, S.; Li, W.; Huang, H.; Li, H.; Yuan, X.; Li, J.; Zhou, W. Surfactant assisted upgrading fuel properties of waste cooking oil biodiesel. J. Clean. Prod. 2019, 210, 1376–1384. [Google Scholar] [CrossRef]
  18. Lin, C.-Y.; Lin, S.A. Effects of emulsification variables on fuel properties of two-and three-phase biodiesel emulsions. Fuel 2007, 86, 210–217. [Google Scholar] [CrossRef]
  19. Vellaiyan, S.; Amirthagadeswaran, K.S. Zinc oxide incorporated water-in-diesel emulsion fuel: Formulation, particle size measurement, and emission characteristics assessment. Pet. Sci. Technol. 2016, 34, 114–122. [Google Scholar] [CrossRef]
  20. Musculus, M.P.B.; Dec, J.E.; Tree, D.R.; Daly, D.; Langer, D.; Ryan, T.W., III; Matheaus, A.C. Effects of water-fuel emulsions on spray and combustion processes in a heavy-duty DI diesel engine. SAE Trans. 2002, 111, 2736–2756. [Google Scholar] [CrossRef]
  21. Ithnin, A.M.; Yahya, W.; Ahmad, M.A.; AtiqahRamlan, N.; Kadir, H.A.; Sidik, N.A.C.; Koga, T. Emulsifier-free Water-in-Diesel emulsion fuel: Its stability behaviour, engine performance and exhaust emission. Fuel 2018, 215, 454–462. [Google Scholar] [CrossRef]
  22. Vellaiyan, S. Enhancement in combustion, performance, and emission characteristics of a biodiesel-fueled diesel engine by using water emulsion and nanoadditive. Renew. Energy 2020, 145, 2108–2120. [Google Scholar] [CrossRef]
  23. Vellaiyan, S.; Subbiah, A.; Chockalingam, P. Combustion, performance, and emission analysis of diesel engine fueled with water-biodiesel emulsion fuel and nanoadditive. Environ. Sci. Pollut. Res. 2018, 25, 33478–33489. [Google Scholar] [CrossRef] [PubMed]
  24. Jeryrajkumar, L.; Anbarasu, G.; Elangovan, T. Effects on nanoadditives on performance and emission characteristics of calophylliminophyllum biodiesel. Int. J. Chem. Tech. Res. 2016, 9, 210–219. [Google Scholar]
  25. Soudagar, M.E.M.; Nik-Ghazali, N.-N.; Kalam, A.; Badruddin, I.A.; Banapurmath, N.R.; Akram, N. The effect of nano-additives in diesel-biodiesel fuel blends: A comprehensive review on stability, engine performance and emission characteristics. Energy Convers. Manag. 2018, 178, 146–177. [Google Scholar] [CrossRef]
  26. Chandrasekaran, V.; Arthanarisamy, M.; Nachiappan, P.; Dhanakotti, S.; Moorthy, B. The role of nanoadditives for biodiesel and diesel blended transportation fuels. Transp. Res. Part D Transp. Environ. 2016, 46, 145–156. [Google Scholar] [CrossRef]
  27. Saravankumar, P.T.; Suresh, V.; Vijayan, V.; Antony, G. Ecological effect of corn oil biofuel with SiO2 nano-additives. Energy Sources Part A: Recover. Util. Environ. Eff. 2019, 41, 2845–2852. [Google Scholar] [CrossRef]
  28. Krishnamoorthy, R.; Asaithambi, K.; Balasubramanian, D.; Murugesan, P.; Rajarajan, A. Effect of Cobalt Chromite on the Investigation of Traditional CI Engine Powered with Raw Citronella Fuel for the Future Sustainable Renewable Source; No. 2020-28-0445; SAE Technical Paper; SAE: Warrendale, PA, USA, 2020. [Google Scholar]
  29. Gad, M.; Jayaraj, S. A comparative study on the effect of nano-additives on the performance and emissions of a diesel engine run on Jatropha biodiesel. Fuel 2020, 267, 117168. [Google Scholar] [CrossRef]
