Next Article in Journal
Performance and Energy Trade-Offs for Parallel Applications on Heterogeneous Multi-Processing Systems
Next Article in Special Issue
Effect of Additivized Biodiesel Blends on Diesel Engine Performance, Emission, Tribological Characteristics, and Lubricant Tribology
Previous Article in Journal
Raiffa-Kalai-Smorodinsky Bargaining Solution for Bilateral Contracts in Electricity Markets
Previous Article in Special Issue
Resource Recovery from Waste Coffee Grounds Using Ultrasonic-Assisted Technology for Bioenergy Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research of Energy and Ecological Indicators of a Compression Ignition Engine Fuelled with Diesel, Biodiesel (RME-Based) and Isopropanol Fuel Blends

by
Alfredas Rimkus
1,2,
Jonas Matijošius
2,3,* and
Sai Manoj Rayapureddy
2
1
Department of Automobile Transport Engineering, Technical Faculty, Vilnius College of Technologies and Design, Olandų g. 16, LT-01100 Vilnius, Lithuania
2
Department of Automobile Engineering, Faculty of Transport Engineering, Vilnius Gediminas Technical University, J. Basanavičiaus g. 28, LT-03224 Vilnius, Lithuania
3
Institute of Mechanical Science, Vilnius Gediminas Technical University, Basanavičiaus g. 28, LT-03224 Vilnius, Lithuania
*
Author to whom correspondence should be addressed.
Energies 2020, 13(9), 2398; https://doi.org/10.3390/en13092398
Submission received: 27 March 2020 / Revised: 5 May 2020 / Accepted: 6 May 2020 / Published: 11 May 2020
(This article belongs to the Special Issue Alternative Fuels and Their Application to Combustion Engines)

Abstract

:
This article presents the results of a study of energy and ecological indicators at different engine loads (BMEP) adjusting the Start of Injection (SOI) of a Compression Ignition Engine fuelled with blends of diesel (D), rapeseed methyl ester (RME)-based biodiesel and isopropanol (P). Fuel blends mixed at D50RME45P5, D50RME40P10 and D50RME30P20 proportions were used. Alcohol-based fuels, such as isopropanol, were chosen because they can be made from different biomass-based feedstocks and used as additives with diesel fuel in diesel engines. Diesel fuel and its blend with 10% alcohol have almost the same thermal efficiency (BTE). In further examination of energy and ecological indicators, combustion parameters were analysed at SOI 6 CAD BTDC using AVL BOOST software (BURN subprogram). Increasing alcohol content in fuel blends led to a reduced cetane number, which prolonged the ignition delay phase and intensified heat release in the premixed combustion phase. Higher combustion temperatures and oxygen content in the fuel blends increased NOx emissions. Lower C/H ratios and higher O2 levels affected by RME and isopropanol reduced smoke emissions.

Graphical Abstract

1. Introduction

The rising demand for fossil fuels and the environmental issues concerning their use are the biggest challenges that people face today. The transportation sector alone contributes up to 30% of the world’s harmful emissions [1]. As the demand for fossil fuels continues to rise faster than its supply, fossil fuels deplete, which in turn drives the price of such fuels as diesel and petrol up [2,3,4]. The absence of a replacement for vehicles that run on liquid fuels, and a high price of electric vehicles made the automotive industry devote its resources to finding alternative fuels as a replacement and to decreasing the emissions concerned with environmental problems [5,6,7]. Recent studies revealed that greenhouse gases and harmful combustion chamber emissions can be significantly reduced by using fuels, such as alcohol and biodiesel, as primary alternatives [8,9,10,11,12,13,14,15,16].
In order to beat the fossil fuel deficiency and control the increasing demand of natural gas, opportunities for using alternative fuels in internal combustion engines have been searched [17]. The ease of handling and storing alcohol and biodiesel makes them promising substitutes for fossil fuels [9,18,19,20,21].
Modern biofuel production technologies are sufficiently developed and focused on the production of alcohols, which are very widely used as fuel additives [22]. Mostly ethanol is produced, the production technologies of which have been well known since ancient times. As a fuel additive ethanol has many disadvantages, the most important of which is its corrosivity [23]. Therefore, the selection of an alternative fuel blend for these studies focused on propanol, which is less corrosive and is close to petroleum diesel in its properties and calorific value [24]. In addition, when blended with diesel (10% v/v), propanol performs better in terms of emissions and noise than ethanol [25]. This choice was also based on its cheapness compared to other higher order alcohols like butanol and pentanol [26].
The use of alternative fuels in standard internal combustion engines most often requires modifying such engines. Our proposed three-component fuel mixture allows the replacement of conventional fuels (diesel) with alternative analogues, the use of biodiesel must take into account many operational aspects, such as e.g., CFPP, and alcohol propanol allows one to partially solve this problem. Therefore, the chosen three-component diesel-biodiesel-propanol blend allows us to solve the difficult task of using alternative fuels without engine modifications. The three-component mixtures mentioned are rarely found in the literature and are usually defined by a narrow analysis of fuel consumption and some exhaust parameters, and a more detailed analysis of heat release and other parameters is postponed for further research [27].
The performance of the diesel engine with the blend of n-propanol at different proportions like 10% by volume is attainable [28]. Diesel engines can use propanol-diesel blends containing 10 to 20% propanol without significantly affecting engine performance. It is further concluded that 10–20% of propanol-diesel blends are beneficial in reducing harmful fumes and NOx emissions [29]. n-Propanol/diesel blends higher than 30% showed lower soot density due to the predominant effect of increasing spontaneous oxygen content in n-propanol/diesel mixtures [30].
It was observed that when the isopropanol concentration exceeds 15%, the combustion temperature and BTE performance starts to increase in biodiesel [31], especially in biodiesel-diesel-isopropanol blends (80%/10%/10%) [27]. Isopropanol improves the cold-flow limit in blends with both diesel and biodiesel, cooling it when operating at low temperatures [32]. 15% Isopropanol content in diesel allows for improved engine performance, lower smoke and NOx emissions (at low to medium loads), but increases brake specific fuel consumption (BSFC) due to its lower calorific value [33]. Increasing the isopropanol content to 55% compared to diesel led to an increase in nitrogen oxide emissions (by an average of 139%), a reduction in carbon monoxide emissions (45%) and an increase in CO2 emissions (by an average of 17%). However, no significant change in unburnt hydrocarbon emissions was observed [34].
Mixing different alcohols (ethanol-isopropanol and butanol) (EPB) with diesel at the ratio of 20 and 40% did not demonstrate a significant increase in heat release and combustion pressure compared to that of diesel. They have been found to have a lower molecular weight flux permeability than diesel, and the flame light index was lower with the use of the aforementioned additives in diesel, which contributed to lower soot emissions [35]. On the other hand, this led to a shorter initial combustion duration (ICD) and major combustion duration (MCD) conditions [11]. Changing injection strategies for EPB blends revealed that pilot injection reduces the heat release rate and the peak pressure, while dual injection improves fuel economy, reduces NOx emissions, at the same time increasing soot [35].
The blend of 30% EPB alcohols with gasoline improves BTE and slightly increases CO (4.2%), hydrocarbon (HC) (18.9%) and NOx emissions (5.5%) compared to mineral gasoline [36]. Correspondingly, adding 1% of water to EPB blends (10% EPB and 90% mineral gasoline) resulted in an even better ecological effect—a decrease in CO of up to 7.5% and in NOx of up to 12.4%, respectively [12]. The emulsion blend of fuel and water reduces local areas where the maximum temperature is reached, resulting in decreased nitrogen activity and NOx concentration [37]. However, using isopropanol additive in gasoline alone significantly increases HC emissions at low inlet air temperatures with increasing isopropanol concentration in isopropanol-gasoline blends [38].
When using 25% isopropanol in petrol blends with combustion control, there is a direct relationship between octane number and combustion parameters of isopropanol. Start of combustion (SOC) is delayed by increasing isopropanol content in the blend because isopropanol is more resistant to jerky engine operation [39]. By reducing spark time in a petrol engine, isopropanol-gasoline blends (with isopropanol content up to 30%) had lower NOx emissions than those found at the initial spark time [40].
Oxygen content is the key parameter that differs in fossil fuels and biodiesel [41,42,43]. Environmentally friendly biodiesel produces clean and renewable energy, thus making it the alternative hope [44,45,46]. A study found that rapeseed methyl ester (RME) produced lower CO2 emissions due to a lower carbon-to-hydrogen ratio as compared to diesel [47]. Methanol, ethanol are most commonly used in fuel blends with biodiesel, and currently the second ACB wave will make butanol cheaper [48]. The use of proponol in fuel blends with biodiesel is a rare case, determined by the greater development of other alcohol production technologies [20,49]. Lower heating values for B90Pr10 (90% biodiesel and 10% propanol fuel mixtures) and cetane number increased BSFC and fuel gas temperatures while reducing BTE [50]. Table 1 lists the properties of pure diesel, RME and isopropanol [51].
The aim of the research is to reveal the performance, combustion and emission characteristics of IC engines using pure diesel and fuel blends with different proportions of diesel, rapeseed methyl ester-based biodiesel and isopropanol.

