Next Article in Journal
Co-Adjuvant Nanoparticles for Radiotherapy Treatments of Oncological Diseases
Next Article in Special Issue
Sundew-Inspired Adhesive Hydrogel Threads through Reversible Complexation of Polyphenol and Boronic Acid
Previous Article in Journal
Dynamic Analysis of Ball Screw Feed System with the Effects of Excitation Amplitude and Design Parameters
Previous Article in Special Issue
Soft Bio-Integrated Multifunctional Devices Using an Intrinsically Stretchable Conducting Nanomembrane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Analysis of Engine Performance and Exhaust Pollutant on a Single-Cylinder Diesel Engine Operated Using Moringa Oleifera Biodiesel

1
Department of Mechanical Engineering, School of Technology, Glocal University, Delhi-Yamunotri Marg, SH-57, Mirzapur Pole, Saharanpur District, Uttar Pradesh 247121, India
2
Department of Chemical, Polymer and Composite Materials Engineering, University of Engineering & Technology, Lahore 54890, Pakistan
3
Research Center for Advanced Materials Science (RCAMS), King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
4
Department of Mechanical Engineering, College of Engineering, King Khalid University, Abha 61421, Saudi Arabia
5
Department of Mechanical Engineering, University of Engineering and Technology, New Campus Lahore 54890, Pakistan
6
Department of Mechanical and Industrial Engineering, Tallinn University of Technology, Ehitajate Tee 5, 12616 Tallinn, Estonia
7
Malaysia-Japan International Institute of Technology (Mjiit), UTM Kuala Lumpur, Jalan Sultan Yahya Petra, Kuala Lumpur 54100, Malaysia
8
Institute for Advanced Studies, Univerity of Malaya, Kuala Lumpur 50603, Malaysia
9
School of Mechanical, Chemical and Materials Engineering, Adama Science and Technology University, Adama 1888, Ethiopia
10
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, Kuala Lumpur 50603, Malaysia
11
Mechanical Engineering Department, College of Engineering, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
12
Department of Mechanical Engineering, CMR Technical Campus, Hyderabad 501401, India
13
Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam
14
Mechanical Engineering Department, Faculty of Engineering, King Abdulaziz University, Jeddah 21589, Saudi Arabia
15
Department of Medical Research, China Medical University Hospital, China Medical University, Taichung 40402, Taiwan
16
Department of Civil and Environmental Engineering, Florida International University, Miami, FL 33174, USA
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(15), 7071; https://doi.org/10.3390/app11157071
Submission received: 2 July 2021 / Revised: 23 July 2021 / Accepted: 25 July 2021 / Published: 30 July 2021
(This article belongs to the Special Issue Biomaterial Synthesis and Application)

Abstract

:
In this investigation, biodiesel was produced from Moringa oleifera oil through a transesterification process at operating conditions including a reaction temperature of 60 °C, catalyst concentration of 1% wt., reaction time of 2 h, stirring speed of 1000 rpm and methanol to oil ratio of 8.50:1. Biodiesel blends, B10 and B20, were tested in a compression ignition engine, and the performance and emission characteristics were analyzed and compared with high-speed diesel. The engine was operated at full load conditions with engine speeds varying from 1000 rpm to 2400 rpm. All the performance and exhaust pollutants results were collected and analyzed. It was found that MOB10 produced lower BP (7.44%), BSFC (7.51%), and CO2 (7.7%). The MOB10 also reduced smoke opacity (24%) and HC (10.27%). Compared to diesel, MOB10 also increased CO (2.5%) and NOx (9%) emissions.

1. Introduction

The growth of the human population and a higher quality of living have increased global energy consumption. One of the most significant consumers of energy is the transportation field [1,2]. Transportation is heavily dependent on gasoline and diesel engines. Nevertheless, compared to gasoline, diesel engines are more cost-effective and energy-efficient [3,4]. Diesel has also become preferable because of its higher fuel efficiency, energy density, and lower carbon dioxide (CO2) emissions [5,6]. Thus, diesel engines provide higher mileage [7]. However, factors such as the increasing price of world crude oil, the decline in fossil fuel, and the increase in greenhouse gas emissions have forced researchers and scientists to find renewable and sustainable energy resources [8,9,10].
Furthermore, the health issues resulting from the exhaust of fossil fuel engines are causing alarm across the world [11,12]. Therefore, scientists and researchers are now searching for more renewable, sustainable and cleaner alternatives to replace fossil fuels [13,14]. Scientists and researchers are looking for ways to develop alternative fuels to deal with escalating energy demands [15,16,17].
In this regard, biodiesel or fatty acid methyl ester (FAME) is a potential substitute for petroleum-derived diesel in vehicles [18,19]. Biodiesel is usually produced by transesterification of edible oil or animal fats [20,21]. However, nowadays, biodiesels are also produced from the transesterification of non-edible oils, waste cooking oil, macroalgae, animal fats, and microalgae [22]. Thus, the sources used to produce biodiesel are sustainable and renewable [23]. Moreover, the biodiesel feedstock can be replenished by cultivating crops and rearing livestock.
In contrast, the sources of fossil fuel are non-renewable [24]. Biodiesel blends have been used without making any significant modifications to diesel engines. Biodiesels are potential alternatives for diesel because of their chemical and physical properties [25,26]. Biodiesel utilization in unmodified diesel engines slightly increases brake-specific fuel consumption and NOx emissions. However, biodiesel consumption significantly decreases CO, unburned hydrocarbon (HC), and particulate emissions due to more oxygen and the lack of aromatic content in biodiesels [27,28]. Various research studies have been performed on engines to examine the performance and emission characteristics of Calophyllum inophyllum and palm biodiesel blends [29].
A number of experimental studies have also been performed on the production of MOB and its physicochemical properties. However, there are no comparative studies to date regarding engine performance and the emission characteristics of MOB and its blends of 10% and 20% with diesel in an SCD. This provided the motivation and purpose of this study, which may also potentially assist in the future generation of alternative fuel. Therefore, the engine performance and emission characteristics resulting from regular fossil diesel and all Moringa oleifera methyl ester blended fuels were investigated.
Liaquat, Masjuki [30] experimentally examined exhaust gas emissions from a compression ignition engine fueled with palm biodiesel. A Bosch gas analyzer was used to analyze the engine exhaust emission parameters for 250 h at a 2000 rpm engine speed. A significant reduction in CO and CO2 emissions was recorded as the biodiesel concentration increased in blends. This was due to an excess amount of oxygen, which results in complete combustion occurring in the combustion chamber. Ozsezen and Canakci [31] examined the performance and emission characteristics of the CI engine palm oil methyl ester (biodiesel) blended with pure diesel. As a result, BSFC and brake power (BP) increased by 7.5 and 2.5%, respectively. A significant reduction of 86.89% in CO, 14.29% in HC, and 67.65% in smoke opacity were observed for palm biodiesel. However, the palm oil methyl ester enhanced NOx emissions by 22.1%. Sharon, Karuppasamy [32] used different palm biodiesel concentrations in diesel using a KIRLOSKAR TV-1 diesel engine. During the test, the engine’s load was changed from 20% to 100% at a constant engine speed of 800 rpm. The BSFC for palm biodiesel and pure diesel was found to be 0.315 and 0.2755 kg/kWh, respectively, at full load conditions. Biodiesel blends, B25, B50 and B75, showed a slightly higher BSFC of 2.6%, 8.9% and 9.3%, respectively, than pure diesel. Ong, Masjuki [33] used Calophyllum inophyllum biodiesel in their study to examine engine performance and the emission characteristics of a CI engine. According to their experimental results, the B10 blend showed a slight improvement in BTE as compared to diesel. However, EGT and BSFC were lower for this blend. Shehata and Razek [34] reported on the performance and emission characteristics of neat SOME at different engine speeds and loads. Resultantly, BSFC increased while BP, BTE, and torque were decreased as compared to diesel. For emissions, NOx was reduced, but CO and CO2 were increased. Roy, Wang [35] experimented using COME to monitor the performance and emission characteristics of a four-stroke two cylinders CI engine. The results suggested that BSFC of 10% COME blended fuel showed no significant increment, but further increasing biodiesel concentration in diesel fuel caused a slight increase in the BSFC, up to 2.3% compared to pure diesel. For emissions, CO emission was reduced for all percentage ratios of blended fuels, while similar trends were observed for NOx emission from the B10 blend and pure diesel. However, an increasing percentage of COME in blended fuel increased the NOx emission. Agarwal and Dhar [36] explored the performance, combustion and emission characteristics of Karanja oil methyl ester blended fuel (10%, 20%, and 50%). With regard to the engine performance, BSFC and EGT increased while BTE decreased as compared to diesel fuel. A significant reduction in HC and smoke opacity was observed with a slightly escalation in NOx emissions as compared to high-speed diesel. Both B10 and B20 blends delivered almost the same performance and emission characteristics.
Moringa oleifera Lamarck is a member of the Moringaceae family, a tropical plant that is easy to disseminate and grows to a height of around 5 m–10 m. It is widely grown in tropical countries and is mainly distributed in India, Bangladesh, Pakistan, Africa, South America, Arabic countries, the Philippines, Thailand, and Malaysia. The seeds of Moringa oleifera contain 40% of oil by weight, and the oil produced is a golden yellow color [37]. Several researchers have reported that Moringa oleifera oil contains a high oleic acid volume, that is, approximately 70% of the total fatty acid summary [38]. Compared to other feedstocks, Moringa oleifera oil is from a non-edible source, which gives it good potential for conversion into biodiesel without affecting food industries [39].
Rajaraman et al. [40] have discussed blended Moringa oleifera methyl ester (B20 and B100) and analyzed the engine performance and emission characteristics using a direct injection CI engine at full load conditions. The performance results show that the brake thermal efficiency of Moringa oleifera blended fuel decreased compared to standard diesel fuel due to its high viscosity and density, as well as the lower calorific value of the blended fuel. The emission results show that Moringa oleifera blended fuel produces lower CO, HC NOx, and PM than regular diesel fuel.
The current energy emergency has negatively affected the worldwide economy. The economies of numerous non-industrial nations have become uncompetitive because of the lack of usable energy. The present study is an effort to reduce the consumption of conventional fossil fuels. Moringa oil is derived from the seeds of Moringa oleifera, a small tree local to the mountains that can be used to prepare biodiesel via the transesterification process.