  30. Hoseini, S.; Najafi, G.; Ghobadian, B.; Ebadi, M.; Mamat, R.; Yusaf, T. Performance and emission characteristics of a CI engine using graphene oxide (GO) nano-particles additives in biodiesel-diesel blends. Renew. Energy 2019, 145, 458–465. [Google Scholar] [CrossRef]
  31. Kumar, S.; Dinesha, P.; Ajay, C.M.; Kabbur, P. Combined effect of oxygenated liquid and metal oxide nanoparticle fuel additives on the combustion characteristics of a biodiesel engine operated with higher blend percentages. Energy 2020, 197, 117194. [Google Scholar] [CrossRef]
  32. Devarajan, Y.; Munuswamy, D.B.; Mahalingam, A. Investigation on behavior of diesel engine performance, emission, and combustion characteristics using nano-additive in neat biodiesel. Heat Mass Transf und Stoffuebertragung. Heat Mass Transf. 2019, 55, 1641–1650. [Google Scholar] [CrossRef]
  33. Radhakrishnan, S.; Munuswamy, D.B.; Devarajan, Y.; Mahalingam, A. Performance, emission and combustion study on neat biodiesel and water blends fuelled research diesel engine. Heat Mass Transf und Stoffuebertragung. Heat Mass Transf. 2019, 55, 1229–1237. [Google Scholar] [CrossRef]
  34. Dhinesh, B.; Annamalai, M. A study on performance, combustion and emission behaviour of diesel engine powered by novel nano nerium oleander biofuel. J. Clean. Prod. 2018, 196, 74–83. [Google Scholar] [CrossRef]
  35. Yuvarajan, D.; Babu, M.D.; BeemKumar, N.; Kishore, P.A. Experimental investigation on the influence of titanium dioxide nanofluid on emission pattern of biodiesel in a diesel engine. Atmos. Pollut. Res. 2018, 9, 47–52. [Google Scholar] [CrossRef]
  36. Venkata, R.M.; Yuvarajan, D. Emission analysis on the influence of magnetite nanofluid on methyl ester in diesel engine. Atmos Pollut. Res. 2016, 7, 477–481. [Google Scholar] [CrossRef]
  37. Dhahad, H.A.; Chaichan, M.T. The impact of adding nano-Al2O3 and nano-ZnO to Iraqi diesel fuel in terms of compression ignition engines’ performance and emitted pollutants. Therm. Sci. Eng. Prog. 2020, 18, 100535. [Google Scholar] [CrossRef]
  38. Basha, J.S. Impact of Carbon Nanotubes and Di-Ethyl Ether as additives with biodiesel emulsion fuels in a diesel engine—An experimental investigation. J. Energy Inst. 2016, 91, 289–303. [Google Scholar] [CrossRef]
  39. Sathiyamoorthi, R.; Sankaranarayanan, G. Effect of antioxidant additives on the performance and emission characteristics of a DICI engine using neat lemongrass oil–diesel blend. Fuel 2016, 174, 89–96. [Google Scholar] [CrossRef]
  40. Bora, P.; Boro, J.; Konwar, L.J.; Deka, D. Formulation of microemulsion based hybrid biofuel from waste cooking oil—A comparative study with biodiesel. J. Energy Inst. 2016, 89, 560–568. [Google Scholar] [CrossRef]
  41. Vinukumar, K.; Azhagurajan, A.; Vettivel, S.C.; Vedaraman, N. Rice husk as nanoadditive in diesel–biodiesel fuel blends used in diesel engine. J. Therm. Anal. 2018, 131, 1333–1343. [Google Scholar] [CrossRef]
  42. Kareemullah, M.; Afzal, A.; Rehman, K.F.; Shahapurkar, K.; Khan, H.; Akram, N. Performance and emission analysis of compression ignition engine using biodiesels from Acid oil, Mahua oil, and Castor oil. Heat Transf. 2019, 49, 858–871. [Google Scholar] [CrossRef]