2. Materials and Methods

2.1. Engine Testing Equipment

The engine used for testing is a 1.9 Turbocharged Direct Injection (TDI) diesel engine with a VP37 (BOSCH, Stuttgart, Germany) electronic controlled distribution type fuel pump. The start of the injection (SOI) was controlled by the Engine Electronic Control Unit (ECU) and it was a single injection strategy. Figure 1 presents a detailed image of the tested engine and its parts, while research conducted by other authors [52,53,54,55] and Table 2 lists the test engine parameters.
The brake torque MB (Nm) was determined on a load bench with a measurement error of ± 1.2 Nm. Hourly fuel consumption Bf (kg/h) was found using electronic scales SK-5000 with a 0.5% measurement error. The pressure in the cylinder was measured using a piezoelectric sensor GG2-1569 mounted on a glow plug with a sensitivity of 15.8 ± 0.09 pC/bar. Cylinder pressure values were recorded using the LabView Real software at an interval of 0.176 CAD. The pressure in the engine intake manifold was measured using an OHM HD 2304.0 pressure gauge (Delta, Padova, Italy) with a measurement error of ± 0.0002 MPa. The intake air and the exhaust gas temperature were measured using K-type thermocouples IR 8839 accurate to ± 1.5 °C. The exhaust gas concentration was determined using a DiCom 4000 gas analyzer (AVL, Graz, Austria). CO2 measurement accuracy was 0.1% vol., CO—0.01% vol., HC—1 ppm, NOx—1 ppm, and smoke absorption coefficient—0.01 m−1.
Statistical calculations of type A uncertainties were used to measure exhaust. Type A uncertainties were used to determine the standard deviation for repeated measurements, where u(x) is uncertainty, n—repeatability of measurements, and s( x ¯ )—reliability [55,56]:
u ( x ) = s ( x ¯ ) = s ( x ) n
where x ¯ is the mean repeated value; s(x) is a standard deviation; s( x ¯ ) is a standard deviation of the mean.
Uncertainty ranges u(x) of exhaust components are presented in Table 3.

2.2. Fuels and Test Conditions

Tests were conducted using 100% pure diesel fuel and fuel blends prepared using different proportions of diesel (D), rapeseed methyl ester (RME)-based biodiesel and isopropanol (P). The first blend contained 50% diesel, 30% rapeseed methyl ester and 20% isopropanol (D50RME30P20), the second blend had 50% diesel, 40% rapeseed methyl ester and 10% isopropanol (D50RME40P10), and the third one 50% diesel, 45% rapeseed methyl ester and 5% isopropanol (D50RME45P5). The properties like density, mass fraction and lower heating value of the blends were calculated using the following formula:
properties of fuel blends = ∑ [(percentage of fuel blend × property)]
Table 3 presents a comparison of the calculated properties of the fuel blends with the standard diesel fuel. Uncertainties were calculated according to the model B [55]. Standard uncertainties were calculated according to the formula:
u c = u c 2 ( f )
where uc2(f) is the total uncertainty dispersion.
The uncertainty calculations for each fuel blend are presented in Table 4.
Engine tests were carried out at the engine speed of n = 2000 rpm and engine brake torque MB was 30, 60 and 90 Nm, which corresponds to the Brake Mean Effective Pressure (BMEP) 0.2 MPa, 0.4 MPa and 0.6 MPa in the first experimental tests step. These are the loads of a city car running of the ≈ 50 km/h, ≈ 80 km/h and ≈ 100 km/h speeds. During load-changing tests, fuel Start or Injection Timing (SOI ≈ 2 CAD BTDC) was controlled by the engine electronic control unit. There were (BMEP = 0.3 MPa) injection timing was adjusted (SOI = 0…16 CAD BTDC) by modulating the SOI control signal in the second experimental test step. Injection timing was adjusted to determine the variation of engine performance using fuel mixtures of different properties under different combustion conditions. The Energy Indicators (Hourly Fuel Consumption Bf, Brake Specific Fuel Consumption BSFC, Brake Thermal Efficiency (BTE)) and Ecological Indicators (emission of carbon monoxide CO, carbon dioxide CO2, nitrogen oxides NOx, hydrocarbons CH, and smoke) were measured and calculated at different engine loads and by adjusting start of injection. The results of all the tested fuel blends will be analysed comparing them with results of diesel fuel. Experimental tests were performed to ensure repeatability of the experiments. Several experiment design parameters were singled out, such as fuel type used, engine rotation speed, engine load torque, and fuel injection angle.
The density difference between the fuels was found to decrease with increasing alcohol concentration as seen in Table 4. Since the molecular mass of alcohols is lower than that of diesel and biodiesel [57,58,59,60], with the alcohol content increasing from 5% to 10%, the density was found to decrease by 1.45%, and with an additional increase from 10% to 20%, the density decreased by nearly 0.17%. When compared to diesel, fuel blends with a 5 and 10% alcohol content tended to be denser, and the D50RME30P20 blend tended to have a lower density compared to conventional fuel as calculated and presented in Table 4.
A lower heat value was found to differ less with an increase in the alcohol percentage share as seen in Table 4. The lower heat value difference increased to 6.5% when increasing the fuel concentration from 5% to 10%, and a further change of the concentration from 10% to 20% led to the difference increasing by nearly 7.7%. The lower heat value highly depends on carbon and hydrogen content in fuel [61,62,63]. So, with a relatively higher carbon and hydrogen content in diesel compared to their content in other fuel blends, the lower heat value of conventional fuel was found to be higher than that of other fuel blends, which declined with increasing alcohol content as seen in Table 4.
The difference in the cetane number decreased with increasing alcohol content as seen in Table 4. The conventional diesel fuel is known for being rich in paraffin, which helps it achieve a higher cetane number compared to other fuel blends [64]. When increasing the alcohol content from 5% to 10%, the difference was found to increase to 37.5%, and a further increase in alcohol content to 20% led to an increase in the difference of nearly 45%. As seen in Table 4, the cetane number steadily decreased with alcohol content compared to that of the conventional fuel. The fuel stability was ensured by producing fuel blends right before testing and feeding them to the engine.

