Article citation information:

Šarkan, B., Skrúcaný, T., Semanová, Š., Madleňák, R., Kuranc, A., Sejkorová, M., Caban, J. Vehicle coast-down method as a tool for calculating total resistance for the purposes of type-approval fuel consumption. Scientific Journal of Silesian University of Technology. Series Transport. 2018, 98, 161-172. ISSN: 0209-3324. DOI: https://doi.org/10.20858/sjsutst.2018.98.15.

 

 

Branislav Šarkan[1], Tomáš Skrúcaný[2], Štefánia Semanová[3],
Radovan MADLEŇÁK[4], Andrzej KURANC[5], Marie SEJKOROVÁ[6], Jacek CABAN[7]

 

 

 

VEHICLE COAST-DOWN METHOD AS A TOOL FOR CALCULATING TOTAL RESISTANCE FOR THE PURPOSES OF TYPE-APPROVAL FUEL CONSUMPTION

 

Summary. A coast-down test is carried out in cases when there is a need for the exact expression of the forces acting on a road vehicle during its coast-down. These forces act mainly against the vehicle’s movement due to air and tyre rolling resistance. Knowledge of the course of these forces throughout the vehicle’s movement range is also a requirement when measuring fuel consumption with a roller performance dynamometer. The reason is that this device has to load the rollers by force or performance corresponding to the given vehicle, while testing it during various driving cycles. For approval purposes, the requirements for this type of test are described in the applicable regulations, i.e., UNECE Regulations No. 83 and No. 101, or the newly developed Worldwide Harmonized Light Vehicles Test Procedure (WLTP). Slovak Technical Standard STN 30 0556 also contains detailed procedures for carrying out the test. The authors of this paper have taken into account both types of technical regulations in order to conduct coast-down tests on chosen vehicles. The results are usable in terms of measuring fuel consumption on a roller dynamometer. Furthermore, the vehicle’s economic performance in normal operation mode can be also assumed with these results.

Keywords: coast-down method, vehicle resistance, aerodynamic drag, tyre rolling resistance, fuel consumption

 

 

1. INTRODUCTION

 

A coast-down test is applied to determine resistance under real conditions. This is the test in which a vehicle running by its inertia is slowed down due to driving resistances from the moment of shifting the neutral position of the transmission until reaching zero speed. The test is carried out on a test track. The driving resistances represent forces that act in opposition to the movement of the vehicle running on a horizontal track. These include air resistance and mechanical resistance. Air resistance is caused by the effect of air flow as the vehicle passes through the air. Mechanical resistance represents tyre rolling resistance and resistances in bearings and parts of the transmission device [5 10,11,15,20].

 

 

2. measurement methodology

 

The vehicle is driven on a straight test track at constant speed under windless conditions; and, after shifting into neutral, it moves further forward by inertia. The vehicle is thus gradually slowed down due to rolling resistance, aerodynamic drag and friction resistance in the transmission device until it completely stops. Changes in the vehicle speed are continuously recorded. The vehicle movement is straight and unevenly slowed during the coast-down. Driving resistances are calculated based on the ascertained course of vehicle deceleration during the coast-down, as well as geometric parameters of the test track and the known vehicle’s inertia parameters. The coast-down characteristic of the vehicle is ascertained by coast-down tests. Based on this characteristic, it is possible to assess the mechanical condition of the vehicle, the chassis settings, the influence of used tyres and the aerodynamic properties of the vehicle body [14,19].

Several requirements relating to a vehicle, test track, test speed, atmospheric conditions and measuring equipment must be met when carrying out coast-down tests. These requirements differ depending on the type of regulation used. The requirements of the individual regulations regarding a vehicle, test track, atmospheric conditions and accuracy of the used measuring and recording equipment are listed in Table 1. The most detailed description of such requirements is provided in the WLTP. The results obtained in accordance with the WLTP are used for transferring the load of a road vehicle in normal operation mode onto the rollers of a roller performance dynamometer in order to determine the type-approval fuel consumption [25,30].