  43. Afzal, A.; Mujeebu, M.A. Thermo-Mechanical and Structural Performances of Automobile Disc Brakes: A Review of Numerical and Experimental Studies. Arch. Comput. Methods Eng. 2019, 26, 1489–1513. [Google Scholar] [CrossRef]
  44. Samuel, O.D.; Okwu, M.; Oyejide, O.J.; Taghinezhad, E.; Afzal, A.; Kaveh, M. Optimizing biodiesel production from abundant waste oils through empirical method and grey wolf optimizer. Fuel 2020, 281, 118701. [Google Scholar] [CrossRef]
  45. Kumar, S.S.; Rajan, K.; Mohanavel, V.; Ravichandran, M.; Rajendran, P.; Rashedi, A.; Sharma, A.; Khan, S.A.; Afzal, A. Combustion, Performance, and Emission Behaviors of Biodiesel Fueled Diesel Engine with the Impact of Alumina Nanoparticle as an Additive. Sustainability 2021, 13, 12103. [Google Scholar] [CrossRef]
  46. Aneeque, M.; Alshahrani, S.; Kareemullah, M.; Afzal, A.; Saleel, C.; Soudagar, M.; Hossain, N.; Subbiah, R.; Ahmed, M. The Combined Effect of Alcohols and Calophylluminophyllum Biodiesel Using Response Surface Methodology Optimization. Sustainability 2021, 13, 7345. [Google Scholar] [CrossRef]
  47. Ağbulut, Ü.; Sarıdemir, S.; Rajak, U.; Polat, F.; Afzal, A.; Verma, T.N. Effects of high-dosage copper oxide nanoparticles addition in diesel fuel on engine characteristics. Energy 2021, 229, 120611. [Google Scholar] [CrossRef]
  48. Razzaq, L.; Mujtaba, M.; Soudagar, M.E.M.; Ahmed, W.; Fayaz, H.; Bashir, S.; Fattah, I.R.; Ong, H.C.; Shahapurkar, K.; Afzal, A.; et al. Engine performance and emission characteristics of palm biodiesel blends with graphene oxide nanoplatelets and dimethyl carbonate additives. J. Environ. Manag. 2021, 282, 111917. [Google Scholar] [CrossRef]
  49. Samuel, O.D.; Samuel, O.D.; Waheed, M.A.; Waheed, M.A.; Taheri-Garavand, A.; Taheri-Garavand, A.; Verma, T.N.; Verma, T.N.; Dairo, O.U.; Dairo, O.U.; et al. Prandtl number of optimum biodiesel from food industrial waste oil and diesel fuel blend for diesel engine. Fuel 2020, 285, 119049. [Google Scholar] [CrossRef]
  50. Ağbulut, Ü.; Elibol, E.; Demirci, T.; Sarıdemir, S.; Gürel, A.E.; Rajak, U.; Afzal, A.; Verma, T.N. Synthesis of graphene oxide nanoparticles and the influences of their usage as fuel additives on CI engine behaviors. Energy 2021, 244, 122603. [Google Scholar] [CrossRef]
  51. Yaliwal, V.; Banapurmath, N.; Soudagar, M.E.M.; Afzal, A.; Ahmadi, P. Effect of manifold and port injection of hydrogen and exhaust gas recirculation (EGR) in dairy scum biodiesel—low energy content gas-fueled CI engine operated on dual fuel mode. Int. J. Hydrogen Energy 2021, 47, 6873–6897. [Google Scholar] [CrossRef]
  52. Khandal, S.V.; Ağbulut, Ü.; Afzal, A.; Sharifpur, M.; Razak, K.A.; Khalilpoor, N. Influences of hydrogen addition from different dual-fuel modes on engine behaviors. Energy Sci. Eng. 2022, 10, 881–891. [Google Scholar] [CrossRef]
  53. Wategave, S.; Banapurmath, N.; Sawant, M.; Soudagar, M.E.M.; Mujtaba, M.; Afzal, A.; Basha, J.; Alazwari, M.A.; Safaei, M.R.; Elfasakhany, A.; et al. Clean combustion and emissions strategy using reactivity controlled compression ignition (RCCI) mode engine powered with CNG-Karanja biodiesel. J. Taiwan Inst. Chem. Eng. 2021, 124, 116–131. [Google Scholar] [CrossRef]