2.3. Tools for Numerical Analysis of the Combustion Process

Due to a significant change in the cetane number in diesel and other prepared fuel blends, analysing changes in performance of the engine by calculating combustion characteristics and comparing them with those of diesel is necessary. AVL BOOST software was used in calculating the combustion characteristics with the help of BURN subprogram. BURN analysis was conducted having created a digital model of the 1.9 TDI diesel engine used in the experiment as seen in Figure 2. The digital model of the engine was constructed by selecting the required elements from a displayed element catalogue in AVL BOOST. The analysis of combustion requires general data on engine parameters, fuel and data describing the operating point.
Engine parameters and experiment results were uploaded in the BURN subprogram and run to get the combustion characteristics. Changes in the in-cylinder pressure and temperature, heat release rate (ROHR) and the mass fraction burnt (MFB) were calculated in this research:
R O H R = d x d α = 6.908 α C D · ( m v + 1 ) · ( α α S O C α C D ) m v · e 6.908 · ( α α S O C α C D ) ( m v + 1 )
d x = d Q Q
M F B = α S O C α dQ d α · Q ( α ) · d α = 1 e 6.908 · ( α α S O C α C D ) ( m v + 1 ) ,   α > α S O C
where Q —total fuel heat input; α —crank angle; m v —combustion shape parameter; α S O C —start of combustion; α C D —combustion duration.

3. Results and Discussion

The experimental tests were carried out in two steps: (a) changing the engine load BMEP (0.2; 0.4 and 0.6 MPa); (b) changing SOI (0…16 CAD BTDC), and BMEP = 0.3 MPa. After plotting the graphs from the (b) experiment results, a polynomial curve was drawn with a degree of 2 for all the energy and ecological parameters to get a change trend.

3.1. Energy Indicators

Brake Specific Fuel consumption (BSFC) of diesel fuel was low at all loads compared to other fuel blends, as observed in Figure 3a. When increasing alcohol content, fuel consumption tended to increase with D50RME30P20 being at the maximum. Having replaced 50% of diesel by a blend of biodiesel and propanol and increased the concentration of propanol up to 20% (D50RME30P20), BSFC increased ~9% due to a 7.7% reduction in LHV (Table 4) and a change in combustion process. With BMEP = 0.3 MPa, the analysis of BSFC from the perspective of injection timing revealed a higher consumption of all the fuel blends (7–10%), and with advancing angle, the value tended to decrease and further increase after 8–12 CAD BTDC as seen in Figure 3b.
At 8 CAD BTDC, D100 fuel was at its lowest on average. The lowest consumption of D-RME-P blends was achieved with injection timing of 10...12 CAD BTDC, as the cetane number of this fuel decreased to 23.06.
BTE of D50RME45P5 and D50RME40P10 fuel blends at BMEP 0.2 and 0.4 MPa was close to that of diesel, and BTE of D50RME30P20 was 3.5% (at BMEP 0.2 MPa) to 1.8% (at BMEP 0.4 MPa) lower as seen in Figure 3a. With a BMEP = 0.6 MPa, BTE of all fuels was slightly different and reached up to 0.35%.
Observing the dependence curve of efficiency on injection timing in Figure 3b, at BMEP = 0.3 MPa, BTE of the D50RME30P20 fuel blend was ~1.8% lower than that of diesel. BTE of fuel blends with 5% and 10% isopropanol content was lower (2.5…1.2%) at SOI = 0…6 CAD BTDC, but at SOI = 8…12 CAD BTDC, BTE of D50RME45P5 and D50RME40P10 was the same as BTE of D100.

3.2. Ecological Indicators

Carbon dioxide comparative emissions (g/kWh) were found to decrease for all the fuels with increasing load as seen in Figure 4a as BTE increased and BSFC decreased. At a low load (BMEP = 0.2 MPa), CO2 emissions of the blends containing isopropanol were 0.8…1.9% higher compared to diesel but increasing the load to BMEP = 0.6 MPa resulted in ~0.8% lower CO2 emissions for the D50RME40P10 fuel blend.
This was due to a 2.1% decrease in the C/H ratio (Table 4) and the fact that at higher loads the BTE of all fuels was similar. Checking the dependence of emissions on injection timing revealed that all the fuels showed a gradual decrease in emissions with advancing injection timing as seen in Figure 4b, BTE increased, and smoke emissions decreased (Figure 5b). Diesel was found to have similar CO2 emissions compared to emissions of other fuel blends. CO2 emissions of the D50RME40P10 fuel blend at various SOIs were only ~0.2% lower than emissions of D100. Even though RME and C/H ratio of isopropanol is lower, increased fuel consumption increases CO2 emissions.
Hydrocarbon emissions of all the fuels were found to decrease with increasing load as seen in Figure 4a, as the combustion temperature increased [65,66]. Having the lowest emissions as compared to all the other fuels, diesel also tended to have a steady decrease pattern of ~38%, ~45% and ~60% (compared to the D50RME45P5, D50RME40P10 and D50RME30P20 fuel blends in hydrocarbon emissions). When increasing the alcohol content, fuel blends tended to have higher hydrocarbon emissions due to increasing alcohol base of the fuel, however, HC emissions were low compared to petrol engine [52]. The observation of the dependence of hydrocarbon emissions on injection timing revealed that with an increase in alcohol content, emissions tended to increase as seen in Figure 4b. HC emissions of D50RME45P5, D50RME40P10 and D50RME30P20 increased by ~15%, ~28% and ~58% compared to D100 at SOI ≈ 6 CAD BTDC. With the engine running on all the fuels, hydrocarbon emissions showed a slight growth trend when injection timing was advanced more than 6 CAD BTDC.
Nitrogen oxide emissions for D50RME45P5, D50RME40P10, D50RME30P20 at a low load (BMEP = 0.2 MPa) were ~10%, ~12%, and ~21% higher compared to those of D100 fuel as seen in Figure 5a. This was mainly due to the increased ignition delay due to a low cetane number of isopropanol (see the Dependence of the rate of heat release and the mass burnt fraction on the crank angle degree figure bellow.) and the increased oxygen concentration in the fuel blend (Table 4).
With an increase in the load (BMEP = 0.6 MPa), the difference in NOx emissions was reduced to ~4%, ~7% and ~10% as the effect of ignition delay on different fuels was reduced. While analysing the dependence of nitrogen oxide emissions on the injection timing, emissions were found to rise steadily with increasing injection timing as seen in Figure 5b, because combustion occurred at a lower volume and higher temperatures [67]. All the fuels were found to follow a similar pattern, but after a more careful analysis, diesel emissions were found to be lower (1…4%) compared to emissions of other fuel blends. An earlier injection timing reduced the difference between NOx emissions of diesel and fuel blends [68,69,70].
The smoke level of fuels tended to increase with increasing load as seen in Figure 5a, as fuel mass increases per cycle and the air-to-fuel ratio decreases. Increasing the isopropanol concentration in fuel blends reduces smoke level, and this effect is more intense with an increasing engine load [71]. At a low load (BMEP = 0.2 MPa), the D50RME45P5, D50RME40P10 and D50RME30P20 fuel smoke emissions decreased by ~6%, ~10% and ~14%, respectively, in comparison to pure diesel. Higher oxygen concentrations and lower C/H ratios resulted in lower D50RME30P20 smoke emissions. Increasing the load to BMEP = 0.6 MPa increased the smoke reduction effect to ~12%, ~16%.
The analysis of the dependence of smoke levels on injection timing revealed that the levels tended to decrease with advancing the injection timing as seen in Figure 5b. Interestingly, throughout the SOI study range (0...16 CAD BTDC), the combustion performance resulted in the largest reduction in diesel smoke emissions from ~7.25 m−1 to ~4.25 m−1 (~40%), while the D50RME30P20 smoke emissions decreased from ~5.7 to ~5 m−1 (~12%). Therefore, at SOI = 0...8 CAD BTDC (low advanced injection timing), D50RME30P20 had the lowest smoke emissions, and at SOI = 10...16 CAD BTDC (high advanced injection timing), smoke emissions of D100 fuel were the lowest.