The methodology recommended by the WLTP involves performing each coast-down measurement without interruption. However, the division of the section is permitted if the data cannot be collected in a continuous manner across the full speed range. When dividing the measurement process, it is necessary to ensure that vehicle conditions remain as stable as possible at each point of division [29].

 

Tab. 1

Different requirements for carrying out the coast-down test [23,27,28,29]

Measurement requirements

Available methodologies for the coast-down

UNECE

No. 101

UNECE

No. 83

WLTP

STN 30 0556

Test track parameters

Longitudinal slope

±2%

±1.5%

±1%

±1%

Local inclination

±0.5%

-

±0.5%

-

Cross-sectional slope

Max. 1.5%

-

Max. 1.5%

Max. 1.5%

Atmospheric conditions

Wind speed (average)

Max. 3 m·s-1

Max. 3 m·s-1

Max. 7 m·s-1

Max. 1.5 m·s-1

Wind speed (gusts)

Max. 5 m·s-1

Max. 5 m·s-1

Max. 10 m·s-1

-

Wind cross vector component

Max. 2 m·s-1

Max. 2 m·s-1

Max. 4 m·s-1

-

Area of measurement above the surface

0.7 m

0.7 m

-

-

Air temperature

5- 35°C

-

5-40°C

5-25°C

Air pressure

91-104 kPa

-

-

97.33-101.25 kPa

Relative humidity

Max. 95%

-

-

-

Vehicle

Vehicle run-in

Min. 300 km

Min. 3,000 km

Min. 3,000 km

-

Tyre run-in

Min. 300 km

-

Min. 200 km

Min. 1,000 km

Tread depth

50-90%

50-90%

80-100%

67%

Time recording (frequency)

-

-

5 Hz

-

Measuring equipment
(accuracy)

Time recording (accuracy)

±0.1 s

±0.1 s

±0.01 s

-

Speed record (accuracy)

±0.5 km·h-1

±2%

±0.5 km·h-1

-

Wind speed

-

-

±0.3 m·s-1

-

Wind direction

-

-

±3°

-

Air temperature

-

-

±1 K

-

Air pressure

-

-

±0.3 kPa

-

Vehicle mass

-

-

±10 kg

-

Tyre pressure

-

-

±5 kPa

-

The measurements of coast-down are carried out in both directions repeatedly until at least three consecutive measurement pairs meet the requirements for statistical accuracy p in percentage terms as shown below (1).

 

                                                                                                                     (1)

 

where  represents the coefficient, which is determined by the WLTP depending on the number of performed pairs of coast-down measurements. For example, the coefficient 2.48 is used for n=3 and the coefficient 1.6 is applied for n=4. The standard deviation expressed in seconds is denoted as σ, while Δtj is the average time of the coast-down at the reference speed in seconds.

 

The standard deviation is calculated according to the following formula:

 

                                                                                                 (2)

 

where Δtji represents the harmonized average time of the coast-down of the i-th pair of measurements and is calculated according to the following formula:

 

                                                                                                                   (3)

 

where Δtjai and Δtjbi represent the time of the coast-down for the i-th measurement in each direction, respectively.

 

The calculated statistical accuracy of at least three consecutive coast-down tests in each direction (three pairs of measurements) should not exceed 3%, according to the WLTP [29].

After meeting the accuracy of the performed coast-down tests, coast-down curves are created based on the recorded data (Figure 1), while vehicle resistances are determined in newtons for the whole range of examined speed. To calculate these resistances, it is necessary to divide the coast-down curve into individual vehicle speed intervals. The WLTP recommends an interval width of 20 km·hod-1 for the coast-down from the speed higher, rather than 60 km·hod-1. The durations are assigned to the respective intervals of speed decreases, and thus the vehicle resistances are determined or the dependence of vehicle resistance on speed is ascertained.

Fig. 1. Coast-down curve

 

The resistance force for individual speed intervals is calculated according to the formula shown below (4). In this formula, mav represents the vehicle mass while tested. This is the average vehicle mass before and after carrying out the coast-down, while considering the consumed fuel. Further, mr is the equivalent effective mass of all the wheels and vehicle components rotating with the wheels during the coast-down on the road. It should be measured or calculated by an appropriate technique. Alternatively, mr may be estimated to be equal to 3% of the unladen vehicle mass when increased by 25 kg.

 

                                                                                                      (4)

 

If necessary, it is also possible to determine resistance forces for individual directions of the vehicle coast-down. This requires calculating the average times of the intervals for the relevant direction; however, the formula is the same as the previous one.

 

                                                                                              (5)

 

 

3. Measurement results

 

Three vehicles were used for practical tests on the coast-down (Figure 2). The first tested vehicle was that used in the laboratory of the Department of Road and Urban Transport, namely, a Kia Ceed 1.6 CVVT, which is a vehicle with a spark-ignition engine. The vehicle kerb weight is 1,163 kg, the air resistance coefficient is 0.33 and the vehicle frontal area is 2.25 m2. The second tested vehicle was a Ford Galaxy 2.0 Ghia, which has a diesel engine. The vehicle kerb weight is 1,799 kg, the air resistance coefficient is 0.314 and the vehicle frontal area is 2.78 m2. The third tested vehicle was a Hyundai i30 Wagon with a 1.6 CRDi diesel engine. The vehicle kerb weight is 1,413 kg, the air resistance coefficient is 0.3 and the vehicle frontal area is 2.136 m2. Before carrying out the coast-down tests, the vehicles were subjected to tyre pressure control and the tyres were inflated as required by the WLTP.

 

20160802_124811

 

Fig. 2. Tested vehicles (Kia Ceed, Ford Galaxy, Hyundai i30)

 

 

The coast-down tests were carried out on the road connecting Water Dam Žilina with the village of Mojš. The test track was 1.25 km long. The rest of the road section was used for vehicle acceleration and turning the vehicle in opposite direction. The road had an asphalted surface of sufficient quality. At the beginning of the section, it was necessary to achieve the required vehicle speed. The initial vehicle speed was 115 km·h-1 and the coast-down was recorded from the speed of 110 km·h-1. As the test track was not sufficiently long enough to carry out the whole coast-down, the test was divided into two measurement parts (from 110 km·h-1 to 60 km·h-1, and from 65 km·h-1 to 0 km·h-1). The test track is graphically illustrated in Figure 3.

 

 

Fig. 3. Test track [Google Maps]

 

 

During the coast-down measurements, atmospheric conditions were controlled and recorded by using a weather station with a thermometer, a moisture meter and an anemometer. All atmospheric conditions complied with the required level for the measurements. The wind speed was at the average level of 1.1 m·s-1 and air temperature was 14°C.

The universal OBD2 diagnostics device, which was compatible with the diagnostic interface equipped with the ELM327 chip and TouchScan software, was used to record vehicle speed over time. This device allows for recording data at a frequency of approximately 3 Hz. Before using the diagnostics device, it is advisable to calibrate a speed indicator, e.g., by using a roller performance dynamometer, which is also among the equipment available from the laboratory of Department of Road and Urban Transport [3,9,17,19]. The record of vehicle speed over time is saved in .txt format, meaning that it is possible to work with it in Microsoft Excel.

After processing the results of the coast-down for tested vehicles, the calculation of accuracy was made according to the formulas mentioned previously in the paper (see Chapter 2). In total, three coast-down tests in each direction of the measuring section for each tested vehicle were carried out. Each measured coast-down was divided into two measurement parts (from 110 km·h-1 to 60 km·h-1, and from 65 km·h-1 to 0 km·h-1) due to the insufficient length of the test track. Based on Table 2 and Figure 4, it can be concluded that the duration of the coast-down in Direction 1 was significantly shorter than the coast-down in Direction 2. This is caused by the longitudinal slope of the measuring section. The value of this slope was 1.1%, which slightly exceeds the required value according to the WLTP. Therefore, the harmonized average time of the coast-downs was applied to the calculations. In the case of first two vehicles, the required statistical accuracy of not exceeding 3% was met. However, the Ford Galaxy slightly exceeded this requirement. In the case of using this information for determining the type-approval fuel consumption or the official measurement of fuel consumption, it would have been necessary to repeat the measurements with this vehicle until the statistical accuracy was achieved at the required level.

 

Tab. 2

Calculation of statistical accuracy of the coast-down for the tested vehicles

 

Vehicle

Coast-down

 Direction 1

[s]

Coast-down

Direction 2

[s]

Coast-down

Direction 1

 [s]

Coast-down

Direction 2

 [s]

Coast-down

Direction 1

 [s]

Coast-down

Direction 2

 [s]

Harmonized average time Δtji

[s]

Average time of the coast-down Δtj

[s]

STD
σ [s]

Statistical accuracy
p

[%]

Kia Ceed

145.80

186.28

145.85

196.02

145.00

190.30

163.57

167.25

164.59

165.14

1.90

2.85

Hyundai i30

143.94

191.25

149.69

184.75

147.74

184.36

164.26

165.38

164.03

164.56

0.72

1.09

Ford Galaxy

149.90

205.77

150.80

209.70

153.33

213.89

173.45

175.44

178.62

175.83

2.61

3.68

 

It is possible to express the resistance force of the vehicle for each direction separately. Mainly in the case of roads with a longitudinal slope near to the tolerance limit, or in the case of wind direction in the longitudinal direction of the measuring section, differences may occur in the values for individual directions. For this reason, the harmonized average time (∆tj) of the alternating measurements of the coast-down must also be determined, so that the average total vehicle resistance (Fj) can be calculated. The processed results of the coast-down test for the Hyundai i30 are presented in detail in Table 3.

The outcome of this type of test is the determination of the total driving resistance of the vehicle while driving at constant speed on a horizontal road. If necessary, it is also possible to calculate the required performance (power) to overcome this resistance. The resultant resistance force expressed in newtons is multiplied by the actual vehicle speed in m·s-1, such  that the calculated performance required to overcome the resistance is expressed in watts. Using these types of calculation, it is also possible to determine the performance needed to overcome any speed or the maximum speed of the vehicle [7,8,10,11,14,32].

 

Fig. 4. Average times of the coast-down in the “there” and “back” directions for
the Hyundai i30

 

 

Tab. 3

Processed results of the coast-down for the Hyundai i30

 

Vehicle speed and speed intervals

[km·h-1]

Time of the coast-down for respective intervals for Direction 1

[s]

∆tja

Time of the coast-down for respective intervals for Direction 2

[s]

∆tjb

Fja

Fjb

∆tj

Fj

105

<110;100)

6.17

6.50

6.14

6.27

6.69

7.78

7.24

7.24

758.52

657.20

6.72

707.86

95

<100;90)

6.62

7.99

7.79

7.47

9.01

8.77

8.31

8.70

636.96

546.87

8.03

591.91

85

<90;80)

8.24

7.87

8.58

8.23

9.40

10.10

10.12

9.87

577.88

481.70

8.98

529.79

75

<80;70)

8.50

9.84

9.95

9.43

11.31

11.92

11.79

11.67

504.34

407.42

10.43

455.88

65

<70;60)

10.80

11.11

9.96

10.62

12.94

12.92

12.82

12.89

447.69

368.87

11.65

408.28

55

<60;50)

12.45

13.54

13.20

13.06

16.14

16.18

16.33

16.22

364.07

293.28

14.47

328.67

45

<50;40)

13.46

14.50

14.66

14.21

18.60

18.14

18.29

18.34

334.77

259.27

16.01

297.02

35

<40;30)

15.00

16.10

15.81

15.64

19.47

19.85

19.65

19.66

304.15

241.95

17.42

273.05

25

<30;20)

17.45

17.35

18.02

17.61

25.76

25.76

25.76

25.76

270.12

184.63

20.92

227.37

15

<20;10)

20.03

19.59

20.39

20.00

29.18

22.72

23.96

25.29

237.76

188.08

22.34

212.92

5

<10;0>

25.22

25.30

23.24

24.59

32.75

30.61

30.09

31.15

193.44

152.68

27.48

173.06

143.94

149.69

147.74

147.12

191.25

184.75

184.36

186.79

-

-

-

-

 

 

For type-approval fuel consumption purposes, the final result of this kind of test is the function of driving resistance curves. These curves are expressed as a quadratic function of velocity. Individual parameters then become the basis for transferring the load of a particular vehicle to the rollers of a roller performance dynamometer. Subsequently, this device is able to simulate respective driving resistances in relation to the actual speed of a vehicle moving on the rollers.

 

 

Fig. 5. Comparison of the average driving resistance of the tested vehicles

 

 

3. Conclusion

 

The issue of measuring vehicle resistance is addressed by several methodologies. For the type-approval fuel consumption purposes in 2018, it will be necessary to follow the methodology presented in the WLTP [20,29], which insists on significant requirements to be met relating not only to measuring equipment but also to the test track, atmospheric conditions and the vehicle itself. This paper has analysed the existing regulations for carrying out the coast-down test, as well as addressed the difficulty in quantifying driving resistances by using practical measurement examples involving three vehicles. The difficulty of the test lies primarily in processing the results with statistical tools in such a way that the resultant vehicle resistance corresponds to normal operation mode as realistically as possible.

 

 

References

 

1.             Barta D. 2012. Kolesové vozidlá – autobusy. Žilina: Žilinská univerzita. ISBN 978-80-554-0644-2 [In Slovak: Wheeled Vehicles – Buses. Žilina: University of Žilina.]

2.             Barta D., M. Mruzek, M. Kendra, P. Kordos, L. Krzywonos. 2016. “Using of non-conventional fuels in hybrid vehicle drives”. Advances in Science and Technology Research Journal 10(32): 240-247. ISSN 2299-8624. DOI: 10.12913/22998624/65108.

3.             Czaban J., D. Szpica. 2013. “Drive test system to be used on roller dynamometer”. Mechanika 19(5): 600-605. DOI: 10.5755/j01.mech.19.5.5542.

4.             Czech P. 2013. “Intelligent approach to valve clearance diagnostic in cars”. Activities of Transport Telematics. TST 2013. Communications in Computer and Information Science 395: 384-391. DOI: https://doi.org/10.1007/978-3-642-41647-7_47.

5.             Czech P. 2017. “Physically disabled pedestrians – road users in terms of road accidents”. Contemporary Challenges of Transport Systems and Traffic Engineering. Lecture Note in Networks and Systems 2: 157-165. DOI: https://doi.org/10.1007/978-3-319-43985-3_14.

6.             Czech P. 2017. “Underage pedestrian road users in terms of road accidents”. Intelligent Transport Systems and Travel Behaviour. Advances in Intelligent Systems and Computing 505: 33-44. DOI: https://doi.org/10.1007/978-3-319-43991-4_4.

7.             Drozdziel P., H. Komsta, L. Krzywonos. 2013. “Repair costs and the intensity of vehicle use”. Transport Problems 8(3): 131-138. ISSN 1896-0596.

8.             Drozdziel P., L. Krzywonos. 2009. “The estimation of the reliability of the first daily diesel engine start-up during its operation in the vehicle”. Eksploatacja i Niezawodnosc - Maintenance and Reliability 41(1): 4-10. ISSN 1507-2711.

9.             Fernandez-Yaneza P., O. Armasa, S. Martinez-Martinez. 2016. “Impact of relative position vehicle-wind blower in a roller test bench under climatic chamber”. Applied Thermal Engineering 106: 266. DOI: 10.1016/j.applthermaleng.2016.06.021.

10.         Figlus T., M. Stańczyk. 2016. “A method for detecting damage to rolling bearings in toothed gears of processing lines”. Metalurgija 55(1): 75-78. ISSN 0543-5846.

11.         Figlus T., M. Stańczyk. 2014. “Diagnosis of the wear of gears in the gearbox using the wavelet packet transform”. Metalurgija 53(4): 673-676. ISSN: 0543-5846.

12.         Gnap J., A. Kalašová, M. Gogola, J. Ondruš. 2010. “The Centre of Excellence for transport service and control”. Communications 12(3): 116-120. ISSN 1335-4205.

13.         Hockicko P., B. Trpišová, J. Ondruš. 2014. “Correcting students’ misconceptions about automobile braking distances and video analysis using interactive program tracker”. Journal of science education and technology 23(6):763-776. ISSN 1059-0145.

14.         James, D.J.G., K.J. Burnham, M.J. Richardson, R.A. Williams. 1999. “Improving vehicle performance using adaptive control techniques”. Artificial Life and Robotics 3(4): 236-241. DOI: 10.1007/BF02481187.

15.         Kalašová A., Ľ. Černický, M. Hamar. 2012. “A new approach to road safety in Slovakia”. In Transport Systems Telematics: 12th International Conference on Transport Systems Telematics: 388-395. 10-13 October 2012, Katowice-Ustroń, Poland. ISBN 978-3-642-34049-9. Available at: http://link.springer.com/content/pdf/10.1007%2F978-3-642-34050-5_44.pdf.

16.         Knez M., B. Jereb, M. Obrecht. 2014. “Factors influencing the purchasing decisions of low emission cars: a study of Slovenia”. Transportation Research Part D: Transport and Environment 30: 53-61. ISSN 1361-9309. DOI: 10.1016/j.trd.2014.05.007.

17.         Kubíková S., A. Kalašová, Ľ. Černický. 2014. “Microscopic simulation of optimal use of communication network”. In Telematics – Support for Transport. TST 2014. Communications in Computer and Information Science: 414-423. 22-25 October 2014, Katowice- Ustroń, Poland. ISBN 978-3-662-45316-2. DOI: https://doi.org/10.1007/978-3-662-45317-9_44.

18.         Kuranc A. 2015. “Exhaust emission test performance with the use of the signal from air flow meter”. Eksploatacja i Niezawodnosc – Maintenance and Reliability 17(1): 69-74.

19.         Labuda R., A. Kovalcik, J. Repka, V. Hlavna. 2012. “Simulation of a wheeled vehicle dynamic regimes in laboratory conditions”. Communications 14(3): 5-9. ISSN 1335-4205.

20.         Moravčík Ľ., Š. Liščák. 2012. “Zlepšovanie legislatívy pre schvaľovanie vozidiel”. [In Slovak: “Improvement of legislation on vehicle approval”.] In CMDTUR: Sixth International Scientific Conference: 228-243. 19-20 April 2012, University of Žilina, Žilina, Slovak Republic. ISBN 978-80-554-0512-4.

21.         Nadolski R., K. Ludwinek, J. Staszak, M. Jaśkiewicz. 2012. “Utilization of BLDC motor in electrical vehicles”. Przegląd Elektrotechniczny (Electrical Review) 2012 (4a). ISSN 0033-2097.

22.         Rievaj V, P. Faith, A. Dávid. 2006. “Measurement by a cylinder test stand and tyre rolling resistance”. Transport 21(1): 25-28. ISSN 1648-4142. DOI: http://dx.doi.org/10.1080/16484142.2006.9638036.

23.         STN 30 0556. Cestné vozidlá. Rýchlostné vlastnosti. Metódy skúšok. Bratislava: Úrad pre normalizáciu, metrológiu a skúšobníctvo SR. [In Slovak: Road Vehicles. Speed Characteristics. Test Methods. Bratislava: Slovak Office of Standards, Metrology and Testing.]

24.         Šipuš D., B. Abramović. 2017. “The possibility of using public transport in rural area”. In TRANSCOM 2017: International Scientific Conference on Sustainable, Modern and Safe Transport. Book Series: Procedia Engineering 192: 788-793.  DOI: 10.1016/j.proeng.2017.06.136. 

25.         TNO Innovation for Life. Road Load Determination of Passenger Cars. Accessed: 2 November 2017. Available at: https://www.tno.nl/media/1971/road_load_determination_passenger_cars_tno_r10237.pdf.

26.         Tomasikova M., M. Lukac, J. Caban, F. Brumercik. 2016. “Controllability and stability of a vehicle”. LOGI – Scientific Journal on Transport and Logistics 7(1): 136-142. ISSN 1804-3216.

27.         UNECE. Regulation No. 101. Uniform Provisions Concerning the Approval of Passenger Cars Powered by an Internal Combustion Engine Only, or Powered by a Hybrid Electric Power Train with Regard to the Measurement of the Emission of Carbon Dioxide and Fuel Consumption and/or the Measurement of Electric Energy Consumption and Electric Range, and of Categories M1 and N1 Vehicles Powered by an Electric Power Train Only with Regard to the Measurement of Electric Energy Consumption and Electric Range. Accessed: 9 November 2017. Available at: http://www.unece.org/fileadmin/DAM/trans/main/wp29/wp29regs/2015/R101r3e.pdf.

28.         UNECE. Regulation No. 83. Uniform Provisions Concerning the Approval of Vehicles with Regard to the Emission of Pollutants According to Engine Fuel Requirements. Accessed: 9 November 2017. Available at: https://www.unece.org/fileadmin/DAM/trans/main/wp29/wp29regs/r083r4e.pdf.

29.         UNECE. Global Technical Regulation No. 15. World Harmonized Light Vehicles Test Procedure. Accessed: 9 November 2017. Available at: https://www.unece.org/fileadmin/DAM/trans/main/wp29/wp29r-1998agr-rules/ECE-TRANS-180a15e.pdf.

30.         UNECE. World Forum for the Harmonization of Vehicle Regulation (WP.29).. Available at: https://www.unece.org/trans/main/welcwp29.html.

31.         Yanowitz J., M.S. Graboski, L.B.A. Ryan, T.L. Alleman, R.L. McCormick. 1999. “Dynamometer study of emissions from 21 in-use heavy-duty diesel vehicles”. Environmental Science Technology 33(2): 209-216. DOI: 10.1021/es980458p.

32.         Zbigniew M., M. Jaśkiewicz, K. Ludwinek, Z. Gawęcki. 2015. “Special characteristics of reliability for serial mechatronic systems”. In Selected Problems of Electrical Engineering and Electronics (WZEE). The Kielce University of Technology and The Polish Society of Theoretical and Applied Electrotechnics – Kielce Branch. 17-19 September 2015, Kielce, Poland. ISBN 978-1-4673-9452-9.

 

 

Received 05.11.2017; accepted in revised form 12.02.2018

 

 

Scientific Journal of Silesian University of Technology. Series Transport is licensed under a Creative Commons Attribution 4.0 International License



[1] Faculty of Operation and Economics of Transport and Communications, University of Žilina, Univerzitná 1, 010 26 Žilina, Slovak Republic. E-mail: branislav.šarkan@fpedas.uniza.sk.

[2] Faculty of Operation and Economics of Transport and Communications, University of Žilina, Univerzitná 1, 010 26 Žilina, Slovak Republic. E-mail: tomas.skrucany@fpedas.uniza.sk.

[3] Faculty of Operation and Economics of Transport and Communications, University of Žilina, Univerzitná 1, 010 26 Žilina, Slovak Republic. E-mail: stefania.semanova@fpedas.uniza.sk.

[4] Faculty of Operation and Economics of Transport and Communications, University of Žilina, Univerzitná 1, 010 26 Žilina, Slovak Republic. E-mail: radovan.madlenak@fpedas.uniza.sk.

[5] Faculty of Production Engineering, University of Life Sciences in Lublin, 28 Głęboka Street, 20-612 Lublin, Poland. E-mail: andrzej.kuranc@up.lublin.pl.

[6] Faculty of Transport Engineering, University of Pardubice, Študentská 95, 532 10 Pardubice, Czech Republic. E-mail: marie.sejkorova@upce.cz.

[7] Faculty of Mechanical Engineering, Lublin University of Technology, Nadbystrzycka 36, 20-618 Lublin, Poland. E-mail: j.caban@pollub.pl.