  54. Arunkumar, M.; Mohanavel, V.; Afzal, A.; Sathish, T.; Ravichandran, M.; Khan, S.; Abdullah, N.; Bin Azami, M.; Asif, M. A Study on Performance and Emission Characteristics of Diesel Engine Using Ricinus Communis (Castor Oil) Ethyl Esters. Energies 2021, 14, 4320. [Google Scholar] [CrossRef]
  55. Verma, T.N.; Rajak, U.; Dasore, A.; Afzal, A.; Manokar, A.M.; Aabid, A.; Baig, M. Experimental and empirical investigation of a CI engine fuelled with blends of diesel and roselle biodiesel. Sci. Rep. 2021, 11, 23. [Google Scholar] [CrossRef] [PubMed]
  56. Elumalai, P.V.; Parthasarathy, M.; Hariharan, V.; Jayakar, J.; Iqbal, S.M. Evaluation of water emulsion in biodiesel for engine performance and emission characteristics. J. Therm. Anal. 2022, 147, 4285–4301. [Google Scholar] [CrossRef]
  57. Sivakandhan, C.; Elumalai, P.V.; Murugan, M.; Saravanan, A.; Ranjit, P.S.; Varaprasad, B.; Murugan, M. Effects of on MnO2 nanoparticles behavior of a sardine oil methyl ester operated in thermal barrier coated engine. J. Therm. Anal. 2022, 13. [Google Scholar] [CrossRef]
  58. Elumalai, P.V.; Senthur, N.S.; Parthasarathy, M.; Das, S.K.; Samuel, O.D.; Reddy, M.S.; Saravana, A.; Anjanidevi, S.; Sitaramamurty, A.S.; Anusha, M.; et al. Hydrogen in Spark Ignition Engines. In Application of Clean Fuels in Combustion Engines. Energy, Environment, and Sustainability; Di Blasio, G., Agarwal, A.K., Belgiorno, G., Shukla, P.C., Eds.; Springer: Singapore, 2022. [Google Scholar] [CrossRef]
  59. Afzal, A.; Samee, A.D.M.; Javad, A.; Shafvan, S.A.P.V.A.; Kabeer, K.M.A. Heat transfer analysis of plain and dimpled tubes with different spacings. Heat Transf. Asian Res. 2017, 47, 556–568. [Google Scholar] [CrossRef]
  60. Soudagar, M.E.M.; Kalam, M.A.; Sajid, M.U.; Afzal, A.; Banapurmath, N.R.; Akram, N.; Mane, S.; Saleel, A.C. Thermal analyses of minichannels and use of mathematical and numerical models. Numer. Heat Transfer. Part A Appl. 2020, 77, 497–537. [Google Scholar] [CrossRef]
  61. Afzal, A.; Samee, A.D.M.; Razak, R.K.A.; Ramis, M.K. Steady and Transient State Analyses on Conjugate Laminar Forced Convection Heat Transfer. Arch. Comput. Methods Eng. 2020, 27, 135–170. [Google Scholar] [CrossRef]
  62. Afzal, A.; Samee, A.M.; Jilte, R.; Islam, T.; Manokar, A.M.; Razak, K.A. Battery thermal management: An optimization study of parallelized conjugate numerical analysis using Cuckoo search and Artificial bee colony algorithm. Int. J. Heat Mass Transf. 2020, 166, 120798. [Google Scholar] [CrossRef]
  63. Elumalai, P.V.; Dhinesh, B.; Jayakar, J.; Nambiraj, M.; Hariharan, V. Effects of antioxidants to reduce the harmful pollutants from diesel engine using preheated palm oil–diesel blend. J. Therm. Anal. Calorim. 2022, 147, 2439–2453. [Google Scholar] [CrossRef]
  64. Elumalai, P.V.; Senthur, N.S.; Parthasarathy, M.; Dash, S.K.; Samuel, O.D.; Reddy, M.S.; Murugan, M.; Das, P.K.; Sitaramamurty, A.S.S.M.; Anjanidevi, S.; et al. Effectiveness of Hydrogen and Nanoparticles Addition in Eucalyptus Biofuel for Improving the Performance and Reduction of Emission in CI Engine. In Greener and Scalable E-fuels for Decarbonization of Transport. Energy, Environment, and Sustainability; Agarwal, A.K., Valera, H., Eds.; Springer: Singapore, 2022. [Google Scholar] [CrossRef]
Figure 1. Configuration of RTEFIS: (a) pictographic representation of RTEFIS setup and (b) schematic representation.