3.3. Combustion Characteristics

The analysis of combustion characteristics was conducted with the engine operating at BMEP = 0.3 MPa (n = 2000 rpm). Changing SOI (0…16 CAD BTDC) allowed comparing energy and ecological performance of the engine running on different fuels (Figure 3b, Figure 4b and Figure 5b) and concluding that with ignition timing being 6 CAD BTDC engine efficiency is close to the maximum, and smoke and NOx emissions are relatively low. The pressure values of all the fuels at SOI = 6 CAD BTDC obtained during the experiment as seen in Figure 6 were uploaded in the AVL BOOST (BURN subprogram) to get the combustion characteristics. Since the result at SOI = 6 CAD BTDC was relatively good, the combustion characteristics of fuel blends, such as in-cylinder pressure, pressure rise, rate of heat release (ROHR) and mass fraction burned (MFB), were analysed and compared to those of pure diesel at that particular degree of ignition timing.
Conventional diesel fuel (D100) showed the maximum pressure of ~7.32 MPa at 367 CAD, while the maximum pressures of D50RME45P5, D50RME40P10 and D50RME30P20 were ~1.0%, ~1.1% and ~1.7% higher as seen in Figure 6. There was also a delayed burning with isopropanol.
A combustion-driven pressure starts to increase the earliest with the engine running on diesel. The maximum pressure rise of ~0.40 MPa/CAD was observed at 361 CAD as seen in Figure 6. The maximum pressure rise increases with an increase in alcohol percentage. ~7.5% (361.5 CAD), ~17% (362 CAD) and ~29% (363 CAD) was the maximum increase for D50RME45P5, D50RME40P10 and D50RME30P20.
The maximum rate of heat release of diesel fuel was lower at 32.0 J/deg than that of other fuel blends as seen in Figure 7. The maximum rate of 34.8 J/deg (~8% higher) was observed with the D50RME4P5 fuel blend, while the maximum rate of D50RME40P10 was 38.1 J/deg (~19% higher) and that of D50RME30P20—46.4 J/deg (~44% higher). Isopropanol reduces the cetane number (Table 4), prolongs the ignition delay phase and significantly increases the maximum ROHR during the premixed combustion phase [58]. The LHV of isopropanol is lower, but this is offset by the higher fuel content (Figure 7), and the diffusion combustion phase produces a similar amount of heat for all fuels. Higher maximum temperatures during premixed combustion phase increase the formation of nitrogen oxides, but allows for a better combustion of soot at the end of the combustion process [69,70,71].
The mass burn fraction diagram in Figure 7 confirms the prolongation of the ignition delay phase in fuel blends with a higher alcohol content. Ignition delay for D100 was ~ 5 CAD, D50RME45P5—~6 CAD, D50RME40P10—~7 CAD and D50RME30P20—~8 CAD.
Although ignition delay is longer for blends with isopropanol [72], oxygen concentration in blends significantly increases due to RME and isopropanol. This significantly accelerates the combustion process during the premixed combustion phase, and 0.5 MBF of all fuels is available at ~7.5 CAD ATDC. The diffusion combustion phase MBF intensity is similar for all fuels, although the fuel consumption of D50RME30P20 increased by ~9% (Figure 3), which was offset by increased injection rate due to a lower isopropanol viscosity and faster combustion driven by a higher oxygen concentration. 99% of the D100 fuel mass ends up burning ~410 CAD ATDC, D50RME45P5—~412 CAD ATDC, D50RME40P10—~413 CAD ATDC and D50RME30P20—~414 CAD ATDC.

4. Conclusions

The analysis of the energy, ecological and combustion parameters of diesel 100, D50RME45P5, D50RME40P10 and D50RME30P20 in a turbocharged direct injection diesel engine at the speed (n) of 2000 rpm and under various loads and injection timings allows making the following conclusions:
(1)
RME and isopropanol reduce LHV and the cetane number of fuel blends, but increase the oxygen concentration in the blend and lower the C/H ratio.
(2)
D50RME30P20 brake specific fuel consumption increased by ~9% compared to D100 and BTE decreased by ~1.8% due to a 7.7% reduction in LHV and a change in the combustion process. The maximum BTE of the D50RME40P10 fuel blend was equal to D100 efficiency having advanced the injection timing of the fuel blend ~2 CAD. This offset the increase in the ignition delay due to a low propanol cetane number (~12).
(3)
Carbon dioxide emissions of all fuels are similar, but the best carbon dioxide effect was obtained with the D50RME40P10 fuel blend. At medium loads, CO2 emissions of this blend declined by ~0.2% compared to diesel, though fuel consumption increased by ~6%, as the C/H ratio of the fuel blend was 2.1% lower. A more advanced injection timing (SOI = 8…10 CAD BTDC) at the minimum fuel consumption allows achieving lower CO2 emissions.
(4)
Isopropanol has a greater impact on nitrogen oxide emissions at low loads. NOx emissions of D50RME45P5, D50RME40P10 and D50RME30P20 increased by ~10%, ~12% and ~21% due to a higher oxygen concentration in the blends and a higher combustion temperature. With increasing load, an increase in NOx emissions was lower (~4%, ~7% and ~10%) as a low isopropanol cetane number had a lesser effect on the ignition delay phase and the heat release rate during the premixed combustion phase. With an early injection timing, NOx emissions increased, but the impact of alcohol was lower.
(5)
Having replaced diesel with fuel blends at a low load resulted in a ~6%, ~10% and ~14% reduction in smoke and an average load reduction of ~12%, ~16% and ~28%. Smoke was reduced by lower C/H ratios and increased oxygen content in the fuels. As the load increased, the BTE of the fuel blends increased more intensively, which further reduced smoke emissions. In the case of early injection timing (SOI = 8…16 CAD BTDC), smoke emissions of the fuel blends changed (decreased) less intensively due to changed fuel characteristics compared to pure diesel.
(6)
At a low engine load (BMEP = 0.3 MPa), the average rotation speed (n = 2000 rpm), the fixed injection timing (SOI = 6 CAD BTDC) and the replacement of diesel by fuel blends with a higher alcohol content (5%, 10% and 20%) resulted in ignition delay changing from ~5CAD to ~6 CAD, ~7 CAD and ~8 CAD. A greater ignition delay (accumulates more fuel) and a higher oxygen content in fuel during the premixed combustion phase increased the heat release intensity by ~8%, ~19% and ~44%, which in turn increased the pressure rise by ~8%, ~17% and ~29% in the thermodynamic load of the crank mechanism. During the diffusion combustion phase, the combustion heat release of all the fuels examined was similar.
(7)
The authors plan to continue research of these three-component blends by increasing the share of alternative fuels in the blends and assessing the impact of the EGR system when using these blends.