2. Materials and Methods

2.1. Biodiesel Preparation

The Moringa oleifera biodiesels were produced through an alkaline-catalyzed transesterification process. Firstly, the Moringa oleifera crude oil was mixed with 25% vol. of methanol and 1% wt. of KOH. A temperature of 60 °C and a stirring speed of 1000 rpm were maintained for 2 h. These conditions, were used to ensure that a homogenous mixture of Moringa oleifera oil, methanol, and potassium hydroxide was obtained, and so that the transesterification process would produce a desirable yield rate. Once the transesterification process was finished, the biodiesel was separated via a separating funnel. After 12 h, the product was transformed into two layers. Two immiscible layers of liquid formed in the separating funnel, the top layer was the methyl ester (biodiesel), and the bottom layer consisted of impurities and glycerin. The bottom layer was drained from the separating funnel, and following this, 50% vol. of distilled water at a temperature of 60 °C was used to spray and wash each methyl ester. Next, the methyl ester was rinsed with hot DW until Moringa oleifera methyl ester was cleaned of all impurities. Then, by using a rotary evaporator, methyl ester was dried and then purified via filter paper. After the purification process, MOME was mixed with diesel at various ratios to produce the biodiesel blends. The blends prepared in this study were as follows: MOB10, MOB20, and diesel. A total of three samples were prepared for the study, comprising two samples of biodiesel blends and one sample of pure diesel.

2.2. Composition of Biodiesel

The FAC of the MOME was analyzed using a gas chromatography (GC) system, Agilent 7890 series, USA. Specifications and operating mode of GC system are summarized in Table 1. The FAC of the MOME is presented in Table 2. The amount of esters, methyl linoleate, monoglycerides, diglycerides, triglycerides, and free and total glycerin was measured according to the EN14103 standard.

2.3. Physiochemical Characteristics of Biodiesel

It was imperative to measure Moringa oleifera biodiesel (MOME) characteristics and their blends (MOB10 and MOB20) to assess the quality and suitability of these fuels for diesel engines. Each biodiesel has different physicochemical properties depending on feedstock type and biodiesel production process, post-production treatment, and fatty acid composition of the biodiesel. Hence, different biodiesel and biodiesel blends shows different effects on the CIDE’s performance and exhaust emissions. In this study, the physicochemical properties (i.e., density, viscosity index, flash point, acid number, oxidation stability, pour point, cloud point, and CFPP and kinematic viscosity) of MOME and the blends were measured using ASTM standards. Results of the measured properties are summarized in Table 3. A Stabinger viscometer (Model: SVM 3000, Anton Paar, UK) was utilized to measure density (at 15 and 40 °C) and kinematic viscosity (at 40 and 100 °C). A bomb calorimeter (Model: C2000 Basic, IKA, UK) was utilized for calorific value measurement.
C C I = 0.0892 T 10 N + 45.2 + 0.131 + 0.901   B T 50 N + 0.0523 0.420 B T 90 N + 107 B + 60 B 2 + 0.00049 T 10 N 2 T 90 N 2
where,
  • CCI = calculated cetane index, D = density at 15 °C (g/mL), B = [e(−3.5)(DN)] − 1, DN = D − 0.85, T10N = T10 − 215, T50N = T50 − 260, and T90N = T90 − 310.
The cetane index of MOME and the blends was calculated based on the recovered temperature values at 10%, 50%, and 90% (T10, T50, and T90) and the fuel density at 15 °C (D) according to ASTM D4737 standard test methods, which is given by the equation in [41].

2.4. Engine Setup

A naturally aspirated, single-cylinder, four-stroke, direct injection diesel engine with an eddy current dynamometer was used in this study. Technical specifications for the tested engine are listed in Table 4. The experimental layout of the test engine is displayed in Figure 1. Engine tests were carried out in full load conditions in triplicates, and the engine speed varied from 1000 to 2400 RPM with an interval of 200 rpm. The exhaust emission parameters (smoke opacity, NOx, HC, and CO) were analyzed using an AVL exhaust gas analyzer (Model: DiCom 4000, AVL Ditest, Austria). In Table 5, the technical specifications of the used gas analyzer (AVL exhaust gas analyzer) are listed. First, the neat diesel fuel was utilized to bring the engine to a stable operating condition. Once this condition was reached, the biodiesel blended fuel was used for investigation. The engine was run for a few minutes, and then the residual diesel was drained. Data acquisition was performed after the drainage of residual diesel. This practice was repeated for each biodiesel blend. After one test was completed for the biodiesel blend, the engine was operated via diesel. This practice helped to drain the residual biodiesel blend used in the previous test from the fuel line.