Figure 1. Configuration of RTEFIS: (a) pictographic representation of RTEFIS setup and (b) schematic representation.
Sustainability 14 05313 g001
Figure 2. (a) Schematic of energy-intensive process and (b) photograph of energy-intensive process.
Figure 2. (a) Schematic of energy-intensive process and (b) photograph of energy-intensive process.
Sustainability 14 05313 g002
Figure 3. FT-IR for Citronella oil.
Figure 3. FT-IR for Citronella oil.
Sustainability 14 05313 g003
Figure 4. GC-MS for Citronella oil.
Figure 4. GC-MS for Citronella oil.
Sustainability 14 05313 g004
Figure 5. (a) SEM Image and (b) EDX Spectra of nano-cobalt chromite.
Figure 5. (a) SEM Image and (b) EDX Spectra of nano-cobalt chromite.
Sustainability 14 05313 g005
Figure 6. XRD for nano-cobalt chromite.
Figure 6. XRD for nano-cobalt chromite.
Sustainability 14 05313 g006
Figure 7. BTE vs. BP.
Figure 7. BTE vs. BP.
Sustainability 14 05313 g007
Figure 8. BSFC vs. BP.
Figure 8. BSFC vs. BP.
Sustainability 14 05313 g008
Figure 9. Nitrogen oxide emissions vs. brake power.
Figure 9. Nitrogen oxide emissions vs. brake power.
Sustainability 14 05313 g009
Figure 10. Hydrocarbon emissions vs. brake power.
Figure 10. Hydrocarbon emissions vs. brake power.
Sustainability 14 05313 g010
Figure 11. Carbon monoxide emissions vs. brake power.
Figure 11. Carbon monoxide emissions vs. brake power.
Sustainability 14 05313 g011
Figure 12. Smoke emissions vs. brake power.
Figure 12. Smoke emissions vs. brake power.
Sustainability 14 05313 g012
Figure 13. In-cylinder pressure vs. brake power.
Figure 13. In-cylinder pressure vs. brake power.
Sustainability 14 05313 g013
Figure 14. Heat release rate vs. brake power.
Figure 14. Heat release rate vs. brake power.
Sustainability 14 05313 g014
Table 1. Comparison of performance and emission responses for assorted nanofuel combinations and diesel.
Table 1. Comparison of performance and emission responses for assorted nanofuel combinations and diesel.
Type of Fuel BlendPerformanceEmission CharacteristicsRef.
BSFCBTECONOxHCPM/Smoke
Jatropha (Al2O3) 100 ppm[30]
Jatropha (Ti) 100 ppm[30]
Jatropha (CNT) 100 ppm[30]
Oenothera (GO) 100 ppm[31]
Garcinia gummi (CeO2, ZrO2, TiO2)[32]
WCO (CeO2 80 ppm)[33]
Neat Palm Oil (SiO2) 10 and 20 ppm[34]
Biodiesel—diesel (CNT, Ago 10, 20, 30 ppm)[35]
WCO—Diesel (CNT 20, 40 and 100 ppm)[36]
WOP (TiO2) 100 ppm[37]
Soyabean nanoemulsion (ZrO2) 100 ppm[38]
Nanoemulsion with Biodiesel (CNT 150 ppm)[39]
Mustard oil—nanoemulsion (TiO2) 50 ppm[40]
Rice bran—nanoemulsion (FeCl2)[41]
Crude oil—nanoemulsion (Al2O3) 80 ppm[42]
Table 2. Thermal, physical, and chemical properties of test fuel.