Author Contributions

Author Contributions: Conceptualization, A.R. and S.M.R.; methodology, A.R., S.M.R. and J.M.; software, S.M.R. and A.R.; formal analysis, A.R. and J.M.; validation, A.R. and S.M.R.; writing—original draft preparation, A.R. and J.M.; writing—review and editing, A.R. and J.M.; supervision, A.R., and J.M.; project administration, J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Part of the combustion characteristics reported in this article were obtained using the engine numerical analysis tool AVL BOOST, acquired by signing a Cooperation Agreement between AVL Advanced Simulation Technologies and the faculty of the Transport Engineering of Vilnius Gediminas Technical University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations and Nomenclature

ACBAcetoneb—utanol producing process
ATDCAfter Top Dead Centre (CAD)
AVLAnstalt für Verbrennungskraftmaschinen List
BDCBottom Dead Center
BfFuel mass consumption (kg/h)
BMEPBrake Mean Effective Pressure (MPa)
BSFCBrake Specific Fuel Consumption (g/kWh)
BTDCBefore Top Dead Center (CAD)
BTEBrake Thermal Efficiency
CACrank Angle (degree)
CFPPCold filter plugging point
COCarbon monoxide
CO2Carbon dioxide
CNCetane Number
CVCalorific Value
DDiesel fuel
ECUElectronic Control Unit
EPBEthanol-Propanol and butanol fuel blend
HCHydrocarbons
ICInternal combustion
ICDInitial combustion duration (CAD)
LHVLower Heating Value (MJ/kg)
MBBrake torque (Nm)
MFBMass fraction burned
MCDMajor combustion duration (CAD)
nRotational speed of the crankshaft (rpm)
NOxNitrogen Oxide
O2Oxygen
PIsopropanol
ROHRRate of heat release (J/deg)
RMERapeseed Methyl Ester
SOIStart of Injection (CAD)
TDCTop Dead Center
TDITurbocharged Direct Injection

References

  1. Enriquez, A.; Benoit, L.; Dalkmann, H.; Brannigan, C. GIZ Sourcebook 5e Transport and Climate Change. Tech. Rep. 2014. [Google Scholar] [CrossRef]
  2. Abas, N.; Kalair, A.; Khan, N. Review of fossil fuels and future energy technologies. Futures 2015, 69, 31–49. [Google Scholar] [CrossRef]
  3. David, M. The role of organized publics in articulating the exnovation of fossil-fuel technologies for intra- and intergenerational energy justice in energy transitions. Appl. Energy 2018, 228, 339–350. [Google Scholar] [CrossRef]
  4. Jia, T.; Dai, Y.; Wang, R. Refining energy sources in winemaking industry by using solar energy as alternatives for fossil fuels: A review and perspective. Renew. Sustain. Energy Rev. 2018, 88, 278–296. [Google Scholar] [CrossRef]
  5. Tamilselvan, P.; Nallusamy, N.; Rajkumar, S. A comprehensive review on performance, combustion and emission characteristics of biodiesel fuelled diesel engines. Renew. Sustain. Energy Rev. 2017, 79, 1134–1159. [Google Scholar] [CrossRef]
  6. Sładkowski, A. (Ed.) Ecology in Transport: Problems and Solutions; Lecture Notes in Networks and Systems; Springer International Publishing: Cham, Switzerland, 2020; Volume 124, ISBN 978-3-030-42322-3. [Google Scholar]
  7. Bureika, G.; Steišūnas, S. Complex Evaluation of Electric Rail Transport Implementation in Vilnius City. Transp. Probl. 2016, 11, 49–60. [Google Scholar] [CrossRef] [Green Version]
  8. Bhasker, J.P.; Porpatham, E. Effects of compression ratio and hydrogen addition on lean combustion characteristics and emission formation in a Compressed Natural Gas fuelled spark ignition engine. Fuel 2017, 208, 260–270. [Google Scholar] [CrossRef]
  9. Chen, H.; He, J.; Zhong, X. Engine combustion and emission fuelled with natural gas: A review. J. Energy Inst. 2019, 92, 1123–1136. [Google Scholar] [CrossRef]
  10. Ghadikolaei, M.A.; Cheung, C.S.; Yung, K.-F. Study of combustion, performance and emissions of diesel engine fueled with diesel/biodiesel/alcohol blends having the same oxygen concentration. Energy 2018, 157, 258–269. [Google Scholar] [CrossRef]
  11. Li, Y.; Chen, Y.; Wu, G.; Lee, C.-F.F.; Liu, J. Experimental comparison of acetone-n-butanol-ethanol (ABE) and isopropanol-n-butanol-ethanol (IBE) as fuel candidate in spark-ignition engine. Appl. Therm. Eng. 2018, 133, 179–187. [Google Scholar] [CrossRef]
  12. Li, Y.; Chen, Y.; Wu, G.; Liu, J. Experimental evaluation of water-containing isopropanol-n-butanol-ethanol and gasoline blend as a fuel candidate in spark-ignition engine. Appl. Energy 2018, 219, 42–52. [Google Scholar] [CrossRef]
  13. Li, G.; Liu, Z.; Lee, T.H.; Lee, C.-F.F.; Zhang, C. Effects of dilute gas on combustion and emission characteristics of a common-rail diesel engine fueled with isopropanol-butanol-ethanol and diesel blends. Energy Convers. Manag. 2018, 165, 373–381. [Google Scholar] [CrossRef]
  14. Li, M.; Xu, J.; Xie, H.; Wang, Y. Transport biofuels technological paradigm based conversion approaches towards a bio-electric energy framework. Energy Convers. Manag. 2018, 172, 554–566. [Google Scholar] [CrossRef]
  15. Moula, M.E.; Nyári, J.; Bartel, A. Public acceptance of biofuels in the transport sector in Finland. Int. J. Sustain. Built Environ. 2017, 6, 434–441. [Google Scholar] [CrossRef]
  16. Yan, X.; Gao, P.; Zhao, G.; Shi, L.; Xu, J.-B.; Zhao, T. Transport of highly concentrated fuel in direct methanol fuel cells. Appl. Therm. Eng. 2017, 126, 290–295. [Google Scholar] [CrossRef]
  17. Zefreh, M.M.; Török, Á. Theoretical Comparison of the Effects of Different Traffic Conditions on Urban Road Traffic Noise. J. Adv. Transp. 2018, 2018, 7949574. [Google Scholar] [CrossRef] [Green Version]
  18. Hoseinpour, M.; Sadrnia, H.; Tabasizadeh, M.; Ghobadian, B. Evaluation of the effect of gasoline fumigation on performance and emission characteristics of a diesel engine fueled with B20 using an experimental investigation and TOPSIS method. Fuel 2018, 223, 277–285. [Google Scholar] [CrossRef]
  19. Jamuwa, D.; Sharma, D.; Soni, S. Experimental investigation of performance, exhaust emission and combustion parameters of stationary compression ignition engine using ethanol fumigation in dual fuel mode. Energy Convers. Manag. 2016, 115, 221–231. [Google Scholar] [CrossRef]
  20. Yusri, I.; Mamat, R.; Najafi, G.; Razman, A.; Awad, O.; Azmi, W.H.; Ishak, W.; Shaiful, A. Alcohol based automotive fuels from first four alcohol family in compression and spark ignition engine: A review on engine performance and exhaust emissions. Renew. Sustain. Energy Rev. 2017, 77, 169–181. [Google Scholar] [CrossRef] [Green Version]
  21. Zaharin, M.; Abdullah, N.; Najafi, G.; Sharudin, H.; Yusaf, T. Effects of physicochemical properties of biodiesel fuel blends with alcohol on diesel engine performance and exhaust emissions: A review. Renew. Sustain. Energy Rev. 2017, 79, 475–493. [Google Scholar] [CrossRef]
  22. Szabados, G.; Bereczky, A. Experimental investigation of physicochemical properties of diesel, biodiesel and TBK-biodiesel fuels and combustion and emission analysis in CI internal combustion engine. Renew. Energy 2018, 121, 568–578. [Google Scholar] [CrossRef]
  23. Žaglinskis, J.; Lukács, K.; Bereczky, A. Comparison of properties of a compression ignition engine operating on diesel–biodiesel blend with methanol additive. Fuel 2016, 170, 245–253. [Google Scholar] [CrossRef]
  24. Matijošius, J.; Sokolovskij, E. Research into the Quality of fuels and their Biocomponents. Transport 2009, 24, 212–217. [Google Scholar] [CrossRef] [Green Version]
  25. Pinzi, S.; Redel-Macías, M.; Candia, D.E.L.; Soriano, J.A.; Dorado, M. Influence of ethanol/diesel fuel and propanol/diesel fuel blends over exhaust and noise emissions. Energy Procedia 2017, 142, 849–854. [Google Scholar] [CrossRef]
  26. Bereczky, A. The Past, Present and Future of the Training of Internal Combustion Engines at the Department of Energy Engineering of BME. In Vehicle and Automotive Engineering; Jármai, K., Bolló, B., Eds.; Springer Science and Business Media LLC: Cham, Switzerland, 2017; Volume 13, pp. 225–234. [Google Scholar]
  27. Bencheikh, K.; Atabani, A.; Shobana, S.; Mohammed, M.; Uğuz, G.; Arpa, O.; Kumar, G.; Ayanoğlu, A.; Bokhari, A. Fuels properties, characterizations and engine and emission performance analyses of ternary waste cooking oil biodiesel–diesel–propanol blends. Sustain. Energy Technol. Assess. 2019, 35, 321–334. [Google Scholar] [CrossRef]
  28. Yogesh, P.; Dinesh Sakthi, S.; Aravinth, C.; Sathiyakeerthy, K.; Susenther, M. Performance Test on Diesel-Biodiesel–Propanol Blended Fuels in CI Engine. IJISRT 2018, 3, 1–6. [Google Scholar]
  29. Muthaiyan, P.; Gomathinayagam, S. Combustion Characteristics of a Diesel Engine Using Propanol Diesel Fuel Blends. J. Inst. Eng. (India) Ser. C 2016, 97, 323–329. [Google Scholar] [CrossRef]
  30. Thillainayagam, M.; Venkatesan, K.; Dipak, R.; Subramani, S.; Sethuramasamyraja, B.; Babu, R.K. Diesel reformulation using bio-derived propanol to control toxic emissions from a light-duty agricultural diesel engine. Environ. Sci. Pollut. Res. 2017, 24, 16725–16734. [Google Scholar] [CrossRef]
  31. Campos-Fernández, J.; Arnal, J.M.; Gomez, J.; LaCalle, N.; Dorado, M.P. Performance tests of a diesel engine fueled with pentanol/diesel fuel blends. Fuel 2013, 107, 866–872. [Google Scholar] [CrossRef]
  32. Atmanli, A.; Atmanlı, A. Comparative analyses of diesel–waste oil biodiesel and propanol, n-butanol or 1-pentanol blends in a diesel engine. Fuel 2016, 176, 209–215. [Google Scholar] [CrossRef]
  33. Şen, M. The effect of the injection pressure on single cylinder diesel engine fueled with propanol–diesel blend. Fuel 2019, 254, 115617. [Google Scholar] [CrossRef]
  34. Jamrozik, A.; Tutak, W.; Pyrc, M.; Gruca, M.; Kocisko, M. Study on co-combustion of diesel fuel with oxygenated alcohols in a compression ignition dual-fuel engine. Fuel 2018, 221, 329–345. [Google Scholar] [CrossRef]
  35. Lee, T.H.; Hansen, A.C.; Li, G.; Lee, T. Effects of isopropanol-butanol-ethanol and diesel fuel blends on combustion characteristics in a constant volume chamber. Fuel 2019, 254, 115613. [Google Scholar] [CrossRef]
  36. Li, G.; Lee, T.H.; Liu, Z.; Lee, C.-F.F.; Zhang, C. Effects of injection strategies on combustion and emission characteristics of a common-rail diesel engine fueled with isopropanol-butanol-ethanol and diesel blends. Renew. Energy 2019, 130, 677–686. [Google Scholar] [CrossRef]
  37. Chybowski, L.; Laskowski, R.; Gawdzińska, K. An overview of systems supplying water into the combustion chamber of diesel engines to decrease the amount of nitrogen oxides in exhaust gas. J. Mar. Sci. Technol. 2015, 20, 393–405. [Google Scholar] [CrossRef]
  38. Uyumaz, A. An experimental investigation into combustion and performance characteristics of an HCCI gasoline engine fueled with n-heptane, isopropanol and n-butanol fuel blends at different inlet air temperatures. Energy Convers. Manag. 2015, 98, 199–207. [Google Scholar] [CrossRef]
  39. Calam, A.; Aydoğan, B.; Halis, S. The comparison of combustion, engine performance and emission characteristics of ethanol, methanol, fusel oil, butanol, isopropanol and naphtha with n-heptane blends on HCCI engine. Fuel 2020, 266, 117071. [Google Scholar] [CrossRef]
  40. Sivasubramanian, H.; Pochareddy, Y.K.; Dhamodaran, G.; Esakkimuthu, G.S. Performance, emission and combustion characteristics of a branched higher mass, C 3 alcohol (isopropanol) blends fuelled medium duty MPFI SI engine. Eng. Sci. Technol. Int. J. 2017, 20, 528–535. [Google Scholar] [CrossRef]
  41. Baloch, H.A.; Nizamuddin, S.; Siddiqui, M.; Riaz, S.; Jatoi, A.S.; Dumbre, D.K.; Mubarak, N.; Srinivasan, M.; Griffin, G. Recent advances in production and upgrading of bio-oil from biomass: A critical overview. J. Environ. Chem. Eng. 2018, 6, 5101–5118. [Google Scholar] [CrossRef]
  42. Dominković, D.F.; Bačeković, I.; Pedersen, A.S.; Krajačić, G. The future of transportation in sustainable energy systems: Opportunities and barriers in a clean energy transition. Renew. Sustain. Energy Rev. 2018, 82, 1823–1838. [Google Scholar] [CrossRef]
  43. Trumbo, J.L.; Tonn, B.E. Biofuels: A sustainable choice for the United States’ energy future? Technol. Forecast. Soc. Chang. 2016, 104, 147–161. [Google Scholar] [CrossRef]
  44. Abbas, S.Z.; Kousar, A.; Razzaq, S.; Saeed, A.; Alam, M.; Mahmood, A. Energy management in South Asia. Energy Strat. Rev. 2018, 21, 25–34. [Google Scholar] [CrossRef]
  45. Patel, P.D.; Lakdawala, A.; Chourasia, S.K.; Patel, R. Bio fuels for compression ignition engine: A review on engine performance, emission and life cycle analysis. Renew. Sustain. Energy Rev. 2016, 65, 24–43. [Google Scholar] [CrossRef]
  46. Thangavelu, S.K.; Ahmed, A.S.; Ani, F.N. Review on bioethanol as alternative fuel for spark ignition engines. Renew. Sustain. Energy Rev. 2016, 56, 820–835. [Google Scholar] [CrossRef]
  47. Imran, S.; Emberson, D.; Wen, D.; Diez, A.; Crookes, R.; Korakianitis, T. Performance and specific emissions contours of a diesel and RME fueled compression-ignition engine throughout its operating speed and power range. Appl. Energy 2013, 111, 771–777. [Google Scholar] [CrossRef]
  48. Luo, H.; Zheng, P.; Bilal, M.; Xie, F.; Zeng, Q.; Zhu, C.; Yang, R.; Wang, Z. Efficient bio-butanol production from lignocellulosic waste by elucidating the mechanisms of Clostridium acetobutylicum response to phenolic inhibitors. Sci. Total. Environ. 2020, 710, 136399. [Google Scholar] [CrossRef] [PubMed]
  49. Çelebi, Y.; Aydın, H. An overview on the light alcohol fuels in diesel engines. Fuel 2019, 236, 890–911. [Google Scholar] [CrossRef]
  50. Yilmaz, N.; Ileri, E.; Atmanli, A. Performance of biodiesel/higher alcohols blends in a diesel engine. Int. J. Energy Res. 2016, 40, 1134–1143. [Google Scholar] [CrossRef]
  51. Markov, V.A.; Gaivoronsky, A.I.; Grechov, L.V.; Ivashchenko, N.A. Work of diesel engines on alternative fuels (Рабoта дизелей на нетрадициoнных тoпливах)-In Russian; Legion-Avtodata: Moscow, Russia, 2008; ISBN 978-5-88850-361-4. [Google Scholar]
  52. Rimkus, A.; Žaglinskis, J.; Stravinskas, S.; Rapalis, P.; Matijošius, J.; Bereczky, A. Research on the Combustion, Energy and Emission Parameters of Various Concentration Blends of Hydrotreated Vegetable Oil Biofuel and Diesel Fuel in a Compression-Ignition Engine. Energies 2019, 12, 2978. [Google Scholar] [CrossRef] [Green Version]
  53. Rimkus, A.; Pukalskas, S.; Matijošius, J.; Sokolovskij, E. Betterment of ecological parameters of a diesel engine using Brown‘s gas. J. Environ. Eng. Landsc. Manag. 2012, 21, 133–140. [Google Scholar] [CrossRef]
  54. Fuć, P.; Lijewski, P.; Ziolkowski, A.; Dobrzyński, M. Dynamic Test Bed Analysis of Gas Energy Balance for a Diesel Exhaust System Fit with a Thermoelectric Generator. J. Electron. Mater. 2017, 46, 3145–3155. [Google Scholar] [CrossRef]
  55. JCGM 100 2008. Evaluation of measurement data—Guide to the expression of uncertainty in measurement. Available online: https://www.bipm.org/utils/common/documents/jcgm/JCGM_100_2008_E.pdf (accessed on 13 April 2020).
  56. Evaluation of measurement data—Supplement 1 to the “Guide to the expression of uncertainty in measurement”—Propagation of distributions using a Monte Carlo method. Available online: https://www.bipm.org/utils/common/documents/jcgm/JCGM_101_2008_E.pdf (accessed on 27 April 2020).
  57. Gutarevych, Y.; Mateichyk, V.; Matijošius, J.; Rimkus, A.; Gritsuk, I.; Syrota, O.; Shuba, Y. Improving Fuel Economy of Spark Ignition Engines Applying the Combined Method of Power Regulation. Energies 2020, 13, 1076. [Google Scholar] [CrossRef] [Green Version]
  58. Zoldy, M. Fuel Properties of Butanol–Hydrogenated Vegetable Oil Blends as a Diesel Extender Option for Internal Combustion Engines. Period. Polytech. Chem. Eng. 2019, 64, 205–212. [Google Scholar] [CrossRef]
  59. Zöldy, M. Investigation of Correlation Between Diesel Fuel Cold Operability and Standardized Cold Flow Properties. Period. Polytech. Transp. Eng. 2019. [Google Scholar] [CrossRef] [Green Version]
  60. Kumar, S.; Pandey, A.K. Current Developments in Biotechnology and Bioengineering and Waste Treatment Processes for Energy Generation. Curr. Dev. Biotechnol. Bioeng. 2019, 1–9. [Google Scholar] [CrossRef]
  61. Zöldy, M. Improving Heavy Duty Vehicles Fuel Consumption with Density and Friction Modifier. Int. J. Automot. Technol. 2019, 20, 971–978. [Google Scholar] [CrossRef]
  62. Ghosh, P.; Jaffe, S.B. Detailed Composition-Based Model for Predicting the Cetane Number of Diesel Fuels. Ind. Eng. Chem. Res. 2006, 45, 346–351. [Google Scholar] [CrossRef]
  63. Gutarevych, Y.; Shuba, Y.; Matijošius, J.; Karev, S.; Sokolovskij, E.; Rimkus, A. Intensification of the combustion process in a gasoline engine by adding a hydrogen-containing gas. Int. J. Hydrog. Energy 2018, 43, 16334–16343. [Google Scholar] [CrossRef]
  64. Kukharonak, H.; Ivashko, V.; Pukalskas, S.; Rimkus, A.; Matijošius, J. Operation of a Spark-ignition Engine on Mixtures of Petrol and N-butanol. Procedia Eng. 2017, 187, 588–598. [Google Scholar] [CrossRef]
  65. Zöldy, M.; Hollo, A.; Thernesz, A. Butanol as a Diesel Extender Option for Internal Combustion Engines. SAE Tech. Paper Ser. 2010. [Google Scholar] [CrossRef]
  66. Hunicz, J.; Matijošius, J.; Rimkus, A.; Kilikevičius, A.; Kordos, P.; Mikulski, M. Efficient hydrotreated vegetable oil combustion under partially premixed conditions with heavy exhaust gas recirculation. Fuel 2020, 268, 117350. [Google Scholar] [CrossRef]
  67. Caban, J.; Droździel, P.; Ignaciuk, P.; Kordos, P. THE IMPACT OF CHANGING THE FUEL DOSE ON CHOSEN PARAMETERS OF THE DIESEL ENGINE START-UP PROCESS. Transp. Probl. 2019, 14, 51–62. [Google Scholar] [CrossRef] [Green Version]
  68. Kilikevičienė, K.; Kačianauskas, R.; Kilikevičius, A.; Maknickas, A.; Matijošius, J.; Rimkus, A.; Vainorius, D. Experimental investigation of acoustic agglomeration of diesel engine exhaust particles using new created acoustic chamber. Powder Technol. 2020, 360, 421–429. [Google Scholar] [CrossRef]
  69. Rimkus, A.; Matijošius, J.; Bogdevicius, M.; Bereczky, A.; Török, Á. An investigation of the efficiency of using O2 and H2 (hydrooxile gas -HHO) gas additives in a ci engine operating on diesel fuel and biodiesel. Energy 2018, 152, 640–651. [Google Scholar] [CrossRef]
  70. Zavadskas, E.K.; Čereška, A.; Matijošius, J.; Rimkus, A.; Baušys, R. Internal Combustion Engine Analysis of Energy Ecological Parameters by Neutrosophic MULTIMOORA and SWARA Methods. Energies 2019, 12, 1415. [Google Scholar] [CrossRef] [Green Version]
  71. Zöldy, M. Potential future renewable fuel challanges for internal combustion engine. Jármûvek és Mobilgépek II. évf 2009, 2, 397–403. [Google Scholar]
  72. Lijewski, P.; Fuć, P.; Dobrzynski, M.; Markiewicz, F. Exhaust emissions from small engines in handheld devices. MATEC Web Conf. 2017, 118, 16. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Image of the test engine: (a) test bench scheme; (b) test rig.
Figure 1. Image of the test engine: (a) test bench scheme; (b) test rig.
Energies 13 02398 g001
Figure 2. Digital Engine Model used in AVL BOOST.
Figure 2. Digital Engine Model used in AVL BOOST.
Energies 13 02398 g002
Figure 3. Dependence of Brake Specific Fuel Consumption and Brake Thermal Efficiency on different: (a) loads; (b) injection timing.
Figure 3. Dependence of Brake Specific Fuel Consumption and Brake Thermal Efficiency on different: (a) loads; (b) injection timing.
Energies 13 02398 g003
Figure 4. Dependence of carbon dioxide emissions and hydrocarbons on different: (a) loads; (b) injection timing.
Figure 4. Dependence of carbon dioxide emissions and hydrocarbons on different: (a) loads; (b) injection timing.
Energies 13 02398 g004
Figure 5. Dependence of nitrogen oxide emissions and smoke on different: (a) loads; (b) injection timing.
Figure 5. Dependence of nitrogen oxide emissions and smoke on different: (a) loads; (b) injection timing.
Energies 13 02398 g005
Figure 6. Dependence of pressure and pressure rise in the cylinder on the crank angle degree.
Figure 6. Dependence of pressure and pressure rise in the cylinder on the crank angle degree.
Energies 13 02398 g006
Figure 7. Dependence of the rate of heat release and the mass burnt fraction on the crank angle degree.
Figure 7. Dependence of the rate of heat release and the mass burnt fraction on the crank angle degree.
Energies 13 02398 g007
Table 1. Properties of 100% pure diesel, rapeseed methyl ester and isopropanol.
Table 1. Properties of 100% pure diesel, rapeseed methyl ester and isopropanol.
PropertiesDieselRapeseed Methyl EsterIsopropanol
Density (kg/m3)843877785.1
Mass Fraction (% mass): Carbon86.377.560
Hydrogen13.71213.4
Oxygen010.526.6
Stoichiometric AFR14.312.510.4
Lower Heating Value (LHV)(MJ/kg)42.337.832.8
Cetane number514812
Auto-ignition temperature (°C)250240399
Table 2. Main parameters of the 1.9 TDI diesel engine.
Table 2. Main parameters of the 1.9 TDI diesel engine.
ParameterValue
Displacement (cm3)1896
No. of cylinders4
Compression ratio19.5
Power (kW)66 (4000 rpm)
Torque (Nm)180 (2000–25,000 rpm)
Bore (mm)79.5
Stroke (mm)95.5
Intake valve opening at13 CAD before TDC
Intake valve closing at25 CAD after BDC
Exhaust valve opening at28 CAD before BDC
Exhaust valve closing at19 CAD after TDC
Fuel injectionDirect injection (single)
Fuel injection-pump designAxial-piston distributor injection pump
Nozzle typeHole-type
Nozzle and holder assemblyTwo-spring
Nozzle opening pressure (bar)200
Table 3. Uncertainty ranges u(x) of exhaust components.
Table 3. Uncertainty ranges u(x) of exhaust components.
Exhaust ComponentNumber of Cycles x ¯ Standard Uncertainly u(x)
CO (g/kWh)48300.0036
CO2 (g/kWh)48950.0041
HC (g/kWh)40.070.0005
NOx (g/kWh)413.50.0079
Smoke (m−1)47.30.0026
Table 4. Comparison of fuel properties of different fuel blends used and uncertainty ranges uc of each fuel blend parameter.
Table 4. Comparison of fuel properties of different fuel blends used and uncertainty ranges uc of each fuel blend parameter.
D100D50RME45P5D50RME40P10D50RME30P20
Density (kg/m3)843855.4850.8844.4
uc of Density (kg/m3)0.0080.00370.00320.0026
Mass Fraction (%): Carbon86.381.02580.1578.4
uc of Carbon (%)0.003330.003220.002150.00203
Hydrogen13.712.9212.9913.13
uc of Hydrogen 0.000640.000580.000470.00044
Oxygen06.0556.868.47
uc of Oxygen 00.000080.000090.00012
Lower Heating Value (MJ/kg)42.339.839.5539.05
uc of Lower Heating Value (MJ/kg)0.008140.007260.006330.00589
Cetane Number (-)5134.4931.8427.94
uc of Cetane Number (-)0.02550.03440.04670.0592

Share and Cite

MDPI and ACS Style

Rimkus, A.; Matijošius, J.; Manoj Rayapureddy, S. Research of Energy and Ecological Indicators of a Compression Ignition Engine Fuelled with Diesel, Biodiesel (RME-Based) and Isopropanol Fuel Blends. Energies 2020, 13, 2398. https://doi.org/10.3390/en13092398

AMA Style

Rimkus A, Matijošius J, Manoj Rayapureddy S. Research of Energy and Ecological Indicators of a Compression Ignition Engine Fuelled with Diesel, Biodiesel (RME-Based) and Isopropanol Fuel Blends. Energies. 2020; 13(9):2398. https://doi.org/10.3390/en13092398

Chicago/Turabian Style

Rimkus, Alfredas, Jonas Matijošius, and Sai Manoj Rayapureddy. 2020. "Research of Energy and Ecological Indicators of a Compression Ignition Engine Fuelled with Diesel, Biodiesel (RME-Based) and Isopropanol Fuel Blends" Energies 13, no. 9: 2398. https://doi.org/10.3390/en13092398

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