2.4.1. BTE and BSFC

Brake thermal efficiency (BTE) is defined as the brake power of an internal combustion engine as a function of the heat input obtained from fuel burning. BTE is calculated using the formula given below:
η B T E = B P × 3600 × 100 m × C v %
where, BP is brake power, m is mass flow rate and C v is the calorific value of the tested fuel.
Measurement of the fuel efficiency of any engine that burns the fuel and generates rotational or shaft power is BSFC.

2.4.2. Smoke Opacity, HC, CO and NOx

Smoke opacity is defined as the amount of light concealed by the particulate matter or soot particles omitted from the combustion of diesel. Smoke opacity reflects the presence of soot in the exhaust gases. Smoke meters, also known as opacity meters, measure the amount of light blocked in the smoke emitted by vehicles. The smoke in engine exhaust depends mainly on the combustion process, formation of the air–fuel mixture, amount of fuel injected before the ignition process, and oxygen content of fuel [42]. In general, incomplete fuel combustion leads to higher smoke opacity. Smoke opacity is influenced by the engine speed, engine load, fuel viscosity, cetane number, air turbulence, and spray pattern in the cylinder [43,44]. HC is produced in the diesel engine when there is an over-rich mixture or over-lean mixtures. Physicochemical properties of the fuel, fuel injection, and engine operating conditions also play a vital role in forming HC emissions. Incomplete combustion leads to CO formation. The lower oxygen content of diesel results in higher CO emissions.
On the other hand, vegetable oil-based biodiesels have a higher oxygen content in their chains, which leads to complete combustion, and hence, lower CO emissions. NOx emissions are influenced by the fuel’s spray characteristics and oxygen content, and adiabatic flame temperature. Spray fuel characteristics refer to the size and momentum of fuel droplets, degree of mixing between fuel droplets with air, penetration rate and evaporation, and radiant heat transfer rate [45,46,47,48,49].
Figure 1. Schematic diagram of the diesel engine (Reprinted with permission from Elsevier, ref. [50]).
Figure 1. Schematic diagram of the diesel engine (Reprinted with permission from Elsevier, ref. [50]).
Applsci 11 07071 g001
Table 4. Specifications of diesel engine [50] (Adapted with permission from the Elsevier).
Table 4. Specifications of diesel engine [50] (Adapted with permission from the Elsevier).
No.DescriptionSpecificationsNo.DescriptionSpecifications
1Engine ModelTF 120M5Compression ratio17.7:1
2Displacement (cm3)6386Maximum power8.8 kW/2400 rpm
3Bore (mm)927Cooling systemWater cooling
4Stroke (mm)968Fuel systemPump line nozzle
Table 5. Specifications of emission analyzer.
Table 5. Specifications of emission analyzer.
EquipmentMeasurementMethodMeasurement RangeResolution
AVL DiGas 4000/AVL DiCom 4000CONon-dispersive infrared0–10% vol.0.01 vol%
HCNon-dispersive infrared0–20,000 ppm vol.1 ppm
NOxElectrochemical detector0–5000 ppm vol.1 ppm
AVL DiSmoke 4000/AVL DiCom 4000OpacityPhotodiode receiver0–100% vol.0.1%

3. Results

3.1. Engine Performance Characteristics

3.1.1. Brake Power (BP)

The performance of CI diesel engines relies on the characteristics of the fuel utilized for the testing engine and fuel injection system. The fuel characteristics include kinematic viscosity, density, oxygen content, and calorific value [48,49]. Figure 2 shows the brake power (BP) of Moringa oleifera biodiesel blends and diesel at different engine speeds. According to the results, BP increases progressively with engine speed until 2200 rpm and then decreases. Consequently, diesel fuel has the highest BP (5.43 kW) at 2200 rpm.
In contrast, the MOB20 blend has the lowest BP (4.68 kW). The average BP is higher for diesel than MOB10 and MOB20 by 12.18% and 17.32%, respectively. The average BP is lower for MOB20 than MOB10 and diesel by 6.85% and 7.17%, respectively. This may be attributed to the larger HHV of biodiesel blends [50]. The MOB10 blend has the highest HHV in comparison with other biodiesel blends examined in this study. Besides, the fuel’s physicochemical properties affect the spray formation during fuel injection, which in turn, affects combustion [51]. Lower viscosity and density of the MOB10 blend may result in loss of engine power due to more significant fuel pump leakage than other fuel blends [52]. Generally, fuels with higher viscosities can reduce fuel pump leakages [53].

3.1.2. Brake Specific Fuel Consumption (BSFC)

Figure 3 shows the BSFC of Moringa oleifera biodiesel blends and diesel at various engine speeds. Diesel fuel shows lower BSFC as compared to biodiesel blends. The MOB20 blend has the highest average BSFC, with a value of 0.6115 kg/kWh. The MOB10 and MOB20 blends have a higher average BSFC than diesel by 7.03% and 12.75%, respectively. In general, biodiesels have a larger HHV due to the fuel-borne oxygen. Hence, a higher amount of fuel mass needs to be injected from the fuel injection pump into the engine due to biodiesel’s higher density than diesel. More biodiesel needs to be injected into the combustion chamber for the same power output as diesel according to volumetric efficiency. The higher kinematic viscosity of Moringa oleifera biodiesel blends is the leading cause of poor air–fuel mixing resulting from slower fuel atomization. Higher density and lower calorific values than diesel are factors that lead to the higher BSFC for biodiesel blends, especially those containing higher concentrations of biodiesels [54].

3.1.3. Brake Thermal Efficiency (BTE)

At full load conditions, the BTE increases, but it declines with an increasing compression ratio; it acts similar to the indicated thermal efficiency. Figure 4 illustrates the engine brake thermal efficiencies for MOB10, MOB20, and diesel fuels. According to our observations, the average brake thermal efficiency for MOB10 was 2% higher than pure diesel. However, the average brake thermal efficiency for MOB20 was 3.45% lower as compared to pure diesel. The curves were plotted by averaging three readings. Various researchers have found similar results whereby the brake thermal efficiency of the biodiesel blends was comparable with pure diesel’s thermal efficiency [55,56]. In addition, they have found that preheating biodiesel fuel before injection increases the brake thermal efficiency.

3.2. Emission Characteristics

3.2.1. Unburned Hydrocarbon Emissions (HC)

Figure 5 shows HC emissions for MOB and its blends with diesel at different engine speeds. Average HC emission is higher for diesel than that for MOB10 and MOB20 by 6.71% and 8.79%, respectively. Furthermore, the fuel blends containing 20% of biodiesel have higher HC emissions at low speeds compared to those containing 10% of biodiesel.
Moreover, it can be observed that each tested fuel had higher HC emissions when the engine was running at lower speeds. Conversely, the amount of HC emissions decreased when the engine’s speed was higher. The lean air–fuel mixture is the primary reason for more HC emissions at lower engine speeds as well as poor fuel distribution. The lower temperature and presence of excess air are responsible for lean air-fuel mixtures [42]. Over-rich and over-lean air–fuel mixtures are typical during heterogeneous combustion in diesel engines, which leads to HC emissions. The oxygen content of biodiesels generally leads to lower HC emissions than diesel at high engine speeds due to improved fuel combustion [57].

3.2.2. Carbon Dioxide Emissions (CO2)

Figure 6 shows the CO2 emissions of MOB blends and diesel at various engine speeds. CO2 emissions from the engine’s exhaust reached a maximum value with MOB20 and were reduced when the biodiesel concentration in the fuel was decreased. The average CO2 emission values for MOB10, MOB20, and diesel were 5.693%, 6.124%, and 6%; the curves were plotted by averaging three readings. MOB20 showed higher CO2 emissions than diesel and MOB10 due to more oxygen in MOB20 relative to neat diesel and MOB10. The higher amount of oxygen in the biodiesel increased the oxidation and combustion process. Due to the higher amount of oxygen, the excess amount of CO is converted to CO2 [52].

3.2.3. Carbon Monoxide Emissions (CO)

According to previous reports, oxygenated fuels reduce up to 30% of CO emissions compared to diesel—however, the magnitude of the reduction depends on the engine type and age, and ambient conditions [58,59]. Figure 7 displays CO emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds. It can be observed that the MOB20 blend produces the highest amount of CO emissions at an engine speed of 1400 rpm.
On the other hand, the MOB20 blend produces the lowest CO emissions at 2400 rpm and the lowest average CO emission in this study. The average CO emission of diesel is 0.82% higher than that for MOB20. However, the average CO emission is 1.99% lower than that for MOB10. In general, for the same blend ratio, the CO emissions decreased as the engine’s speed changed from a lower to a higher value for all fuels. This is due to higher oxygen content and the higher cetane number of biodiesel fuel than diesel fuel. Higher cylinder pressure and temperature promote complete combustion at high engine speed, especially for biodiesel fuel that contains higher oxygen content. This enables the conversion of CO to CO2, reducing the amount of CO emission [60,61,62].

3.2.4. Nitrogen Oxide Emissions (NOx)

Many studies have shown that biodiesel fuels produce higher engine NOx emissions compared to diesel [63,64,65,66,67,68,69,70]. Figure 8 displays NOx emissions of MOB blends and diesel at various engine speeds. Several factors influence the production of NOx, and one of them is the oxygen content. In general, vegetable oil-based biodiesels have higher oxygen content (with a difference of 12% relative to diesel) as well as low nitrogen content. This results in higher NOx emissions when there is an increase in the combustion chamber temperature, which improves the combustion process [66]. The MOB20 blend has the highest NOx emissions (416 ppm) at an engine speed of 2400 rpm. Moringa oleifera biodiesel has more oxygen content as compared to neat diesel fuel. Besides, NOx emissions increase with an increase in the concentration of biodiesel in fuel blends. Average NOx emissions are lower for diesel compared to MOB10 and MOB20 by 4.71% and 8.12%, respectively. Abedin et al. [71] found that fuel blends containing 10% and 20% of palm biodiesel reduce NOx emissions by approximately 3.3%. Rahman et al. [70] discovered that a fuel blend containing 10% biodiesel produces higher NOx emissions by 9% relative to diesel. In general, biodiesels have a higher adiabatic flame temperature because of their high unsaturated fatty acid content, leading to more NOx emissions [69]. The higher viscosity and density of biodiesels are also responsible for higher NOx emissions [33].

3.2.5. Smoke Opacity

Figure 9 displays the smoke opacity for Moringa oleifera biodiesel and its blends with diesel tested at different engine speeds. For diesel, the average smoke opacity is higher than MOB10 and MOB20 by 33.49% and 22.73%, respectively. The MOB10 blend has the lowest average smoke opacity (32.2%) compared to MOB20 and neat diesel. At higher engine speeds, the smoke opacity of MOB blends increased significantly. Several studies have shown that the smoke opacity was lower due to more oxygen contents in biodiesel. A lower ratio of carbon–hydrogen and non-availability of aromatic compounds in the biodiesel reduced the smoke emissions [72]. According to Gumus and Kasifoglu [73], more oxygen in biodiesel blends can reduce smoke exposure in exhaust gasses. Zhang et al. [74] found that combustion of biodiesel blends occurs earlier than diesel. Smoke emissions are reduced due to advanced injection timing, which results from the combustion process’s quick start. In contrast, diesel has higher sulfur content than biodiesel blends, which is the main reason for high smoke opacity [75].

4. Conclusions

The performance and exhaust emission characteristics of Moringa oleifera biodiesel blends were analyzed in this study. The results of the experimental investigation show that the MOB10 blend is the best blend ratio based on the following criteria:
  • At optimum speed, the BTE for MOB10 and MOB20 was 2.54% higher and 3.45% lower, respectively, than that of pure diesel.
  • MOB10 and MOB20 blends had a higher average BSFC than diesel by 7.03% and 12.75%, respectively, due to the higher density and lower calorific values of biodiesel blends.
  • MOB10 produced slightly lower BP when compared to diesel, by 0.26 kW. The MOB20 blend was the worst performer, producing less usable power than diesel by 0.36 kW.
  • The average HC emission for MOB10 and MOB20 were lower than diesel, with a difference of 8 ppm.
  • The average NOx emission for blended fuels was significantly higher than the neat diesel, and the MOB20 blend produces more NOx emissions due to increased oxygen content in fuel blends.
  • MOB10 produced lower smoke opacity than those of neat diesel and MOB20 due to good combustion.
  • Therefore, MOB10 is suitable to use in conventional compression-ignition diesel engines.
Future Recommendation: The NOx emissions slightly increased in the combustion of biodiesel blends compared to conventional diesel. The researchers could pursue this work using different fuel additives such as nanoparticles or alcohols to reduce NOx emissions.

Author Contributions

Conceptualization, M.A.M. and M.E.M.S.; methodology, M.A.M., T.A. and M.E.M.S.; software, M.A.M., W.A., H.M.K., L.R. and M.E.M.S.; validation, A.H., M.S.G. and T.M.Y.K.; formal analysis, investigation, resources, data curation, writing—original draft preparation, M.A.M., L.R., A.A., M.S.G., U.I., T.M.I. and M.E.M.S.; review, editing and support, A.E., K.S., M.A.A.B. and M.F.; supervision, M.A.M.; project administration, M.S.G. and M.R.S.; funding acquisition, T.M.Y.K., A.E. and M.R.S. All authors have read and agreed to the published version of the manuscript.

Funding

The Deanship of Scientific Research at King Khalid University, Grant Number R.G.P. 2/147/42. Taif University Researchers Supporting Project number (TURSP–2020/40), Taif University, Taif, Saudi Arabia.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through Grant Number R.G.P. 2/147/42. This work was supported by Taif University Researchers Supporting Project number (TURSP–2020/40), Taif University, Taif, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

MOBMoringa Oleifera Biodiesel
CICompression Ignition
BPBrake Power
BTEBrake Thermal Efficiency
BSFCBrake Specific Fuel Consumption
MOB10Moringa Oil Biodiesel 10% + Diesel 90%
MOB20Moringa Oil Biodiesel 20% + Diesel 80%
COCarbon Monoxide
CO2Carbon Dioxide
NOxNitrogen Oxides
HCHydrocarbons
EGTExhaust Gas Temperature
FACFatty Acid Composition
MOMEMoringa Oleifera Methyl Ester
COMECanola Oil Methyl Ester
SOMESoybean Oil Methyl Ester
SCDESingle Cylinder Diesel Engine
CFPPCold Filter Plugging Point
HHVHigher Heating Value
DWDeionized Water

References

  1. Wategave, S.; Banapurmath, N.; Sawant, M.; Soudagar, M.E.M.; Mujtaba, M.; Afzal, A.; Basha, J.S.; Alazwari, M.A.; Safaei, M.R.; Elfasakhany, A. 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]
  2. Li, Z.; Sarafraz, M.; Mazinani, A.; Hayat, T.; Alsulami, H.; Goodarzi, M. Pool boiling heat transfer to CuO-H2O nanofluid on finned surfaces. Int. J. Heat Mass Transf. 2020, 156, 119780. [Google Scholar] [CrossRef]
  3. Sarafraz, M.M.; Christo, F.C. Thermodynamic assessment and techno-economic analysis of a liquid indium-based chemical looping system for biomass gasification. Energy Convers. Manag. 2020, 225, 113428. [Google Scholar] [CrossRef] [PubMed]
  4. Ahmadi, A.A.; Arabbeiki, M.; Ali, H.M.; Goodarzi, M.; Safaei, M.R. Configuration and optimization of a minichannel using water–alumina nanofluid by non-dominated sorting genetic algorithm and response surface method. Nanomaterials 2020, 10, 901. [Google Scholar] [CrossRef] [PubMed]
  5. Di Sarli, V.; Di Benedetto, A. Combined effects of soot load and catalyst activity on the regeneration dynamics of catalytic diesel particulate filters. AIChE J. 2018, 64, 1714–1722. [Google Scholar] [CrossRef]
  6. Di Sarli, V.; Landi, G.; Lisi, L.; Di Benedetto, A. Ceria-coated diesel particulate filters for continuous regeneration. AIChE J. 2017, 63, 3442–3449. [Google Scholar] [CrossRef]
  7. Bala Prasad, K.; Dhana Raju, V.; Ahamad Shaik, A.; Gopidesi, R.K.; Sreekara Reddy, M.B.S.; Soudagar, M.E.M.; Mujtaba, M.A. Impact of injection timings and exhaust gas recirculation rates on the characteristics of diesel engine operated with neat tamarind biodiesel. Energy Sources Part A Recovery Util. Environ. Eff. 2021, 1–19. [Google Scholar] [CrossRef]
  8. Rahmanian, B.; Safaei, M.R.; Kazi, S.N.; Ahmadi, G.; Oztop, H.F.; Vafai, K. Investigation of pollutant reduction by simulation of turbulent non-premixed pulverized coal combustion. Appl. Therm. Eng. 2014, 73, 1222–1235. [Google Scholar] [CrossRef]
  9. Kang, J.; Chu, S.; Lee, J.; Kim, G.; Min, K. Effect of operating parameters on diesel/propane dual fuel premixed compression ignition in a diesel engine. Int. J. Automot. Technol. 2018, 19, 27–35. [Google Scholar] [CrossRef]
  10. Hussain, A.; Mehdi, S.M.; Akhtar, M.; Ani, F.N.; Ahmed, I. Combustion Performance of Diesel Palm Olein Fuel: A Combined CFD and Experimental Approach. Arab. J. Sci. Eng. 2018, 43, 1291–1300. [Google Scholar] [CrossRef]
  11. Soudagar, M.E.M.; Afzal, A.; Safaei, M.R.; Manokar, A.M.; EL-Seesy, A.I.; Mujtaba, M.; Samuel, O.D.; Badruddin, I.A.; Ahmed, W.; Shahapurkar, K. Investigation on the effect of cottonseed oil blended with different percentages of octanol and suspended MWCNT nanoparticles on diesel engine characteristics. J. Therm. Anal. Calorim. 2020, 1–18. [Google Scholar] [CrossRef]
  12. Bahiraei, M.; Jamshidmofid, M.; Goodarzi, M. Efficacy of a hybrid nanofluid in a new microchannel heat sink equipped with both secondary channels and ribs. J. Mol. Liq. 2019, 273, 88–98. [Google Scholar] [CrossRef]
  13. Sarafraz, M.M.; Goodarzi, M.; Tlili, I.; Alkanhal, T.A.; Arjomandi, M. Thermodynamic potential of a high-concentration hybrid photovoltaic/thermal plant for co-production of steam and electricity. J. Therm. Anal. Calorim. 2021, 143, 1389–1398. [Google Scholar] [CrossRef]
  14. Nazari, S.; Karami, A.; Bahiraei, M.; Olfati, M.; Goodarzi, M.; Khorasanizadeh, H. A novel technique based on artificial intelligence for modeling the required temperature of a solar bread cooker equipped with concentrator through experimental data. Food Bioprod. Process. 2020, 123, 437–449. [Google Scholar] [CrossRef]
  15. Bagherzadeh, S.A.; Jalali, E.; Sarafraz, M.M.; Akbari, O.A.; Karimipour, A.; Goodarzi, M.; Bach, Q.-V. Effects of magnetic field on micro cross jet injection of dispersed nanoparticles in a microchannel. Int. J. Numer. Methods Heat Fluid Flow 2019, 30. [Google Scholar] [CrossRef]
  16. Bagherzadeh, S.A.; D’Orazio, A.; Karimipour, A.; Goodarzi, M.; Bach, Q.-V. A novel sensitivity analysis model of EANN for F-MWCNTs–Fe3O4/EG nanofluid thermal conductivity: Outputs predicted analytically instead of numerically to more accuracy and less costs. Phys. A Stat. Mech. Appl. 2019, 521, 406–415. [Google Scholar] [CrossRef]
  17. Akkoli, K.; Banapurmath, N.; Shivashimpi, M.; Soudagar, M.E.M.; Badruddin, I.A.; Alazwari, M.A.; Yaliwal, V.; Mujtaba, M.; Akram, N.; Goodarzi, M. Effect of injection parameters and producer gas derived from redgram stalk on the performance and emission characteristics of a diesel engine. Alex. Eng. J. 2021, 60, 3133–3142. [Google Scholar] [CrossRef]
  18. Khan, H.M.; Iqbal, T.; Yasin, S.; Irfan, M.; Kazmi, M.; Fayaz, H.; Mujtaba, M.; Ali, C.H.; Kalam, M.; Soudagar, M.E.M. Production and utilization aspects of waste cooking oil based biodiesel in Pakistan. Alex. Eng. J. 2021, 60, 5831–5849. [Google Scholar] [CrossRef]
  19. Mahdisoozani, H.; Mohsenizadeh, M.; Bahiraei, M.; Kasaeian, A.; Daneshvar, A.; Goodarzi, M.; Safaei, M.R. Performance enhancement of internal combustion engines through vibration control: State of the art and challenges. Appl. Sci. 2019, 9, 406. [Google Scholar] [CrossRef] [Green Version]
  20. Choudhary, A.K.; Chelladurai, H.; Kannan, C. Optimization of Combustion Performance of Bioethanol (Water Hyacinth) Diesel Blends on Diesel Engine Using Response Surface Methodology. Arab. J. Sci. Eng. 2015, 40, 3675–3695. [Google Scholar] [CrossRef]
  21. Soudagar, M.E.M.; Banapurmath, N.; Afzal, A.; Hossain, N.; Abbas, M.M.; Haniffa, M.A.C.M.; Naik, B.; Ahmed, W.; Nizamuddin, S.; Mubarak, N. Study of diesel engine characteristics by adding nanosized zinc oxide and diethyl ether additives in Mahua biodiesel–diesel fuel blend. Sci. Rep. 2020, 10, 1–17. [Google Scholar] [CrossRef]
  22. Sadeghinezhad, E.; Kazi, S.; Sadeghinejad, F.; Badarudin, A.; Mehrali, M.; Sadri, R.; Safaei, M.R. A comprehensive literature review of bio-fuel performance in internal combustion engine and relevant costs involvement. Renew. Sustain. Energy Rev. 2014, 30, 29–44. [Google Scholar] [CrossRef]
  23. Goodarzi, M.; Toghraie, D.; Reiszadeh, M.; Afrand, M. Experimental evaluation of dynamic viscosity of ZnO–MWCNTs/engine oil hybrid nanolubricant based on changes in temperature and concentration. J. Therm. Anal. Calorim. 2019, 136, 513–525. [Google Scholar] [CrossRef]
  24. Hussain, F.; Alshahrani, S.; Abbas, M.M.; Khan, H.M.; Jamil, A.; Yaqoob, H.; Soudagar, M.E.M.; Imran, M.; Ahmad, M.; Munir, M. Waste Animal Bones as Catalysts for Biodiesel Production; A Mini Review. Catalysts 2021, 11, 630. [Google Scholar] [CrossRef]
  25. Hosseini, S.M.; Safaei, M.R.; Goodarzi, M.; Alrashed, A.A.; Nguyen, T.K. New temperature, interfacial shell dependent dimensionless model for thermal conductivity of nanofluids. Int. J. Heat Mass Transf. 2017, 114, 207–210. [Google Scholar] [CrossRef]
  26. Senthil, R.; Silambarasan, R. Influence of Injection Timing and Compression Ratio on Performance, Emission and Combustion Characteristics of Jatropha Methyl Ester Operated Di Diesel Engine. Iran. J. Sci. Technol. Trans. Mech. Eng. 2015, 39, 61–76. [Google Scholar]
  27. Agarwal, A.K. Biofuels (alcohols and biodiesel) applications as fuels for internal combustion engines. Prog. Energ. Combust. 2007, 33, 233–271. [Google Scholar] [CrossRef]
  28. S Gavhane, R.; M Kate, A.; Pawar, A.; Safaei, M.R.; M Soudagar, M.E.; Mujtaba Abbas, M.; Muhammad Ali, H.; R Banapurmath, N.; Goodarzi, M.; Badruddin, I.A. Effect of zinc oxide nano-additives and soybean biodiesel at varying loads and compression ratios on VCR diesel engine characteristics. Symmetry 2020, 12, 1042. [Google Scholar] [CrossRef]
  29. Lapuerta, M.; Armas, O.; Rodriguez-Fernandez, J. Effect of biodiesel fuels on diesel engine emissions. Prog. Energ. Combust. 2008, 34, 198–223. [Google Scholar] [CrossRef]
  30. Khan, H.; Soudagar, M.E.M.; Kumar, R.H.; Safaei, M.R.; Farooq, M.; Khidmatgar, A.; Banapurmath, N.R.; Farade, R.A.; Abbas, M.M.; Afzal, A. Effect of nano-graphene oxide and n-butanol fuel additives blended with diesel—Nigella sativa biodiesel fuel emulsion on diesel engine characteristics. Symmetry 2020, 12, 961. [Google Scholar] [CrossRef]
  31. Mosarof, M.H.; Kalam, M.A.; Masjuki, H.H.; Alabdulkarem, A.; Ashraful, A.M.; Arslan, A.; Rashedul, H.K.; Monirul, I.M. Optimization of performance, emission, friction and wear characteristics of palm and Calophyllum inophyllum biodiesel blends. Energy Convers. Manag. 2016, 118, 119–134. [Google Scholar] [CrossRef]
  32. Liaquat, A.; Masjuki, H.; Kalam, M.; Fazal, M.; Khan, A.F.; Fayaz, H.; Varman, M. Impact of palm biodiesel blend on injector deposit formation. Appl. Energy 2013, 111, 882–893. [Google Scholar] [CrossRef]
  33. Ozsezen, A.N.; Canakci, M. Determination of performance and combustion characteristics of a diesel engine fueled with canola and waste palm oil methyl esters. Energy Convers. Manag. 2011, 52, 108–116. [Google Scholar] [CrossRef]
  34. Sharon, H.; Karuppasamy, K.; Kumar, D.S.; Sundaresan, A. A test on DI diesel engine fueled with methyl esters of used palm oil. Renew. Energy 2012, 47, 160–166. [Google Scholar] [CrossRef]
  35. Ong, H.C.; Masjuki, H.; Mahlia, T.; Silitonga, A.; Chong, W.; Leong, K. Optimization of biodiesel production and engine performance from high free fatty acid Calophyllum inophyllum oil in CI diesel engine. Energy Convers. Manag. 2014, 81, 30–40. [Google Scholar] [CrossRef]
  36. Shehata, M.S.; Razek, S.M.A. Experimental investigation of diesel engine performance and emission characteristics using jojoba/diesel blend and sunflower oil. Fuel 2011, 90, 886–897. [Google Scholar] [CrossRef]
  37. Roy, M.M.; Wang, W.; Bujold, J. Biodiesel production and comparison of emissions of a DI diesel engine fueled by biodiesel–diesel and canola oil—Diesel blends at high idling operations. Appl. Energy 2013, 106, 198–208. [Google Scholar] [CrossRef]
  38. Agarwal, A.K.; Dhar, A. Experimental investigations of performance, emission and combustion characteristics of Karanja oil blends fuelled DICI engine. Renew. Energy 2013, 52, 283–291. [Google Scholar] [CrossRef]
  39. Atabani, A.E.; Silitonga, A.S.; Ong, H.C.; Mahlia, T.M.I.; Masjuki, H.H.; Badruddin, I.A.; Fayaz, H. Non-edible vegetable oils: A critical evaluation of oil extraction, fatty acid compositions, biodiesel production, characteristics, engine performance and emissions production. Renew. Sust. Energ. Rev. 2013, 18, 211–245. [Google Scholar] [CrossRef]
  40. Azam, M.M.; Waris, A.; Nahar, N.M. Prospects and potential of fatty acid methyl esters of some non-traditional seed oils for use as biodiesel in India. Biomass Bioenerg. 2005, 29, 293–302. [Google Scholar] [CrossRef]
  41. Mofijur, M.; Masjuki, H.H.; Kalam, M.A.; Rasul, M.G.; Atabani, A.E.; Hazrat, M.A.; Mahmudul, H.M. Effect of Biodiesel-Diesel Blending on Physico-Chemical Properties of Biodiesel produced from Moringa oleifera. Procedia Eng. 2015, 105, 665–669. [Google Scholar] [CrossRef]
  42. Rajaraman, S.; Yashwanth, G.K.; Rajan, T.; Kumaran, R.S.; Raghu, P. Experimental Investigations of Performance and Emission Characteristics of Moringa Oil Methyl Ester and Its Diesel Blends in a Single Cylinder Direct Injection Diesel Engine. In ASME International Mechanical Engineering Congress and Exposition; ASME: Lake Buena Vista, FL, USA, 2010; Volume 3, pp. 27–34. [Google Scholar]
  43. Aleme, H.G.; Barbeira, P.J. Determination of flash point and cetane index in diesel using distillation curves and multivariate calibration. Fuel 2012, 102, 129–134. [Google Scholar] [CrossRef] [Green Version]
  44. Sakthivel, G. Prediction of CI engine performance, emission and combustion characteristics using fish oil as a biodiesel at different injection timing using fuzzy logic. Fuel 2016, 183, 214–229. [Google Scholar] [CrossRef]
  45. Dhamodaran, G.; Krishnan, R.; Pochareddy, Y.K.; Pyarelal, H.M.; Sivasubramanian, H.; Ganeshram, A.K. A comparative study of combustion, emission, and performance characteristics of rice-bran-, neem-, and cottonseed-oil biodiesels with varying degree of unsaturation. Fuel 2017, 187, 296–305. [Google Scholar] [CrossRef]
  46. Razzaq, L.; Farooq, M.; Mujtaba, M.; Sher, F.; Farhan, M.; Hassan, M.T.; Soudagar, M.E.M.; Atabani, A.; Kalam, M.; Imran, M. Modeling viscosity and density of ethanol-diesel-biodiesel ternary blends for sustainable environment. Sustainability 2020, 12, 5186. [Google Scholar] [CrossRef]
  47. Nikkhah, V.; Sarafraz, M.; Hormozi, F.; Peyghambarzadeh, S. Particulate fouling of CuO–water nanofluid at isothermal diffusive condition inside the conventional heat exchanger-experimental and modeling. Exp. Therm. Fluid Sci. 2015, 60, 83–95. [Google Scholar] [CrossRef]
  48. Sarafraz, M.; Hormozi, F.; Peyghambarzadeh, S. Pool boiling heat transfer to aqueous alumina nano-fluids on the plain and concentric circular micro-structured (CCM) surfaces. Exp. Therm. Fluid Sci. 2016, 72, 125–139. [Google Scholar] [CrossRef]
  49. Salari, E.; Peyghambarzadeh, S.; Sarafraz, M.; Hormozi, F.; Nikkhah, V. Thermal behavior of aqueous iron oxide nano-fluid as a coolant on a flat disc heater under the pool boiling condition. Heat Mass Transf. 2017, 53, 265–275. [Google Scholar] [CrossRef]
  50. Mujtaba, M.; Kalam, M.; Masjuki, H.; Gul, M.; Soudagar, M.E.M.; Ong, H.C.; Ahmed, W.; Atabani, A.; Razzaq, L.; Yusoff, M. Comparative study of nanoparticles and alcoholic fuel additives-biodiesel-diesel blend for performance and emission improvements. Fuel 2020, 279, 118434. [Google Scholar] [CrossRef]
  51. Habibullah, M.; Fattah, I.M.R.; Masjuki, H.H.; Kalam, M.A. Effects of Palm-Coconut Biodiesel Blends on the Performance and Emission of a Single-Cylinder Diesel Engine. Energy Fuel 2015, 29, 734–743. [Google Scholar] [CrossRef]
  52. Sarafraz, M.M.; Christo, F. Sustainable three-stage chemical looping ammonia production (3CLAP) process. Energy Convers. Manag. 2021, 229, 113735. [Google Scholar] [CrossRef]
  53. Lin, Y.-C.; Hsu, K.-H.; Chen, C.-B. Experimental investigation of the performance and emissions of a heavy-duty diesel engine fueled with waste cooking oil biodiesel/ultra-low sulfur diesel blends. Energy 2011, 36, 241–248. [Google Scholar] [CrossRef]
  54. Murillo, S.; Miguez, J.; Porteiro, J.; Granada, E.; Moran, J. Performance and exhaust emissions in the use of biodiesel in outboard diesel engines. Fuel 2007, 86, 1765–1771. [Google Scholar] [CrossRef]
  55. Nabi, M.N.; Rahman, M.M.; Akhter, M.S. Biodiesel from cotton seed oil and its effect on engine performance and exhaust emissions. Appl. Therm. Eng. 2009, 29, 2265–2270. [Google Scholar] [CrossRef]
  56. Gürü, M.; Artukoğlu, B.D.; Keskin, A.; Koca, A. Biodiesel production from waste animal fat and improvement of its characteristics by synthesized nickel and magnesium additive. Energy Convers. Manag. 2009, 50, 498–502. [Google Scholar] [CrossRef]
  57. Hussain, F.; Soudagar, M.E.M.; Afzal, A.; Mujtaba, M.; Fattah, I.; Naik, B.; Mulla, M.H.; Badruddin, I.A.; Khan, T.; Raju, V.D. Enhancement in Combustion, Performance, and Emission Characteristics of a Diesel Engine Fueled with Ce-ZnO Nanoparticle Additive Added to Soybean Biodiesel Blends. Energies 2020, 13, 4578. [Google Scholar] [CrossRef]
  58. Teoh, Y.H.; How, H.G.; Masjuki, H.H.; Nguyen, H.T.; Kalam, M.A.; Alabdulkarem, A. Investigation on particulate emissions and combustion characteristics of a common-rail diesel engine fueled with Moringa oleifera biodiesel-diesel blends. Renew. Energy 2019, 136, 521–534. [Google Scholar] [CrossRef]
  59. Mofijur, M.; Masjuki, H.H.; Kalam, M.A.; Atabani, A.E.; Arbab, M.I.; Cheng, S.F.; Gouk, S.W. Properties and use of Moringa oleifera biodiesel and diesel fuel blends in a multi-cylinder diesel engine. Energy Convers. Manag. 2014, 82, 169–176. [Google Scholar] [CrossRef]
  60. Sateesh, K.A.; Yaliwal, V.S.; Soudagar, M.E.M.; Banapurmath, N.R.; Fayaz, H.; Safaei, M.R.; Elfasakhany, A.; El-Seesy, A.I. Utilization of biodiesel/Al2O3 nanoparticles for combustion behavior enhancement of a diesel engine operated on dual fuel mode. J. Therm. Anal. Calorim. 2021, 1–15. [Google Scholar] [CrossRef]
  61. Mofijur, M.; Masjuki, H.H.; Kalam, M.A.; Atabani, A.E.; Fattah, I.M.R.; Mobarak, H.M. Comparative evaluation of performance and emission characteristics of Moringa oleifera and Palm oil based biodiesel in a diesel engine. Ind. Crop Prod. 2014, 53, 78–84. [Google Scholar] [CrossRef]
  62. Rashed, M.M.; Kalam, M.A.; Masjuki, H.H.; Mofijur, M.; Rasul, M.G.; Zulkifli, N.W.M. Performance and emission characteristics of a diesel engine fueled with palm, jatropha, and moringa oil methyl ester. Ind. Crop Prod. 2016, 79, 70–76. [Google Scholar] [CrossRef]
  63. Soudagar, M.E.M.; Mujtaba, M.; Safaei, M.R.; Afzal, A.; Ahmed, W.; Banapurmath, N.; Hossain, N.; Bashir, S.; Badruddin, I.A.; Goodarzi, M. Effect of Sr@ ZnO nanoparticles and Ricinus communis biodiesel-diesel fuel blends on modified CRDI diesel engine characteristics. Energy 2021, 215, 119094. [Google Scholar] [CrossRef]
  64. Nalgundwar, A.; Paul, B.; Sharma, S.K. Comparison of performance and emissions characteristics of DI CI engine fueled with dual biodiesel blends of palm and jatropha. Fuel 2016, 173, 172–179. [Google Scholar] [CrossRef]
  65. Sarafraz, M.; Hormozi, F.; Peyghambarzadeh, S. Role of nanofluid fouling on thermal performance of a thermosyphon: Are nanofluids reliable working fluid? Appl. Therm. Eng. 2015, 82, 212–224. [Google Scholar] [CrossRef]
  66. Srithar, K.; Balasubramanian, K.A.; Pavendan, V.; Kumar, B.A. Experimental investigations on mixing of two biodiesels blended with diesel as alternative fuel for diesel engines. J. King Saud Univ. Eng. Sci. 2014, 29, 50–56. [Google Scholar] [CrossRef]
  67. Harari, P.; Banapurmath, N.; Yaliwal, V.; Soudagar, M.E.M.; Khan, T.Y.; Mujtaba, M.; Safaei, M.R.; Akram, N.; Goodarzi, M.; Elfasakhany, A. Experimental investigation on compression ignition engine powered with pentanol and thevetia peruviana methyl ester under reactivity controlled compression ignition mode of operation. Case Stud. Therm. Eng. 2021, 25, 100921. [Google Scholar] [CrossRef]
  68. Fayaz, H.; Mujtaba, M.; Soudagar, M.E.M.; Razzaq, L.; Nawaz, S.; Nawaz, M.A.; Farooq, M.; Afzal, A.; Ahmed, W.; Khan, T.Y. Collective effect of ternary nano fuel blends on the diesel engine performance and emissions characteristics. Fuel 2021, 293, 120420. [Google Scholar] [CrossRef]
  69. Mohammed, E.-K.; Nemit-allah, M.A. Experimental investigations of ignition delay period and performance of a diesel engine operated with Jatropha oil biodiesel. Alex. Eng. J. 2013, 52, 141–149. [Google Scholar]
  70. Rahman, M.M.; Hassan, M.H.; Kalam, M.A.; Atabani, A.E.; Memon, L.A.; Rahman, S.M.A. Performance and emission analysis of Jatropha curcas and Moringa oleifera methyl ester fuel blends in a multi-cylinder diesel engine. J. Clean. Prod. 2014, 65, 304–310. [Google Scholar] [CrossRef]
  71. Abedin, M.J.; Masjuki, H.H.; Kalam, M.A.; Sanjid, A.; Rahman, S.M.A.; Fattah, I.M.R. Performance, emissions, and heat losses of palm and jatropha biodiesel blends in a diesel engine. Ind. Crop Prod. 2014, 59, 96–104. [Google Scholar] [CrossRef]
  72. Yadav, A.K.; Khan, M.E.; Dubey, A.M.; Pal, A. Performance and emission characteristics of a transportation diesel engine operated with non-edible vegetable oils biodiesel. Case Stud. Therm. Eng. 2016, 8, 236–244. [Google Scholar] [CrossRef] [Green Version]
  73. Gumus, M.; Kasifoglu, S. Performance and emission evaluation of a compression ignition engine using a biodiesel (apricot seed kernel oil methyl ester) and its blends with diesel fuel. Biomass Bioenergy 2010, 34, 134–139. [Google Scholar] [CrossRef]
  74. Zhang, X.; Gao, G.; Li, L.; Wu, Z.; Hu, Z.; Deng, J. Characteristics of Combustion and Emissions in A DI Engine Fueled with Biodiesel Blends from Soybean Oil; SAE Technical Paper; SAE: Warrendale, PA, USA, 2008; pp. 0148–7191. [Google Scholar]
  75. Mujtaba, M.A.; Kalam, M.A.; Masjuki, H.H.; Razzaq, L.; Khan, H.M.; Soudagar, M.E.M.; Gul, M.; Ahmed, W.; Raju, V.D.; Kumar, R.; et al. Development of empirical correlations for density and viscosity estimation of ternary biodiesel blends. Renew. Energy 2021. [Google Scholar] [CrossRef]
Figure 2. Variation in BP with engine speed for biodiesel blends and diesel.
Figure 2. Variation in BP with engine speed for biodiesel blends and diesel.
Applsci 11 07071 g002
Figure 3. Variation in BSFC with varying engine speed for biodiesel blends and diesel.
Figure 3. Variation in BSFC with varying engine speed for biodiesel blends and diesel.
Applsci 11 07071 g003
Figure 4. Variation in BTE with varying engine speed for biodiesel blends and diesel.
Figure 4. Variation in BTE with varying engine speed for biodiesel blends and diesel.
Applsci 11 07071 g004
Figure 5. Unburned hydrocarbon emissions of Moringa oleifera biodiesel blends and diesel at varying engine speeds.
Figure 5. Unburned hydrocarbon emissions of Moringa oleifera biodiesel blends and diesel at varying engine speeds.
Applsci 11 07071 g005
Figure 6. Carbon dioxide emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Figure 6. Carbon dioxide emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Applsci 11 07071 g006
Figure 7. Carbon monoxide emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Figure 7. Carbon monoxide emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Applsci 11 07071 g007
Figure 8. Nitrogen oxides emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Figure 8. Nitrogen oxides emissions of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Applsci 11 07071 g008
Figure 9. Smoke opacity of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Figure 9. Smoke opacity of Moringa oleifera biodiesel blends and diesel at various engine speeds.
Applsci 11 07071 g009
Table 1. Operating parameters of gas chromatography.
Table 1. Operating parameters of gas chromatography.
ParametersSpecifications
InjectorSplit type
Injection volume1 μL
Oven temperature210 °C, isothermal
Split flow100 mL/min
Carrier gasHelium, 83 kPa
Column60 °C for the initial time
Column 2 flowThe constant flow of helium at 1 L/min
TemperatureRise from 60 °C to 200 °C at an interval of 10 °C
Detector250 °C, flame ionization detector with electronic flow control
Table 2. Composition of moringa oleifera methyl ester.
Table 2. Composition of moringa oleifera methyl ester.
Fatty AcidsChemical StructureMolecular Mass (g/mol)Composition of MOME (w/w %)
LaurateC12:0214.34
MyristateC14:0242.40.1
PalmitateC16:0270.458.1
PalmitoleateC16:1268.431.6
StearateC18:0298.55.4
OleateC18:1296.4974.3
LinoleateC18:2294.474.1
LinolenateC18:3292.460.2
ArachidateC20:0326.562.3
EicosenoateC20:1324.241.5
BehenateC22:0354.612.6
Saturated 18.7
Monounsaturated 76.9
Polyunsaturated 4.4
Table 3. Physicochemical characteristics of Moringa oleifera biodiesel and their blends.
Table 3. Physicochemical characteristics of Moringa oleifera biodiesel and their blends.
PropertyUnitASTM StandardsDieselMOB10MOB20MOB100
Density at 15 °Ckg/m3D4052856.9860.1861.7877.6
Density at 40 °Ckg/m3D4052828.4831.2834.6860.7
Kinematic viscosity at 40 °Cmm2/sD4453.25253.55723.67724.8338
Viscosity index91101.5112.1185.2
Oxidation stabilityhEN ISO 141123533.532.726.4
Cetane indexD473748.948.349.658.5
Flashpoint°CD9368.780.182.9151.2
Cloud point°CD250087818
Pour point°CD9703618
Cold filter plugging point°CD637156618
Calorific valueMJ/kgD24045.8644.1843.6139.98
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Soudagar, M.E.M.; Khan, H.M.; Khan, T.M.Y.; Razzaq, L.; Asif, T.; Mujtaba, M.A.; Hussain, A.; Farooq, M.; Ahmed, W.; Shahapurkar, K.; et al. Experimental Analysis of Engine Performance and Exhaust Pollutant on a Single-Cylinder Diesel Engine Operated Using Moringa Oleifera Biodiesel. Appl. Sci. 2021, 11, 7071. https://doi.org/10.3390/app11157071

AMA Style

Soudagar MEM, Khan HM, Khan TMY, Razzaq L, Asif T, Mujtaba MA, Hussain A, Farooq M, Ahmed W, Shahapurkar K, et al. Experimental Analysis of Engine Performance and Exhaust Pollutant on a Single-Cylinder Diesel Engine Operated Using Moringa Oleifera Biodiesel. Applied Sciences. 2021; 11(15):7071. https://doi.org/10.3390/app11157071

Chicago/Turabian Style

Soudagar, Manzoore Elahi M., Haris Mahmood Khan, T. M. Yunus Khan, Luqman Razzaq, Tahir Asif, M. A. Mujtaba, Abrar Hussain, Muhammad Farooq, Waqar Ahmed, Kiran Shahapurkar, and et al. 2021. "Experimental Analysis of Engine Performance and Exhaust Pollutant on a Single-Cylinder Diesel Engine Operated Using Moringa Oleifera Biodiesel" Applied Sciences 11, no. 15: 7071. https://doi.org/10.3390/app11157071

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