Table 2. Thermal, physical, and chemical properties of test fuel.
PropertiesD100
(100% Diesel)
B100
(100% Citronella)
SDE
(Surfactant-Free Diesel Emulsion)
SBE
(Surfactant-Free Bioemulsion)
SNBE
(Surfactant-Free Nano-Bioemulsion)
Calorific Value (MJ/kg)44.123843.00941.08941.183
Cetane number 4755465051
Kinematic Viscosity (cSt)2.94.93.453.914.15
Flash Pt (°C)6655686466
Density (kg/m3)820900843871844
Table 3. Engine specifications.
Table 3. Engine specifications.
MakeKirloskar TV-1
TypeFour-stroke engine
Cooling Water
Bae Engine Bore 86.6 mm
Bae Engine Stroke112 mm
Compression ratio 17.6:1
Power Rating5.1 kW
Speed1500 rpm
IT23deg before TDC
Nozzle 0.3 mm and three nozzles
Bowl Hemispherical
Table 4. Emission analyzer specifications.
Table 4. Emission analyzer specifications.
Instruments TypeManufacturerMeasuring RangeStandard Error
Smoke MeterAVL Smoke meterAVL India Pvt. Ltd. (Gurugram, India)0 to 100 HSU±0.1
Five-gas analyzer Krypton 290 five-gas analyzerSMS Auto Line Equipment Pvt. Ltd. (Chennai, India)CO—(0–10%)
CO2—(0 to 20%)
HC—(0 to10,000 ppm)
O2—(0 to 25%)
NOx—(0 to 5000 ppm)
±0.1
±0.1
±0.05
±0.02
Table 5. Test matrix.
Table 5. Test matrix.
Test No.Test
Abbreviations
LoadBase FuelWater
Concentration
Nanoadditive Concentration
1D100Variable 0–100% (Constant speed 1500 rpm, IT 23 degCA, IP 200 bar)Diesel fuel (100%)--
2B100Variable 0–100% (Constant speed 1500 rpm, IT 23 degCA, IP 200 bar)Citronella fuel (100%)--
3SDEVariable 0–100% (Constant speed 1500 rpm, IT 23 degCA, IP 200 bar)Diesel (95%) 5%-
4SBEVariable 0–100% (Constant speed 1500 rpm, IT 23 degCA, IP 200 bar)Citronella fuel (25%) + Diesel (70%)5%-
5SNBEVariable 0–100% (Constant speed 1500 rpm, IT 23 degCA, IP 200 bar)Citronella fuel (25%) + Diesel (70%)5%100 ppm
Table 6. Ambiguity examination.
Table 6. Ambiguity examination.
S.noMeasured Parameters Percent Uncertainty
1.BTE0.4%
2.BSFC0.6%
3.Pressure 0.2%
4.BP0.4%
5.Crank angle encoder0.05%
6.Engine Speed0.1%
7.Dynamometer Load0.2%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramalingam, K.; Perumal Venkatesan, E.; Aabid, A.; Baig, M. Assessment of CI Engine Performance and Exhaust Air Quality Outfitted with Real-Time Emulsion Fuel Injection System. Sustainability 2022, 14, 5313. https://doi.org/10.3390/su14095313

AMA Style

Ramalingam K, Perumal Venkatesan E, Aabid A, Baig M. Assessment of CI Engine Performance and Exhaust Air Quality Outfitted with Real-Time Emulsion Fuel Injection System. Sustainability. 2022; 14(9):5313. https://doi.org/10.3390/su14095313

Chicago/Turabian Style

Ramalingam, Krishnamoorthy, Elumalai Perumal Venkatesan, Abdul Aabid, and Muneer Baig. 2022. "Assessment of CI Engine Performance and Exhaust Air Quality Outfitted with Real-Time Emulsion Fuel Injection System" Sustainability 14, no. 9: 5313. https://doi.org/10.3390/su14095313

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop