You are on page 1of 183

Dissertation presented to the Instituto Tecnológico de Aeronáutica, in partial

fulfillment of the requirements for the degree of Master of Science in the Program
of Aeronautical and Mechanical Engineering, Field of Aerodynamics, Propulsion,
and Energy.

Santiago Daniel Martinez Boggio

EXPERIMENTAL INVESTIGATION ON THE COMBUSTION

PROCESS IN A SPARK IGNITION OPTICALLY

ACCESSIBLE ENGINE FUELED WITH SYNGAS

Dissertation approved in its final version the signatories below:

Prof. Dr. Pedro Teixeira Lacava


Advisor

Prof. Dr. Pedro Luis Curto Risso


Co-advisor

Prof. Dr. Pedro Teixeira Lacava


Pro-Rector of Graduate Courses

Campo Montenegro
São José dos Campos, SP – Brasil
2018
Cataloging-in-Publication Data
Documentation and Information Division
Martinez Boggio, Santiago Daniel
Experimental investigation on the combustion process in a spark ignition optically accessible engine
fueled with Syngas/ Santiago Daniel Martinez Boggio.
São José dos Campos, 2018.
Número de folhas no formato 183

Dissertation of Master of Science – Aeronautical and Mechanical Engineering, Field of


Aerodynamics, Propulsion and Energy – Instituto Tecnológico de Aeronáutica, 2018. Advisor: Pedro
Teixeira Lacava, Co-Advisor: Pedro Luis Curto Risso.

1. Syngas. 2. Hydrogen. 3. Optical Investigation. I. Instituto Tecnológico de Aeronáutica.


II. Experimental investigation on the combustion process in a spark ignition optically accessible engine
fueled with Syngas.

BIBLIOGRAPHIC REFERENCE
MARTINEZ BOGGIO, Santiago Daniel. Experimental investigation on the combustion
process in a spark ignition optically accessible engine fueled with Syngas. Dissertation of
Master of Science in Aeronautical and Mechanical Engineering, Field of Aerodynamics,
Propulsion, and Energy – Instituto Tecnológico de Aeronáutica, São José dos Campos.

CESSION OF RIGHTS
AUTOR NAME: Santiago Daniel Martinez Boggio
PUBLICATION TITLE: Experimental investigation on the combustion process in a spark
ignition optically accessible engine fueled with Syngas.
PUBLICATION KIND/YEAR: Dissertation / 2018

It is granted to Instituto Tecnológico de Aeronáutica permission to reproduce copies of this


dissertation to only loan or sell copies for academic and scientific purposes. The author reserves
other publication rights and no part of this dissertation can be reproduced without his
authorization.

__________________________________
Santiago Daniel Martinez Boggio
Praça Paris, 114
CEP: 12216-780, São José dos Campos - SP
EXPERIMENTAL INVESTIGATION ON THE COMBUSTION

PROCESS IN A SPARK IGNITION OPTICALLY

ACCESSIBLE ENGINE FUELED WITH SYNGAS

Santiago Daniel Martinez Boggio

Thesis Committee Composition:

Prof. Dr. Ezio Castejon Garcia ChairPerson - ITA


Prof. Dr. Pedro Teixeira Lacava Adviser - ITA
Prof. Dr. Pedro Luis Curto Risso Co-adviser - UDELAR
Prof. Dr. Waldir Antonio Bizzo - UNICAMP
Prof. Dr. Nicolau André Silveira Rodrigues - ITA

ITA
iv

Acknowledgments

I would like to thank my advisor Prof. Dr. Pedro Teixeira Lacava for the opportunity
that gives me to work in the LCPE. Also, his guidance and support in the performing of the
experiments and the correction of the present manuscript.
I would like to express my special gratitude and thanks to my co-advisor Prof. Dr. Pedro
Luis Curto Risso for the support and time dedicated to my development as a student and as a
person. Moreover, his guidance and support in the performance of the numerical simulation test
have enriched this work, as well as in the correction of the present manuscript.
My gratitude also goes to Dra. Leila Ribeiro, Dra. Esther Sbampato, and Dr. Gilberto
Barreta for their support in the experimental tests and constructive comments. I am also thankful
for their capacity to create a stimulating and friendly research environment at the Lab.
The accuracy of the optical methodology and the international recognition of this work
would never have happened if not for the support of Dra. Simona Merola and Dr. Adrian
Irimescu from Istituto Motori. I would like to thanks for their time, patient and efforts dedicated
to my development as a student.
My gratitude also goes to the post-graduate students and friends of LCPE, for their help
in various aspects of the post-graduate life. I want to thanks: Alexander, Maycon, Armando,
Fernanda, Pedro, Flavio, Diego and Gabriel.
My deepest and sincerest thanks to my parents Daniel and Liliana and my siblings
Guillermo and Guzman for believing in me and supporting my decisions during my entire life.
Also, thanks to my girlfriend Luciana for the love, support and happiness that you gave me
along these years of Master.
Finally, I would like to thank the Brazilian federal agency CAPES for the financial
support of a two years scholarship and Universidad de La Republica for the financial support
in several moments of the master.
v

"Nacimos para triunfar”.

(Avelino Sanchez)
vi

Abstract

The optimization of SI engines fueled with syngas requires a deeper understanding of the related
thermal mechanisms in engine-like conditions. Therefore, to improve current knowledge and
provide reference data for modeling and simulation of internal combustion engines,
experimental and numerical investigations were carried out on a port fuel injection (PFI) spark-
ignition (SI) optical research engine fueled with pure syngas (CO/H2). Methane and methane-
H2 blends was considered for comparison. The analysis of different air dilution, spark timing
and fuel dilution was performed. The engine was operated at fixed rotational speed (900 RPM)
at wide open throttle (WOT) as representative condition of energy production applications. In-
cylinder pressure and related thermodynamic parameters were analyzed as indicators of
combustion behavior. Exhaust emission was measured in order to compare the different
operative conditions. 2D cycle resolved digital visualization was performed to follow the flame
front propagation. Custom image processing was applied to estimate the flame speed and others
morphology parameters. A quasi-dimensional model was validate to simulate the first phase of
the combustion process with experimental thermodynamic parameters (in-cylinder pressure and
mass fraction burned) and then a comparison with the optical results was performed. The
hydrogen addition improves the flame propagation speed, reduce flame distortion and center
displacement for syngas and methane blends. In addition, was seen a decrease in terms of cycle-
by-cycle variation and extension of the flammability limit for both fuels. Also, the advance of
spark timing shows an increase in terms of flame propagation speed and flame distortion. The
optimum spark timing (MBT) was found in the expansion stroke for pure syngas in
stoichiometric and intermediate air-fuel ratio. Moreover, spark timing re-calibration is required
in order to fully take advantage of fuel properties such as higher laminar flame speed and
increased stability. Lastly, air and fuel dilution were found to be a good strategy for decrease
NOx emissions without losing efficiency and good flame propagation properties (speed,
distortion and flame center displacement).
vii

Resumo
A otimização de motores alimentados com gás de síntese exige um entendimento mais profundo
dos mecanismos térmicos relacionados em condições de operação reais. Portanto, para melhorar
o conhecimento atual e fornecer dados de referência para modelagem e simulação de motores
de combustão interna, investigações numéricas e experimentais foram realizadas em um motor
de pesquisa de ignição por centelha (SI) com injeção na porta de admissão (PFI) com gás de
sínteses puro (misturas de CO/H2). Metano e misturas de metano-H2 foram consideradas para
comparação. A análise de diferentes diluições do ar, tempo da centelha e diluição do
combustível foram estudadas. O motor foi operado em velocidade rotacional fixa (900 RPM)
em aceleração máxima (WOT) como condição representativa de aplicações para produção de
energia. A pressão no cilindro e os parâmetros termodinâmicos relacionados foram analisados
como indicadores do comportamento de combustão. As emissões no escapamento foram
medidas para comparar as diferentes condições operacionais. A visualização 2D da propagação
da chama foi obtida a partir de imagens com uma câmera de alta velocidade ao longo dos ciclos.
Uma metodologia para o processamento de imagens foi desenvolvida para estimar a velocidade
de propagação, além de outros parâmetros morfológicos. Por outro lado, um modelo quase-
dimensional foi validado para fase inicial da queima contra parâmetros termodinâmicos
experimentais (pressão no cilindro e fração de massa queimada) e com isso uma comparação
com dados ópticos foram realizadas. A adição de hidrogênio melhorou a velocidade de
propagação da chama, além de reduzir a distorção e o deslocamento do centro da chama para
misturas de gás de sínteses e metano. Além disso, foi observado uma diminuição em termos de
variação ciclo a ciclo e uma extensão do limite de inflamabilidade para ambos combustíveis. O
avanço do momento da ignição mostra um aumento em termos de velocidade de propagação e
distorção da chama. O tempo ótimo de centelha foi encontrado no curso de expansão para o gás
de sínteses na relação estequiométrica ar-combustível. Além disso, foi observado que a
re- calibragem do tempo de ignição é necessária para tirar o máximo proveito das propriedades
do combustível, como maior velocidade de chama laminar e maior estabilidade. A diluição do
ar e do combustível foram consideradas uma boa estratégia para diminuição das emissões de
NOx sem perder eficiência e boas propriedades na propagação da chama (velocidade, distorção
e afastamento do centro da chama da centelha).
viii

List of Figures

Figure 1 - Processes involved in biomass gasification. ............................................................ 30

Figure 2 – Scheme of a Downdraft Fixed-Bed Gasifier, adapted from [21]. ........................... 32

Figure 3 – Experimental setup: a) single cylinder research engine mounted in the LCPE b)

scheme of the optical experimental arrangement with the camera set up for taking the bottom

view of the combustion chamber. Adapted from [55]. ............................................................. 45

Figure 4 – Optical access: a) cylinder head bottom view from an inclined mirror located below

the piston and b) piston design with the detail of optical window, top land region, and piston

ring positions. Adapted from [55]. ........................................................................................... 45

Figure 5 – Fuel system with representation of: gas storage, panel to prepare the mixture, gas

chromatographer and PFI fuel injector. .................................................................................... 48

Figure 6 – In-cylinder pressure and auxiliary measurements (motored pressure, spark plug

signal, camera trigger and PFI injection pulse) for methane at baseline conditions (λ=1.0, SA =

-7 CAD ATDC). ....................................................................................................................... 50

Figure 7- a) In-cylinder pressure trace for the average of 200 cycles and several single pressure

traces. b) Indicate mean effective pressure (IMEP) and Maximum In-cylinder pressure (Pmax)

values for 200 consecutive cycles for methane at baseline conditions (900 RPM, λ=1.0, SA = -

7 CAD ATDC). ........................................................................................................................ 51

Figure 8 – Sketch of the image processing steps; intake, exhaust valves, spark plug, and optical

limit are also shown. The axes x, y mark the positive values for the calculus of the center

movement. ................................................................................................................................ 57

Figure 9 – Overlap between the binary image of the flame front and the calculated equivalent

diameter with the WDD method............................................................................................... 58


ix

Figure 10 - Flame size measurements as the average of 25 consecutive cycles of methane for

λ=1.0 and SA= -7 ATDC a) Normalize flame area b) Flame diameter calculated as WDD. .. 58

Figure 11 – Representation of calculus of flame front propagation speed. .............................. 59

Figure 12 - Flame propagation speed measured as an average of 25 consecutive cycles for

methane with λ=1.0 and SA=7 BTDC...................................................................................... 60

Figure 13 – Distortion of the flame front measured as an average of 25 consecutive cycles

calculated with the HCF equation for methane with λ=1.0 and SA=7 BTDC. ........................ 61

Figure 14 – Representation of flame outline for the calculus of HCF and position of the

geometric centroid with respect to the center of the combustion chamber. ............................. 61

Figure 15 - Average flame front movement in the combustion chamber for methane in λ=1.0

and SA=7 BTDC. ..................................................................................................................... 62

Figure 16 – Combustion process lateral view of methane at 900 RPM, λ=1.0,

SA = - 7 CAD ATDC and wide open throttle (WOT). ............................................................ 63

Figure 17 – In-cylinder average pressure of 200 consecutives cycles to methane, S50 and S75

at SA = 7 CAD BTDC in: (a) λ=1.0 and (b) λ=1.4. ................................................................. 65

Figure 18 – Average pressure with S50 and SA = 7 CAD BTDC for the three air-fuel ratios.66

Figure 19 – IMEP average values for 200 consecutive cycles, for M, S50, and S75 in λ range

from 1.0 to 1.4 at SA = 7 CAD BTDC. .................................................................................... 67

Figure 20 – CovIMEP average values for 200 consecutive cycles, for M, S50, and S75 in λ

range from 1.0 to 1.4 at SA = 7 CAD BTDC. .......................................................................... 68

Figure 21 – Pmax (top) and CovPmax (bottom) average values for 200 consecutive cycles, for

M, S50, and S75 in λ range from 1.0 to 1.4 at SA = 7 CAD BTDC. ....................................... 69

Figure 22 – MFB average values for 200 consecutive cycles, for M, S50, and S75 at

SA=- 7 CAD ATDC and λ: (a) 1.0 and (b) 1.4. ....................................................................... 70
x

Figure 23 – MFB average values for 200 consecutive cycles S50 at SA = 7 CAD BTDC and λ

range from 1.0 to 1.4. ............................................................................................................... 71

Figure 24 – Fuel conversion efficiency for M, S50, and S75 with three fuel ratios and SA = 7

CAD BTDC. ............................................................................................................................. 72

Figure 25 – Emission concentration on the exhaust port for M, S50, and S75 at λ=1.4. ......... 73

Figure 26– NOx emission concentration on the exhaust port for M, S50, and S75 at λ=1.4.... 73

Figure 27 – Images of M, S50 and S75 combustion process in the position of 5%, 10% 20%,

30%, 40% and 60% with respect to the total piston area at λ 1.0 and SA = 7 CAD BTDC. ... 75

Figure 28 – Images of M, S50 and S75 combustion process in several crank angle position

positions at λ=1.0 and λ=1.4 with SA = 7 CAD BTDC. .......................................................... 76

Figure 29 – Flame front evolution for M, S50, and S75 at λ=1.0 and SA = -7 CAD ATDC

obtained by averaged data over 25 consecutive engine cycles: (a) Normalized flame area (b)

Flame propagation speed. ......................................................................................................... 77

Figure 30 – Flame front evolution for M, S50, and S75 at λ=1.2 and SA = 7 CAD BTDC

obtained by averaged data over 25 consecutive engine cycles: (a) Normalized flame area (b)

Flame propagation speed. ......................................................................................................... 78

Figure 31 – Flame front evolution for M, S50, and S75 at λ=1.4 and SA = 7 CAD BTDC

obtained by averaged data over 25 consecutive engine cycles: (a) Normalized flame area (b)

Flame propagation speed. ......................................................................................................... 78

Figure 32 – Maximum Flame Speed Propagation for M, S50, and S75 with λ range from 1.0 to

1.4 and SA = 7 CAD BTDC. .................................................................................................... 79

Figure 33 – HCF for M, S50 and S75 at λ=1.0 and SA = 7 CAD BTDC obtained by averaged

data over 25 consecutive engine cycles. ................................................................................... 80

Figure 34 – HCF obtained by averaged data over 25 consecutive engine cycles for λ range from

1.0 to 1.4 and SA = -7 CAD ATDC fueled with (a) S50 and (b) S75. ..................................... 80
xi

Figure 35 – Maximum HCF for M, S50, and S75 with λ range from 1.0 to 1.4 and SA = - 7 CAD

ATDC. ...................................................................................................................................... 81

Figure 36 – Mean flame center movement for M, S50 and S75 at SA = -7 CAD ATDC in (a)

λ=1.0 and (b) λ=1.4. ................................................................................................................. 82

Figure 37 – Maximum distance from the center of the combustion chamber for M, S50 and S75

with λ range from 1.0 to 1.4 and SA = 7 CAD BTDC. ............................................................ 82

Figure 38 – In-cylinder pressure traces averaged over 200 consecutive cycles for λ=1.4

a) methane-H2 blends and b) pure syngas blends compare with methane................................ 85

Figure 39 – In-cylinder pressure traces averaged over 200 consecutive cycles for extreme lean

condition a) methane-H2 blends and b) pure syngas blends compare with methane. .............. 86

Figure 40 – Mass fraction burned traces averaged over 200 consecutive cycles for a) λ=1.4 and

b) lean extreme case. ................................................................................................................ 87

Figure 41 – CO (top) and CH4 (bottom) concentration in ppm on the exhaust port for λ=1.4 and

extreme lean condition.............................................................................................................. 89

Figure 42 – NOx concentration in ppm on the exhaust port for λ=1.4 and extreme lean

conditions. ................................................................................................................................ 89

Figure 43 – Flame image sequence for all fuels with λ=1.4 at 5%, 10%, 20%, 30%, 40% and

60% area of the piston cross-section with exposure time 1.0 CAD for methane blends and 0.5

CAD for syngas. ....................................................................................................................... 91

Figure 44 – Flame image sequence for all fuels in extreme lean conditions, at 5%, 10%, 20%,

30%, 40% and 60% area of the piston cross-section with exposure time 1.0 CAD for methane

blends and 0.5 CAD for syngas. ............................................................................................... 91

Figure 45 – Evolution of a) flame front area and b) flame propagation speed for λ=1.4 obtained

by averaged data over 25 consecutive engine cycles. .............................................................. 93


xii

Figure 46 – Evolution of a) flame front area and b) flame propagation speed for extreme lean

cases obtained by averaged data over 25 consecutive engine cycles. ...................................... 93

Figure 47 – Heywood circularity factor of the flame for (a) λ=1.4 and (b) extreme lean cases

obtained by averaged data over 25 consecutive engine cycles. ................................................ 94

Figure 48 – Flame centroid position for (a) λ=1.4 and (b) extreme lean cases, obtained by

averaged data over 25 consecutive engine cycles. ................................................................... 95

Figure 49 – Maximum flame displacement in vertical axes in intake direction....................... 95

Figure 50 – Performance parameters (IMEP, CovIMEP, and fuel conversion efficiency) at

stoichiometric air-fuel ratio for S50 and several spark timing. .............................................. 100

Figure 51 – In-cylinder pressure traces for stoichiometric fueling of syngas 50% H2 and several

spark timing versus a) crank angle degree and b) normalized cylinder volume .................... 102

Figure 52 - Comparison of maximum in-cylinder pressure and NOx emission for several spark

timings. ................................................................................................................................... 103

Figure 53 – Mass fraction burned to stoichiometric air-fuel ratio with S50 and several sparks

timing with respect to: (a) top dead center (TDC) and (b) start of spark (SA). ..................... 104

Figure 54 – Zoom view of mass fraction burned for stoichiometric air-fuel ratio with S50 and

several sparks timing with respect to the start of spark (SOS). .............................................. 104

Figure 55 – Images of the flame at the same CAD after the start of the spark of Syngas with

50% of H2 in λ=1.0 for different spark timing. F-number: 32, exposure time: 0.5 CAD and

intensification: 55000. ............................................................................................................ 106

Figure 56 – Evolution of the flame front area and flame propagation speed for S50 and several

spark timings by averaged data over 30 consecutive engine cycles. ...................................... 107

Figure 57 – Maximum flame speed S50 and several spark timings by averaged data over 25

consecutive engine cycles....................................................................................................... 108


xiii

Figure 58 – Evolution of average Heywood circularity factor for S50 at several spark timings.

................................................................................................................................................ 109

Figure 59 – Maximum HCF for S50 at several spark timings................................................ 109

Figure 60 – The average flame center movement for S50 at several spark timings. .............. 109

Figure 61 – In-cylinder average pressure curve for 200 cycles at λ=1.4 for: a) methane-hydrogen

blends and b) pure syngas blends compare with pure methane. ............................................. 112

Figure 62 – In-cylinder average pressure curve for 200 cycles at extreme lean condition for: a)

methane-hydrogen blends and b) pure syngas blends compare with pure methane. .............. 113

Figure 63– MFB curve for 200 cycles of methane, M25, M50, S50 and S75 at spark timing for

MBT a) λ=1.4 and b) extreme lean condition. ....................................................................... 114

Figure 64 – Fuel conversion efficiency for M, M25, M50, S50 and S75 at spark timing MBT

for a) λ=1.4 and b) extreme lean condition. ........................................................................... 115

Figure 65 - CO emission in [ppm] for the five blends study at (a) lean and (b) extreme lean

condition for MBT spark advance timing. ............................................................................. 117

Figure 66 - CH4 emission in [ppm] for the five blends study at (a) lean and (b) extreme lean

condition for MBT spark advance timing. ............................................................................. 118

Figure 67- NOx emission in [ppm] for the five blends study at (a) lean and (b) extreme lean

condition for MBT spark advance timing. ............................................................................. 119

Figure 68 – Evolution of the flame front area and flame propagation speed for λ=1.4 obtained

by averaged data over 25 consecutive engine cycles in MBT condition................................ 120

Figure 69 – Evolution of the flame front area and flame propagation speed for λ=1.4 obtained

by averaged data over 25 consecutive engine cycles in MBT condition................................ 121

Figure 70 – Evolution of the flame front area and flame propagation speed for extreme lean

condition obtained by averaged data over 25 consecutive engine cycles in MBT condition. 122
xiv

Figure 71 - Heywood circularity of the flame for (a) λ=1.4 and (b) extreme lean condition

obtained by averaged data over 25 consecutive engine cycles in MBT condition. ................ 123

Figure 72 – Flame centroid position for (a) λ=1.4 and (b) extreme lean condition, obtained by

averaged data over 25 consecutive engine cycles in MBT condition..................................... 124

Figure 73 – In-cylinder pressure traces averaged over 200 consecutive cycles at λ=1.0 for: (a)

50% H2 blends and (b) 75% H2 blends, compare with methane. .......................................... 131

Figure 74 – In-cylinder pressure traces averaged over 200 consecutive cycles at λ=1.4 for: (a)

50% H2 blends and (b) 75% H2 blends, compare with methane. .......................................... 132

Figure 75 – IMEP evaluated average on 200 consecutive engine cycles for different air-fuel

ratios (λ=1.0-1.4) for M, S50, S5050 and S7550 in fixed SA (7 CAD BTDC). .................... 133

Figure 76 – CovIMEP evaluated on 200 consecutive engine cycles for different air-fuel ratios.

................................................................................................................................................ 134

Figure 77 – Fuel conversion efficiency for all the selected conditions and fuels. ................. 136

Figure 78 – Images of flame evolution for stoichiometric air-fuel ratio with M, S50, and S5050.

................................................................................................................................................ 138

Figure 79 – Images of flame evolution for stoichiometric air-fuel ratio with M, S75, and S7550.

................................................................................................................................................ 139

Figure 80 – Images of flame evolution for λ=1.4 with M, S50, and S5050. .......................... 140

Figure 81 – Images of flame evolution for λ=1.4 with M, S75, and S7550. .......................... 140

Figure 82 - Evolution of the flame front area and propagation speed for selected stoichiometric

air-fuel ratio obtained by averaged data over 25 consecutive engine cycles.......................... 141

Figure 83 - Evolution of the flame front area and propagation speed for λ=1.2 obtained by

averaged data over 25 consecutive engine cycles. ................................................................. 142

Figure 84 - Evolution of the flame front area and propagation speed for λ=1.4 obtained by

averaged data over 25 consecutive engine cycles. ................................................................. 142


xv

Figure 85 – Maximum flame propagation speed obtained by averaged data over 25 consecutive

engine cycles........................................................................................................................... 143

Figure 86 – Heywood Circularity Factor for stoichiometric air-fuel ratio obtained by average

data over 25 consecutive engine cycles. ................................................................................. 144

Figure 87 – Maximum Heywood Circularity Factor obtained by average data over 25

consecutive engine cycles for all air-fuel ratios. .................................................................... 144

Figure 88 – Evolution of the luminous centroid with respect to the geometrical center of the

combustion chamber; each step of the path is obtained by averaging data over 25 consecutive

engine cycles for the stoichiometric air-fuel ratio. ................................................................. 145

Figure 89 – Maximum centroid displacement with respect to the geometrical center of the

combustion chamber in the y-direction; each step of the path is obtained by averaging data over

25 consecutive engine cycles. ................................................................................................. 145

Figure 90- Experimental and simulated pressure traces for stoichiometric air-fuel ratio and SA

-7 CAD ATDC for: a) methane and b) S50. ........................................................................... 157

Figure 91- Experimental and simulated mass fraction burned traces for stoichiometric air-fuel

ratio and SA -7 CAD ATDC for: a) methane and b) S50....................................................... 158

Figure 92 – Comparison between simulated and experimental normalized flame area for

stoichiometric air-fuel ratio and SA -7 CAD ATDC.............................................................. 159

Figure 93 - Comparison of the average measured (exp) and simulated (sim) results recorded

during 25 consecutive cycles for a) CAD position of 30% of the normalized flame area and

b) flame diameter in 5 CAD ASOS, at 900 RPM and spark timing of 7 CAD BTDC. ......... 160

Figure 94 – Flame area against volume fraction burned for simulated and experimental

measurements for stoichiometric air-fuel ratio and fixed spark timing for a) methane b) S50.

................................................................................................................................................ 160
xvi

Figure 95 - Flame Area with dispersion bars for methane λ=1.0, fn=11, exp time=0.5 CAD,

Intensification=65000. ............................................................................................................ 175

Figure 96- Flame Area comparison between different amount of cycles for methane λ=1.0,

fn=11, exp time=0.5 CAD, Intensification=55000. ................................................................ 176

Figure 97 - Comparison Flame Area in the same case for three repetitive combustion process

for methane λ=1.0, f-number=11, intensification=55000 and exposure time 0.5 CAD. ........ 177

Figure 98- Comparison Flame Area with different values of exposure time for methane λ=1.0,

f-number=11, intensification=55000. ..................................................................................... 178

Figure 99- Comparison Flame Area with different values of intensification for methane λ=1.0,

f-number=32, exposure time=0.5 CAD. ................................................................................. 179

Figure 100- Comparison Flame Area with different values of f-number for methane λ=1.0,

exposure time=0.5 CAD, intensification=55000. ................................................................... 180

Figure 101 - Numerical and experimental data of 𝑪𝑫 for the calculus of mass flow rate through

the valves [135]. ..................................................................................................................... 182

Figure 102 – Valve lift for intake and exhaust valve. ............................................................ 182
xvii

List of Tables

Table 1 - Characteristics of the syngas for different types of gasifiers with dry wood as biomass

feedstock and air as oxidizer [2]. .............................................................................................. 32

Table 2 - Effect of gasify agent on syngas composition for different types of gasifiers with dry

wood [2].................................................................................................................................... 33

Table 3 – List of important studies in SI engines with syngas in recent years. ........................ 36

Table 4 – Geometric specifications of the PFI-SI single-cylinder research engine AVL 5406.

.................................................................................................................................................. 46

Table 5 - Fuel composition in the percentage of total volume used for study the effect of air-

fuel ratio in syngas combustion process. .................................................................................. 47

Table 6 - Properties of basic components present in the fuels used for study the effect of air-

fuel ratio in syngas combustion process. .................................................................................. 47

Table 7 – Main properties of the fuels used for study the effect of air-fuel ratio in syngas

combustion process. ................................................................................................................. 48

Table 8 – Operating conditions set for study the effect of air-fuel ratio from stoichiometric to

lean conditions in syngas combustion process. Engine speed: 900 RPM; SA: 7 CAD BTDC;

Load: WOT. .............................................................................................................................. 64

Table 9 - Camera set up for imaging acquisition set for study the effect of air-fuel ratio from

stoichiometric to lean conditions in syngas combustion process. F-number: 32, Exposure Time:

0.5 CAD .................................................................................................................................... 74

Table 10 – Operating conditions set to study the effect of relative air-fuel ratio from lean to

extremely lean conditions in syngas combustion process. Engine speed: 900 RPM; SA: 13 CAD

BTDC; Load: WOT. ................................................................................................................. 84


xviii

Table 11 – Main thermodynamic parameters extracted from pressure traces for lean to

extremely lean conditions at fixed spark timing. Engine speed: 900 RPM; SA: 13 CAD BTDC;

Load: WOT. .............................................................................................................................. 86

Table 12 – Crank angle durations from the start of spark to fixed mass fraction burned

thresholds calculated from pressure traces for lean to extremely lean conditions at fixed spark

timing. Engine speed: 900 RPM; SA: 13 CAD BTDC; Load: WOT. ...................................... 87

Table 13 - Camera set up for imaging acquisition set for study the effect of air-fuel ratio from

lean to extremely lean conditions in syngas combustion process at fixed spark timing. F-

number: 5.6, Intensification: 55000.......................................................................................... 90

Table 14 – Operating conditions set for study the effect of spark timing in stoichiometric air-

fuel ratio fueled with S50. Engine speed: 900 RPM; λ=1.0; Load: WOT; DOI= 204 CAD.. 101

Table 15 – Operating conditions set for study the effect of spark timing from lean to extremely

lean conditions in syngas combustion process. Engine speed: 900 RPM; Load: WOT. ........ 111

Table 16 – Maximum flame speed comparison between fixed spark advanced and MBT

condition for M, M25, M50, S50, and S75 at λ=1.4. ............................................................. 121

Table 17 - Maximum HCF and distances from the center of the combustion chamber

comparison between fixed spark advanced and MBT condition for M, M25, M50, S50, and S75

at λ=1.4. .................................................................................................................................. 124

Table 18 – Fuel composition in the percentage of total volume used for study the effect of

diluent content in pure syngas combustion process. .............................................................. 128

Table 19 – Main properties of the fuels used for study the effect of diluent content in pure

syngas combustion process. .................................................................................................... 129

Table 20 – Operating conditions used for study the effect of diluent content in pure syngas

combustion process. Engine speed: 900 RPM; SA: 7 CAD BTDC; Load: WOT.................. 130
xix

Table 21 – Crank angle durations from the start of spark to fixed mass fraction burned

thresholds calculated from pressure traces for study the effect of diluent content in pure syngas

combustion process. Engine speed: 900 RPM; SA: 7 CAD BTDC; Load: WOT.................. 135

Table 22 – Exhaust gas emissions measured at engine speed: 900 RPM; SA: 7 CAD BTDC;

Load: WOT, and λ=1.4. .......................................................................................................... 137

Table 23 – Camera set up for study the effect of diluent content in pure syngas mixtures at fixed

spark timing (7 CAD BTDC). Exposure Time: 0.5 CAD; Intensification: 55000. ................ 137

Table 24 - Comparison between Laminar Flame Speed (LFS) and Engine like Condition Flame

Speed (EFS) for the five fuels study in stoichiometric condition. Engine speed: 900 RPM; SA:

7 CAD BTDC; Load: WOT.................................................................................................... 143

Table 25 – Constant parameters used for the calculus of laminar flame speed of methane and

S50. ......................................................................................................................................... 154

Table 26 – Complementary specifications of the set up used in the PFI-SI single-cylinder

research engine AVL 5406. .................................................................................................... 156


xx

List of Symbols

λ relative air-fuel ratio (AFRrel)

AFR air-fuel ratio

ATDC after top dead center

ASOS after the start of spark

BFBG bubbling fluidized bed gasifier

BSFC brake specific fuel consumption

BTDC before the top dead center

CAD crank angle degrees

CCD charge-coupled device

CFBG circulating fluidized beds

CH4 methane

CI compression ignition

CMOS complementary metal-oxide semiconductor

CNG compress natural gas

CO carbon monoxide

CO2 carbon dioxide

Cov coefficient of variation

CR compression ratio

DFBG downdraft fixed bed gasifier

DI direct injection

DISI direct injection spark ignition

DOD degree of diluents

DOI duration of injection

EOI end of injection


xxi

FHWM full width at half maximum

Fps frame per second

H2 hydrogen

HCCI homogeneous charge compression ignition

HCF Heywood circularity factor

ICCD Intensified charge-coupled device

ICE internal combustion engine

IMEP indicated mean effective pressure

IR infrared region

LHV Lower heating value

M pure methane

M25 fuel blend containing 75%vol methane and 25%vol hydrogen

M50 fuel blend containing 50%vol methane and 50%vol hydrogen

MBT maximum brake torque

mf fuel mass

MFB mass fraction burned

N Crankshaft rotational speed

N2 nitrogen

NG natural gas

NOx nitrogen oxides

Nu nulsset number

O2 oxygen

PDF probability density function

PFI port-fuel injection

Pmax maximum in-cylinder pressure

Pow power

Pr prandtl number
xxii

Re reynolds number

RPM revolutions per minute

S50 syngas blend containing 50%vol carbon monoxide and 50%vol


hydrogen

S5050 syngas blend containing 50%vol carbon monoxide and 50%vol


hydrogen in fuel basis and 50% of diluent.

S75 syngas blend containing 25%vol carbon monoxide and 75%vol


hydrogen

S7550 syngas blend containing 25%vol carbon monoxide and 75%vol


hydrogen in fuel basis and 50% of diluent.

SA spark advance (spark timing)

SI spark ignition

SOx sulphur oxides

TDC top dead center with A for after and B for before

UV ultra-violet

UFBG updraft fixed bed gasifier

UV ultraviolet region

Vd displacement volume

WDD waddel disk diameter

WOT wide open throttle


xxiii

Contents

1 INTRODUCTION .......................................................................................................... 25

2 FUNDAMENTAL CONCEPTS .................................................................................... 29


2.1 Gasification Process ................................................................................................ 29
2.2 Syngas in SI engines ........................................................................................................ 34
2.3 Combustion Analysis in Engines with Optical Access ........................................ 38

3 EFFECT OF AIR-FUEL RATIO IN SYNGAS COMBUSTION PROCESS .......... 42


3.1 Experimental apparatus and measurement procedure ...................................... 44
3.1.1 Engine test bed ......................................................................................................... 44
3.1.2 Fuels ......................................................................................................................... 46
3.1.3 Test Procedure .......................................................................................................... 49
3.1.4 Thermodynamic analysis .......................................................................................... 51
3.1.5 Exhaust measurements ............................................................................................. 54
3.1.6 Optical Investigations ............................................................................................... 55
3.2 Stoichiometric to lean conditions .......................................................................... 63
3.2.1 Operating Points ....................................................................................................... 63
3.2.2 Thermodynamic Results ........................................................................................... 64
3.2.3 Optical investigations ............................................................................................... 73
3.3 Extreme lean conditions ......................................................................................... 83
3.3.1 Operating Points ....................................................................................................... 83
3.3.2 Thermodynamic Results ........................................................................................... 84
3.3.3 Optical Investigations ............................................................................................... 90
3.4 Summary ................................................................................................................. 96

4 EFFECT OF SPARK TIMING IN SYNGAS COMBUSTION PROCESS .............. 98


4.1 Stoichiometric condition ........................................................................................ 99
4.1.1 Operating Points ....................................................................................................... 99
4.1.2 Thermodynamic Results ......................................................................................... 100
4.1.3 Optical Investigation .............................................................................................. 105
4.2 Extreme lean conditions ....................................................................................... 110
4.2.1 Operating points ..................................................................................................... 110
4.2.2 Thermodynamic Results ......................................................................................... 112
xxiv

4.2.3 Optical Investigation .............................................................................................. 119


4.3 Summary ............................................................................................................... 125

5 EFFECT OF DILUENT CONTENT IN SYNGAS COMBUSTION PROCESS ... 127


5.1 Operating Points ................................................................................................... 128
5.2 Results .................................................................................................................... 130
5.2.1 Thermodynamic results .......................................................................................... 130
5.2.2 Optical Investigation .............................................................................................. 137
5.3 Summary ............................................................................................................... 146

6 QUASI-DIMENSIONAL MODELLING IN AN OPTICALLY ACCESSIBLE SI


ENGINE WITH SYNGAS ................................................................................................... 147
6.1 Simulation Model.................................................................................................. 150
6.2 Validation .............................................................................................................. 156
6.3 Results .................................................................................................................... 158
6.4 Summary ............................................................................................................... 161

7 CONCLUSIONS ........................................................................................................... 162

8 BIBLIOGRAPHY ......................................................................................................... 165

9 APPENDIX ................................................................................................................... 174


9.1 Appendix A............................................................................................................ 174
9.1.1 Image processing tests ............................................................................................... 174
9.1.2 Camera set up tests .................................................................................................... 177
9.2 Appendix B ............................................................................................................ 181
9.2.1 Gas flow rates model ................................................................................................. 181
9.2.2 Heat Transfer model .................................................................................................. 182
9.2.3 Friction model ........................................................................................................... 183
25

1 Introduction

Depleting fossil fuel reserves and the increase in climate changes are the main reasons
to seek alternative methods for more efficient use of energy and wider application of renewable
resources. New policies are addressed to reduce the greenhouse gases emissions and increase
the use of sustainable fuels from renewable sources, in both transportation and power generation
sectors [1]. The concept of sustainability lies in the ability to satisfy the necessity of the present
without compromising the resources for the future [1]. In this way, biomass is the 4th most
abundant energy source after coal, oil and natural gas throughout the world, contributing with
almost 14% of the total energy demand globally [2]. Only about 40% of potential biomass
energy is currently utilized [3]. In the United States was estimated that without any changes in
land use and without interfering with the production of food grains, 1.3 billion tons of biomass
could be converted each year on a sustainable basis for biofuel production. To put in context,
1.3 billion tons of biomass is equivalent to 3.8 billion barrels of oil in energy content [4]. US
equivalent energy consumption is about 7 billion barrels per year [5]. Only in Asia does the
current biomass usage slightly exceed the sustainable biomass potential.
Biomass combines solar energy and carbon dioxide into chemical energy in the form of
carbohydrates via photosynthesis. Its use as a fuel is a carbon neutral process since the carbon
dioxide captured during photosynthesis is released during a complete combustion [3]. Biomass
includes wood, crops, agricultural and forestry residues, byproducts from the processing of
biological materials, organic parts, and sludge of municipal wastes. Biomass-derived fuels of
different forms can replace the petrol-fuels in the energy generation and transportation sectors
[1]. For example, ethanol from corn and sugarcane, and biodiesel from soy, rapeseed, and oil
palm have been in the international market as biofuels for quite some time now. However, these
biomass feedstocks are edible and their abundant use gives rise to the food versus fuel crisis.
The present challenge is to produce good quality fuels for engine applications from non-edible
biomass resources like agricultural wastes, municipal solid wastes among others [6].
Agricultural wastes represent an attractive potential feedstock for sustainable energy
production, having the advantage of low cost, easy access, and greenhouse neutrality. It is
considered as “CO2 neutral” because emissions of sulfur dioxides and nitrogen oxides are very
small. Thus, the use of agricultural waste is a real option for clean fuel with zero emissions [7].
Despite the high production of biomass in industrial plants and its energy potential, most of
26

them produce electricity only for their own consumption. On the other hand, because an excess
of biomass is normally generated during biomass processing, boilers in the industry are
basically operated to discard excess of agricultural waste, as there must be a balance between
the production and the use of the same.
On the other hand, a significant increase in the amount of waste from cities has become
a globally environmental problem. The statistics warned that global solid waste generation was
on pace to increase 70% by 2025. Moreover, the global cost of dealing with all that trash is
rising too: from $205 billion a year in 2010 to $375 billion by 2025, with the sharpest cost
increases in developing countries [8]. Thus, proper strategy for waste management is inevitable
as it is a vital part of sustainability and environmental protection. Therefore, new solutions to
convert agricultural and municipal waste in energy are necessary.
Thermochemical conversion technologies include combustion, gasification, and
pyrolysis. While combustion of biomass is the most direct and technically easiest process, the
overall efficiency of generating heat from biomass energy is low [3]. Gasification has many
advantages over direct combustion, due to it can use low-value feedstock and convert them not
only into electricity but also into transportation fuels [8]. The gas produced in the gasification
process of biomass is generally called syngas (also synthetic gas or production gas). The major
components of syngas are hydrogen (H2), carbon monoxide (CO), methane (CH4) as fuels and
nitrogen (N2) and carbon dioxide (CO2) as dilution components of the mixture. In this work, a
convention was adopted in which pure syngas is the one that only contains fuels components
(CH4, CO, and H2) in its composition.
Currently, small-scale electricity generation using biomass gasification is increasingly
attracting interest as a prospective way to provide remote districts with electrical power using
local renewable fuels [1]. An additional benefit in such a rural electrification mechanism is the
possibility of the utilization of various organic wastes from the local industry and agriculture
with a considerable CO2 emission reduction. With negative environmental and social effects
caused by a rapid depletion of resources of natural gas and crude oil, research and development
projects on electricity generation with biomass gasification have gained a new momentum [9].
In this way, different researchers are working on the problem of coupling a biomass gasifier to
internal combustion engine from an experimental point of view [10, 11]. Other researchers use
computational techniques to estimate the influence of the gas composition on the engine
performance [12, 13]. The present work is aimed to investigate the fundamental characteristics
of combustion process of a spark ignition engines operating with various blends of syngas and
several operative conditions.
27

To fully investigate SI internal combustion engine operating regimes, experimental and


numerical investigations were carried out to focus on:

• study the effect of different H2/CO ratios in the syngas blends (pure syngas);
• compare the behavior of the combustion process with different air-fuel ratios (λ),
from stoichiometric to extreme lean conditions;
• analyze the effect of spark timing in the development of the flame front;
• quantify the effect of fuel dilution content (syngas) in relation to pure syngas during
the combustion process;
• apply a numerical code to simulate the first stage of the combustion process of pure
syngas in an optical research engine.

After a technical review of the fundamental concepts (chapter 2), these five topics will
be developed in subsequent chapters (chapter 3 to chapter 6). Combustion process was studied
in detail through combined methodologies based on thermodynamic analysis, exhaust
measurements and optical diagnostics. Specifically, cycle resolved UV-visible digital imaging
was applied to follow the flame front propagation. A high-speed CMOS camera coupled with
an intensifier was used for this study. Image processing was applied to evaluate flame speed
and other morphology parameters, including flame distortion and centroid motion.
Chapter 3 and 4 studies the combustion behavior of two equivalent pure syngas with
CO/H2 50-50% (S50) and CO/H2 75-25% (S75) against methane (M). Also, methane-hydrogen
blends with CH4/H2 75-25% (M25) and CH4/H2 50-50% (M50) was studied. For these chapters,
different air-fuel ratios (air dilution) and spark ignition timing (SA) was tested. The pure syngas
compositions were taken to be representative of the family of various ratios of CO/H2 [14]. In
this work the notation X a% always refers to the percentage (a) of the chemical component (X)
per volume in the total fuel volume. In the case that the mixture is in percentage of the total fuel
mass will be clarified (%mass).
On the other hand, the focus of chapter 5 is to quantify the effect of inert addition in
pure syngas fuel (fuel dilution). For this purpose, 50% of CO2+N2 added to S50 and S75 as the
more representative fuel of real gasification process (S5050 and S7525). In chapter 6, a quasi-
dimensional numerical simulation code was applied to pure syngas with 50% H2 (S50).
Validation process with thermodynamic data and comparison between simulation data against
optical measurements was performed. Finally, chapter 7 presents the conclusions of this work.
28

In all cases, the engine was operated at a fixed crankshaft rotational speed (RPM) and
constant open throttle. Usually, stationary piston engines that power electric generators run at
750, 1500 or 3000 RPM and maximum load [9]. However, 900 RPM (low range) was selected
to meet the engine safety requirements. The maximum load condition was reached with the
wide-open throttle (WOT). Therefore, the condition used is typical of electrical production
engines fueled with gas.
As a resume, the scope of this manuscript is to investigate the use of syngas in an SI engine
as a potential replacement for fossil-derived fuels. This work is part of an ongoing research
project to study the combustion properties for different syngas composition. It is important to
understand how various syngas compositions affect SI combustion fundamentals due to its high
variability of the production process. The results from this master thesis shed light into the
fundamental aspects of syngas combustion and provide a foundation for future gasification
plant designers and syngas producers. Moreover, the work can contribute to calibrate simulation
codes and further to the development of numerical models.
29

2 Fundamental Concepts

2.1 Gasification Process

Direct combustion of solid biomass in engines is not practical and so it requires its
conversion to liquid or gaseous forms. Many studies have proved the superiority of gasification
over other biomass conversion processes [3]. Gasification is a process that converts solid
biomass into a combustible gas. Biomass-derived syngas produced via gasification is
considered as a promising option for replacing the use of conventional fossil fuels in the power
generation sector [1]. As was mentioned in the Introduction, the major components of syngas
are hydrogen (H2), carbon monoxide (CO), methane (CH4) as fuels and nitrogen (N2) and
carbon dioxide (CO2) as dilution components of the mixture. Its composition depends on
biomass source, the process of gasification and oxidizer used. The problem with having a
variable feedstock and variable production methods is that the syngas composition can vary
significantly [15]. With that, properties as heating value can also have high variations. This is
one of the limitations of using syngas in piston engines.
Biomass gasification differs from coal gasification because is a carbon-neutral and
sustainable energy source. Because biomass is more reactive and has higher volatiles content
than coal, gasification process occurs at a lower temperature [14]. Lower temperature reduces
the extent of heat loss, emissions and material problems associated with high temperatures.
Biomass also has low sulfur content, which results in lower SOx emission [15]. In Figure 1, a
diagram of the gasification process is depicted. The first phase is upstream processing, which
includes processing of the biomass to make it suitable for gasification operations. Size reduction
is needed to obtain appropriate particle sizes. In addition, drying is desirable to achieve
appropriate moisture so the process can work efficiently.
Gasification is the central step, in which the main operating parameters of the gasifier
include type and design of gasifier, gasification temperature, flow rates of biomass and
oxidizing agents (air or steam), type and amount of catalysts [3]. In this process, a limited
amount of oxygen/air is supplied to biomass placed in a reactor in such a way that the air-fuel
ratio (AFR) is below the stoichiometric one (λ < 1.0). This result in burning of a relatively small
part of biomass that generates heat to maintain a series of thermochemical processes with a
30

mixture of gases being generated as a final product called syngas [3]. Four key processes occur
inside the reactor, namely: drying, pyrolysis, oxidation and reduction, and each of these
processes has certain physical and chemical features [1]. For more information about each
specific process, see Susastriawan et al. review [8].

Figure 1 - Processes involved in biomass gasification.

A variety of biomass gasifier types has been developed. They can be grouped into two
major classifications: fixed-bed and fluidized-bed. Differentiation is based on the means of
supporting the biomass in the reactor vessel, the direction of flow of both the biomass and
oxidant and the way heat is supplied to the reactor [2]. All these types have certain benefits and
drawbacks with respect to fuel type, application, and operation.
Fixed bed gasifier can be classified further as updraft (countercurrent) or downdraft
(concurrent). In the updraft gasifier (UFBG), the feed (biomass) is introduced from the top and
moves downwards while gasifying agents (air, steam, etc.) are introduced at the bottom of the
grate, so the product gas moves upwards. In this case, the combustion takes place at the bottom
of the bed, which is the hottest part of the gasifier, and product gas exits from the top at a lower
temperature (around 500°C) [2]. Because of the lower exit temperature, the product gas contains
large amounts of tar. Tar is a generic term used for all organic compounds found in the product
gas with the exception of gaseous hydrocarbons. It is the part of the biomass, which does not
decompose completely into lighter gases. In a downdraft gasifier (DFBG), the feed and the
product gas moves downward and the syngas exits from the bottom at a higher temperature,
around 800°C [2]. In this case, most of the tars are consumed. However, heat needs to be
recovered from the high-temperature syngas to increase the energy efficiency.
In the fluidized bed gasifier, the feed is introduced at the bottom, which is fluidized
using air, nitrogen and/or steam and the syngas then moves upward. There are more particulates
in the syngas from this type of gasifier [16]. However, fluidization of the bed enhances the heat
transfer to the biomass particle leading to increases in reaction rates and conversion efficiencies.
31

Also, fluidized beds are able to tolerate a wide variation in fuel types [3]. A fluidized bed can
be either a bubbling fluidized bed (BFBG) or a circulating fluidized bed (CFBG).
In order to choose the type of gasifier that best suits to the industrial plant and the source
of the biomass, it is strongly recommended to study the final application first. Syngas can be
utilized in various combustion technologies, such as power plant boilers, burners, gas turbine
combustors and internal combustion engines. For the latter aforementioned application, it is
required that the syngas have to be cooled down first to not compromise the engine volumetric
efficiency [9]. Available literature includes reports on both, spark ignition (SI) [11, 18] and
compression ignition (CI) engines [19, 20] being charged with the syngas. However, those are
mostly pilot or research set-ups, with the expectation of a few commercial installations. The
syngas is a poor compression ignition engine fuel and its application in CI engine requires that
a pilot injection of high cetane number fuel is done, making this a dual fuel CI engine, with the
potential to use a biodiesel fuel for the pilot injection [9].
For small and medium thermal and electrical power as SI applications, fixed-bed
gasifiers (Figure 2) shows strong advantages [21]. In addition, the downdraft configuration
reduces the amount of high molecular weight hydrocarbons and particles as compared to the
updraft configuration. For these reasons, the downdraft layout is the most satisfactory when the
gasifier must be coupled with a reciprocating internal combustion engine [9]. For small-scale
applications, biomass gasification in downdraft reactors has been studied extensively and
currently is considered to be a mature technology [22-24].
The operating range of the λ in the gasifier modifies the composition of the produced
gas and then its properties such as the energy content (heating value) among others (density,
air-fuel ratio, combustion speed). Generally, the heating value of gasification gas should be
maximized. However, the rest of the properties of the gas are also important, as the laminar
flame speed or the tendency to auto-ignition [21]. Table 1 depicts different ranges of
composition and respective properties for the different type of gasifiers when the same fuel and
oxidizer is used (dry wood and air, respectively) [2]. It is observable that high variation in the
composition can be produced depends on each system and operative condition.
In respect of the influence of the oxidizer, Table 2 depicts several syngas compositions
for air, O2 and steam as an oxidizer in three types of gasifiers [24-28]. It is well known that the
heating value and the hydrogen content of syngas are higher when gasification is made with
steam than air [17]. In addition, it can be extracted from Table 2 the decrease in diluent content
when steam or O2 is used.
32

Figure 2 – Scheme of a Downdraft Fixed-Bed Gasifier, adapted from [21].

Table 1 - Characteristics of the syngas for different types of gasifiers with dry wood as
biomass feedstock and air as oxidizer [2].
Property DFBG UFBG BFBG CFBG
H2 [%] 15-21 10-14 5-16 15-22
CO [%] 10-22 15-20 10-22 13-15
CH4 [%] 1-5 2-3 2-6 2-4
CO2 [%] 11-13 8-10 9-19 13-15
N2 [%] 40-60 40-60 40-60 40-60
LHV [MJ/Nm3] 4.0-5.6 3.7-5.1 3.7-8.4 3.6-5.9
Tar [mg/Nm3] 10-6000 10000-150000 Nd 2000-30000
Particles [mg/Nm3] 100-8000 10-3000 Nd 8000-100000
33

Table 2 - Effect of gasify agent on syngas composition for different types of gasifiers with dry
wood [2].

Property Air O2 Air Steam Air Steam


DFBG DFBG BFBG BFBG CFBG CFBG
H2 [%] 17 32 9 52 14 34
CO [%] 21 48 14 23 19 27
CH4 [%] 1 2 7 7 4 11
CO2 [%] 13 15 20 18 15 23
N2 [%] 48 3 50 0 48 5
LHV [MJ/kg] 4.3 10.5 4.2 15.0 4.6 11.5

Summarizing, it is possible to characterize syngas as fuel content in the range of


15- 50% CO, 10-50%l H2 and 2-5% CH4. The remaining composition is inert gases (N2 and
CO2) with a wide range of 0-70%; which depends on the oxidizer used [2]. Regarding the
influence of various parameters involved in the process of gasification in the final features of
syngas, there is some discrepancy in the values given by various authors. This highlights the
strong dependence on the final composition of the syngas depending on the biomass used, the
type of gasifier and the conditions of pressure and temperature.
The range of heating value for syngas is defined as low (4–6 MJ/Nm3), medium (12–18
MJ/Nm3) and high (40 MJ/Nm3) values [1]. A gas with high purity (low quantities of inert) is
extremely beneficial for fuels and chemical synthesis since it substantially reduces the size and
cost of downstream equipment. In addition, for transportation application could improve the
energy density. Therefore, the type of gasifier and related parameters must be chosen depending
on the final use.
Before the product gas from biomass gasification could be used in ICEs, it needs to be
processed (downstream processing, see Figure 1). The product gas contains particulates, tar,
alkali compounds, and sulfur-containing compounds, which typically need to be removed
before the product gas is used [3]. Tolerable amounts of the contaminants in the syngas depends
on the application. The critical issue of the syngas in ICEs is the removal of undesirable solid
particles and tars. For particle removal, cyclone separators and filters are widely and routinely
used for reducing the engine wear. Moreover, condensation of tars can result in deposits on
engine's components (intake manifold, valves and piston crowns and cylinder walls) and in
effect de-rating the engine operation and performance [2]. According to Insig et al. [29], the
tolerance limit of tar contents for the engine is between 50-100 g/Nm3. Even if no general and
34

comprehensive survey is available, the technical and economic problems related to tars have
caused a cancellation of several investments in the past. Removal of tar is one of the biggest
technical challenges facing the commercialization of gasification technology [30]. As was
describe above, the main advantage of DFBG is the lower tar concentration in the syngas, which
is very important for the durability of ICEs.
The challenges with gasification are to understand the effects of operating conditions on
gasification reactions for reliably predicting and optimizing the product compositions, and for
obtaining maximal efficiencies. It was estimated that with the current technology, the cost of
electricity from gasification of locally available biomass will be lower than the cost of
electricity from a diesel engine [31]. However, the major challenges in the commercialization
of small-scale gasification systems are to effectively clean the product gas, design the system
flexible enough to operate with ICEs using varying biomass qualities, and to decrease the
capital cost [32].

2.2 Syngas in SI engines

The internal combustion engine (ICE) has had a significant impact on the world over
the last century [33]. Even though it was invented over 100 years ago, there is still a vast amount
of research that is devoted to improving ICE operation. The purpose of an ICE, however,
remains the same; to convert chemical fuel energy into useable shaft power [14]. The main
mode of this conversion is currently accomplished using either four-stroke spark ignition (SI)
or compression ignition (CI) engines. An excellent review of conventional SI and CI engine
operation can be found in various textbooks [34-36].
At was mentioned, it is well accepted that due to the high auto-ignition temperature of
the components gases, syngas alone is not a suitable fuel in compression ignition engines.
Therefore, such fuels need to be operated in spark ignition mode [1]. Methane fueling in SI
engines was widely investigated, not only in terms of performance and exhaust emissions
[37- 40] but also the fundamental characterization of combustion process [41-44]. On the other
hand, few works are focused on the combustion process investigation when syngas fuels are
used [45-53], and practically none by optical diagnostics in ‘real-world’ engine-relevant
conditions.
Deeper insight into the effects of different fuel compositions, the relative air-fuel ratio
(λ) and spark ignition timing (SA) on in-cylinder processes can be useful for the development
35

of advanced solutions for the optimization of SI engines with syngas as fuel. Syngas is
characterized by high variability in composition and fundamental properties depending on the
process conditions and feedstock. For that reason, its use may pose some difficulties so it
requires new fundamental understanding for developing the next generation of SI engines.
Development of gas engines using syngas has been explored ever since World War II.
It is estimated that over seven million vehicles in Europe, Australia, South America and Pacific
Islands were converted to run on syngas [53]. These engines were spark-ignited engines, mostly
in the lower compression ratio bracket operating either on charcoal or biomass-derived gas. As
ready-made syngas engines are not available in the market, two options are available for
modifying existing engines to run on 100% syngas [53]. Some researchers choose to modify a
diesel engine [53] while others prefer to use the already available natural gas engine (CNG)
[47-50].
In addition, the question of power generation using syngas has been addressed in recent
times by a few researchers [45-47] to understand how different gas compositions affect SI
combustion fundamentals, such as performance, stability, knock and burn duration among
others. Improvements in gasification technologies, as better clean gas methods, and improve
syngas composition (increase LHV and H2 content and decrease of diluents content); make that
more funds and efforts were dedicated to the study of syngas in ICEs. Table 3 shows the most
important works published in the last 20 years with SI engines applying syngas as fuel. Is
remarkable that the gas composition used by different authors vary significantly. Each of the
work cited in Table 3 will be described and linked with the present work in the subsequent
chapters.
36

Table 3 – List of important studies in SI engines with syngas in recent years.

Crankshaft
Year
rotational
CO H2 CH4 CO2 N2 Type of of
Researcher velocity Ref
[%] [%] [%] [%] [%] gasifier the
tested
work
[RPM]
DFBG, oxy-
Shivqpji and
18 18 1 11 52 steam 1500 2017 [45]
Dasappa
gasification
1500, 1800,
Hagos et al. 40 40 20 0 0 - 2017 [46]
2100, 2400
19 18 3 12 48 DFBG, air,
Przybyła et
28 11 2 9 52 sewage 1500 2016 [10]
al.
23 5 3 11 59 sludge
Gobato et
20 18 1 12 49 DFBG, Wood 1500 2015 [18]
al.
23 23 26 28 0 Catalytic
Arroyo et al. decomposition 2500 2015 [47]
39 40 11 10 0 from biogas
18 18 1 11 52
12 13 2 11 63 DFBG, oxy-
Shivqpji and
steam 1500 2015 [48]
Dasappa 14 26 3 19 38 gasification
16 37 4 25 18

Hagos et al. 50 50 0 0 0 - 1500-2400 2014 [49]

Hagos et al. 50 50 0 0 0 - 1500-2400 2014 [49]

18 18 1 11 52
12 13 2 11 63 DFBG, oxy-
Shivqpji and
steam 1500 2014 [11]
Dasappa 14 26 3 19 38 gasification
16 37 4 25 18
Olive kernels,
23 17 3 5 52
BFB
Tsiakmakis peach
20 12 11 5 52 2800 2014 [50]
et al. kernels, BFB
grape kernels,
17 19 2 6 56
BFB
23 23 26 28 0 Catalytic
Arroyo et al. decomposition 2500-4000 2014 [51]
39 40 11 10 0 from biogas
Raman and DFBG, air,
21 23 1 9 46 1500 2013 [12]
Ram wood
Centeno et DFBG, air,
al.
20 17 1 10 52
wood
500-3000 2012 [13]
25 75 0 0 0
Bika et al. - 900 2011 [14]
50 50 0 0 0
Mustafi et
52 44 0 4 0 - 1000-2000 2006 [52]
al.
Sridhar et DFBG, air,
19 19 2 12 48 1500 2001 [53]
al. wood
37

Another point taken from the Table 3 is that generally, the research groups perform two
types of analysis. The first one is with pure syngas (H2/CO blends). Understanding how varying
H2/CO ratio (syngas fuel component) can affect the combustion fundamentals in an internal
combustion engine (ICE) is a critical first step in characterizing a syngas fuel [49]. This family
of syngas was under investigation for its laminar flame velocity and for its combustion and
knock characteristics in SI engine [14]. There have been several investigations of pure syngas
supplemented in spark ignition engines [14, 33, 39] and even in HCCI [54] applications. Bika
et al. [14] studied various ratios of H2/CO syngas in a port-injection SI engine for their
combustion characteristics and knock limit. The fuels investigated included pure H2, 75-25%
H2/CO, and 50-50% H2/CO. In the study λ vary between 1.25-1.70 and compression ratios (CR)
of 6:1 to 10:1. It was reported that an increase in the content of CO increased the knock limit
of the syngas. The study also indicated that the increase in the H2 content advanced the ignition
timing of the maximum brake torque (MBT), therefore decrease the combustion duration.
Hagos et al. [49] study performance and emissions of H2/CO (50-50%) mixture in the
lean range in a direct injection spark ignition (DISI) engine and compared with compress natural
gas (CNG). The engine speed ranging from 1500 to 2400 RPM, with the throttle being held in
the wide-open position (WOT) and λ were vary from 1.5 to 3.0. The results indicated that a
wider relative air-fuel ratio is possible with syngas with a very low CovIMEP when spark
advance for MBT condition was set. In addition, the thermodynamic analysis showed that
syngas produced a higher in-cylinder peak pressure and faster combustion than for CNG.
However, natural gas produced a higher brake thermal efficiency and lower brake specific fuel
consumption (BSFC). For the syngas, the total hydrocarbon emission was negligible at all load
conditions, and the carbon monoxide emission was negligible at higher loads and increased
under lower load conditions. However, the emission of nitrogen oxides was higher at higher
loads with H2/CO blends.
Other works were dedicated to studying real syngas mixtures [52], to understand their
applicability in real engine operation, observing the engine performance, the global efficiency,
combustion stability, and durability. In general, the engine was coupled with real gasifier
operation. In general, the main advantage of the use of syngas in SI engines is the presence of
high quantities of hydrogen. It can be interpreted as an excellent additive; increasing the laminar
flame speed until seven times higher than the methane speed [43]. Moreover, hydrogen is
characterized by a more stable combustion due to a wider flammability range and low ignition
energy, thus allowing the extension of the lean operation limit [44]. Other interesting point
found in the works (listed in Table 3), is the increase in power de-rating when fueling modern
38

spark ignition engines with syngas than in adapted diesel engines. This shows the effect of
compression ratio, generally 8 to 12 for SI than 17 to 22 for CI engines [53].
The main drawback of syngas is its lower heating value compared with natural gas,
caused by the high content of inert gases in its composition, which causes that part of the energy
released in combustion was absorbed by these components instead of generating power [14]. In
addition, the presence of inert gases displaces air in the intake manifold when PFI mode of
injection is used, thus the volumetric efficiency decreases, resulting in the reduction of the
power output [14]. Flame stability is another problem encountered in the combustion of syngas
with high quantities of diluents, due to low propagation speed [48].
In order to avoid the effects of its low heating value, syngas can be upgraded by
removing fractions of inert gases, obtaining gaseous fuels (pure syngas) with properties closer
to natural gas, but with renewable origin [11]. These upgrading methods improved gaseous
fuels for use in ICEs. Therefore, syngas has considerable potential as an alternative fuel for
internal combustion engines [46]. More fundamentals studies are needed to continue the
development of this fuels for ICEs applications.

2.3 Combustion Analysis in Engines with Optical Access

Gaseous fuels are generally studied by employing methods mainly based on in-cylinder
pressure measurements and exhaust analysis; nonetheless, complementary techniques such as
optical investigations can provide valuable information on combustion evolution [55]. The
product development phase, where the request for a fast response and precise diagnostic
information is especially high, sensor technology, data acquisition and interpretation techniques
can provide a substantial contribution to innovative and development products.
In general, research engines with optical access is either maximized to allow application
of complex optical diagnostic techniques or is maintained full engine operability whilst optical
access are tailored to the diagnostic demand. Therefore, the modifications necessary for the
application of optical techniques imply numerous compromises to the original behavior of a
commercial engine [56]. Nevertheless, even if an ever-growing field of operating conditions
can be covered with single cylinder optical engines, it imposes limits on the comparability with
standard multi-cylinder engines. Special geometry with relatively high crevices volume was
found a limitation of this type of research setup; which consists of the top-land region between
39

the piston, first ring [55]. The increase of blow-by rates also represents an issue that can have a
significant impact on the analysis performed on such engines.
Moreover, low compression ratio (CR) is another disadvantage of this type of engine,
due to the limitation of maximum in-cylinder pressure and temperatures [57]. Therefore, the
CR used in this type of application range from 8:1 to 10:1 (commercial engines ranges from
10:1 to 12:1). In addition, this parameter is difficult to determine with usual methods applied to
commercial engines because it contains large crevices. Irimescu et al. [57] proposed a method
for estimating compression ratio based on in-cylinder pressure traces, which has a higher
precision for this type of engine than conventional measurements methods.
Despite these difficulties, interesting information can be taken from optical analysis in
research engines, like those reported in the literature [58-76]. Optics-based diagnostic
techniques are well established in the analysis of combustion in the stationary regimen, and
stable conditions of pressure and temperature. The challenge in applying them to combustion
processes in ICEs is the adaptation of the diagnostic technique to the operation of the engine.
The requirement of maximum optical access for the illumination of the combustion chamber
and the inspection of the illuminated area is met with large optical windows forming parts of
the cylinder head, the cylinder liner or the piston [56]. Generally, studies deal with topics as
flow field [73], fuel injection and mixture formation [64], as well as the ignition of the mixture
and combustion development [65]. In addition, the characterization of the later combustion
phase with fuel deposit is the object of study of several researchers [61].
A deeper understanding of flame front characteristics can be achieved with laser-based
techniques [44], which allow the identification of variations between different fuel types with
regard to local length scales. On the other hand, spectroscopic investigations offer a glimpse
into the evolution of multiple chemical species at the same instance during combustion [74]
[75], which can be used for measuring local air-fuel ratio. Other technique, which is the focus
of the present work, is direct flame visualization. This is the most widely used technique [58],
in which cycle-resolved imaging being able to provide insight into ignition process as well as
flame front propagation and late combustion phase.
Combustion in pent roof-head geometries (commercial cylinder head geometry of
modern engines) has been investigated by direct imaging using Intensified Charge Couple
Device (ICCD) [41] and high-speed CMOS cameras [69]. In general, the spectral range of the
CMOS camera extended from 380 nm to 900 nm with the highest quantum efficiency at 550
nm. For ICCD, the spectral range of the camera is extended from 200 nm to 900 nm with a quite
flat quantum efficiency between 240 and 400 nm [76]. In particular, chemical species as OH
40

radicals (characteristic wavelength at 310 nm) that represents a consolidated marker of both
reaction and burned gas region were detectable only by ICCD [77]. On the other hand, the
CMOS camera not only ensured cycle resolved visualization but also presented higher
sensitivity to visible and near infra-red (IR) emissions [78]. This allowed the detection of
excited CO2 in burned gas and CH radicals (characteristic wavelength at 431 nm) in the reaction
zone [76].
A major aspect of understanding combustion of fuels in SI engines is their laminar and
turbulent burning velocities in controlled environments [79, 80] and at engine-relevant
conditions [81]. Direct visualization of the flame from the piston window provided data that
could satisfy the requirements and could be used for direct comparison (combustion vessel
experiments and engine application) because of similar methods of flame radius growth
quantification [82]. Although the view from the bottom of the combustion chamber allows only
a line-of-sight evaluation and ‘projected burnt area’ would be a correct phrasing, in the
scientific community is generally references as ‘flame area’ and ‘flame speed’ [58]. Combined
visualization from below and the side has resulted in a comparable flame front propagation
speed during the initial combustion stages [44]; therefore, the methodology can be considered
as representative for the overall propagation speed up to the optical limit.
Imaging of the initial flame-kernel growth has proven particularly useful in investigating
combustion behavior at CAD closer to ignition timing, where the thermodynamic prediction is
not accurate [72]. Image processing allowed the evaluation of macroscopic parameters related
to flame morphology, such as flame area, displacement, and deformation [58]. Actual flame
front delimitation is a matter of debate, with no clear chemical species accepted as markers that
identify reaction zones (even though CH is generally accepted as representative in this sense),
and even different types of camera can give close but nonetheless dissimilar flame areas [76].
Nonetheless, the natural flame visualization given by the emission of chemical reactions can be
considered as a good evidence of flame front evolution.
In the case of syngas, to the best of the authors’ knowledge, no major quantitative studies
in the literature have compared comprehensively the in-cylinder pressure evolution with optical
analysis. However, gaseous fuels like methane, methane-hydrogen blends, and pure hydrogen
were already analyzed with optical techniques in ‘real’ engine like condition. Di Lorio et al.
[41] and Catapano et al. [42] tested in an optically accessible engine pure methane and
methane/hydrogen blends (20% and at 40% of hydrogen) at the stoichiometric air-fuel ratio.
Moreover, Aleiferis et al. [44] performed an optical study of the combustion process in a direct-
injection spark-ignition (DISI) research engine with methane, gasoline, isooctane, ethanol, and
41

butanol. In addition, the same authors [82] performed an optical study with pure hydrogen,
between λ=1.2 and 2.0. In this work applied crank-angle resolved flame chemiluminescence
image of the combustion process. Heywood [83], used Schlieren imaging to compare propane
and hydrogen combustion in a square section optical SI engine. Therefore, several authors apply
flame visualization in liquid and gaseous fuels with relevant results in fuel properties in SI
engines. The same line of investigation should be followed to improve the knowledge of the
combustion process of syngas in SI engine.
42

3 Effect of Air-fuel Ratio in Syngas Combustion Process

Previous research has shown the potential benefits of running an engine with air excess
(λ > 1.0). The challenges of running lean have also been identified [58, 84] but not all of them
have been fundamentally explained. Under high dilution levels, a lean limit is reached where
combustion becomes unstable, with significantly deteriorating drivability and engine
efficiency, thus limiting the full potential of lean combustion. By Ayala et al. [84] was
demonstrated that for gasoline a rapid increase in combustion variability near the lean limit is
due to the inverse dependence of the burning time of the turbulent mixture eddies on the laminar
flame speed. In addition, was shown that by reducing the eddy-burning time, the full burn
duration curve can be shortened increasing the lean relative air/fuel ratio (λ) at peak efficiency
and the lean combustion variability limit. This can be done by increasing turbulence levels,
effectively decreasing its microscale structure, or by increasing the laminar flame speed, for
example, through hydrogen enhancement.
Martinez et al. [58] analyze a similar case with application cycle resolved digital
imaging and spectroscopy technique. In this work, lean operation ensured an increase of around
10% in fuel conversion efficiency compared to stoichiometric fueling, and acceptable stability,
even close to the flammability limit of gasoline (λ=1.6). Another interesting result was the
identification of a prolonged flame kernel phase as the main reason for the requirement of
advanced spark timing, as air-fuel mixtures were leaner. The analysis of exhaust gas emissions
confirmed the potential of lean operation for reducing NOx emissions. Moreover, lean burn
operation demonstrated increased flame distortion and center movement from the location of
the spark plug compared to the stoichiometric case (λ=1.0).
The slow-burning velocity of fuels as methane or high-diluted syngas mixtures could be
improved by mixing with hydrogen, whose burning rate is seven times higher than methane in
stoichiometric conditions [39]. Its wide flammability limits and its low quenching gap
contribute to the extension of lean operation limit enhancing thermal efficiency. The absence
of carbon in hydrogen fuel contributes to further reduce the CO, CO2 and HC emissions [42].
The concept of an internal combustion engine running on pure hydrogen is as old as the
engine itself [82]. The lack of established technology necessary to handle some issues related
to the properties of hydrogen, the loss of performance and volumetric efficiency [85] as well as
the projected infrastructure costs for the safe production and delivery of hydrogen on a large
43

scale have discouraged most engine manufacturers from promoting hydrogen as a fuel for their
engines. Nevertheless, sustainability issues and stricter exhaust emissions legislation have made
hydrogen the subject of much discussion with new research for the fundamental understanding
of in-cylinder phenomena. Syngas is believed to be a transition from the carbon-based on the
hydrogen-based energy in the energy sector [1]. Also, represents an option of used hydrogen
without the need for complex systems for its extraction [2].
Although various technical problems of syngas in SI engines have been tackled and
solved, no work has been published on the use of optical diagnostics to investigate in-cylinder
process of H2/CO combustion in modern engine designs. Deeper insight into the effects of fuel
composition on in-cylinder process can be a useful support in the development of advanced
solutions for the new generation of SI engines. The objective of the current chapter was divided
into three sections. The first one, describe the experimental apparatus and the methodology used
to perform the thermodynamic and optical analysis of the combustion process. The second
section, investigate the effect of air-fuel ratio on the combustion characteristic of pure syngas
with two different H2 content (50% and 75%) in CO. The last section, have the focus in the
comparison of hydrogen addition in CH4 (methane blends) and CO (pure syngas). This last
study was centered in lean and extreme lean conditions. It is defined as an extreme limit, the λ
value that not stable combustion is reached.
Therefore, this chapter expands the understanding of lean combustion in gaseous fuels
by explaining the fundamentals behind this rapid rise in combustion variability and the
hydrogen influence in these phenomena. To this aim, the combustion process in a PFI-SI
optically accessible single-cylinder engine was investigated. The experiments were performed
at fixed engine speed (900 RPM) and open throttle (WOT). The spark timing (SA) was fixed
for the maximum brake torque (MBT) of the baseline condition. The same SA allowed the
analysis of flame propagation and combustion behavior at roughly the same fluid dynamics
conditions [55]. In addition, this approach is the most likely scenario of ‘feed-in and run’, for
which engines designed for natural gas are simply ran on syngas, without re-calibration [58].
High spatial resolution cycle-resolved digital imaging was used to characterize the flame
morphology in different crank angle positions. The optical results were correlated with the
thermodynamic data. The main goal of the work is to contribute to the understanding of
fundamental in-cylinder processes of gas combustion and support numerical models.
44

3.1 Experimental apparatus and measurement procedure

3.1.1 Engine test bed

All tests were performed on an optically accessible single cylinder PFI-SI engine AVL
5406 (Figure 3-a). During combustion, the light emission passed through the sapphire window
and was reflected towards an optical detection assembly by a 45° inclined UV-visible mirror
located in the elongated piston, and then recorded by the acquisition system aimed at studying
flame front propagation (Figure 3-b). The cylinder head featured four valves and a spark plug
located 5 mm from the center of the combustion chamber (Figure 4-a). Optical accessibility
was provided via an elongated piston with a wide flat sapphire window in its crown and a quartz
ring replacing the upper part of the cylinder liner. To reduce window contamination by
lubricating oil, self-lubricating Teflon-bronze piston rings were used in the optical section
(Figure 4-b). As can be seen in the picture a large region called top land is present in this type
of research engine. This region enhances the mass and heat transfer and reduces the effective
compression ratio (CR), producing difference with respect to conventional commercial engines
in certain parameters. This aspect is being detailed in the manuscript.
The optical set-up allowed a bottom field of view (64 mm in diameter) that corresponded
to 78% of the piston diameter and 61% of piston area (Figure 4-b). The crankshaft was equipped
with a shaft encoder resolving 3600 increments per revolution. An AVL427 engine-timing unit
was employed for ignition and injection control, as well as for the provision of synchronized
triggering for image acquisition and recording in-cylinder pressure data. Further details of
engine specification are shown in Table 4.
45

a) b)

Figure 3 – Experimental setup: a) single cylinder research engine mounted in the LCPE b)
scheme of the optical experimental arrangement with the camera set up for taking the bottom
view of the combustion chamber. Adapted from [55].

Exhaust Valves

Spark Plug
Optical x
Limit
Pressure
Transducer
Intake Valves

a) b)

Figure 4 – Optical access: a) cylinder head bottom view from an inclined mirror located
below the piston and b) piston design with the detail of optical window, top land region, and
piston ring positions. Adapted from [55].
46

Table 4 – Geometric specifications of the PFI-SI single-cylinder research engine AVL 5406.

Component Size Unit


Total volume 530 cm3
Piston bore 82 mm
Stroke 90 mm
Geometric Compression Ratio 9.7:1 -
Number of valves 4 2 int, 2 exh
Connecting rod 144 mm
Intake valve diameter 34 mm
Exhaust valve diameter 26 mm
Open Intake Valve 718 CAD
Close Intake Valve 204 CAD
Open Exhaust Valve 480 CAD
Close Exhaust Valve 716 CAD
Intake valve lift 10.49 mm
Exhaust valve lift 9.25 mm

3.1.2 Fuels

Although many compounds contain hydrogen, this element is not found in an un-
combined form in a significant quantity on earth [85]. Various processes exist to obtain pure
hydrogen. However, an alternative to using hydrogen is in combined form as syngas. The major
fuel components of a syngas are H2 and CO. In this work, two ratios (by volume) of H2/CO
were investigated; 1) 75%H2 / 25%CO (S75), and 2) 50%H2 / 50%CO (S50). This two syngas
was taken to be representative of the family of various ratios of H2/CO syngas. Methane was
used as baseline fuel, because of their use in several commercial and research application as
pure or as natural gas (NG). In this chapter, hydrogen was also added to the baseline fuel (CH4)
to compare the effects with pure syngas. Two hydrogen content was used, 25% (M25) and 50%
(M50). In addition, several works have studied these fuels in SI engines with optical techniques
[43, 44]. At this point, it is possible to analyze the difference of influence between H2 content
in CH4 and CO. A resume of the fuel used can be seen in Table 5.
Main properties of the fuel used are presented in Table 6 and Table 7. The heating value
of the fuel determines the power that an engine of a given size can provide. The elemental
composition of biomass used determines its heating value. It could be measured experimentally
or alternatively for gaseous fuels it can be calculated from their composition, as it was in the
present case. Moreover, laminar burning velocity referring to stoichiometric conditions has
47

been determined using the Flame-speed calculator CHEMKIN-PRO software module in


conjunction with GRI-Mech 3 [86] reaction mechanism and with its thermodynamic data and
transport properties. Although the addition of hydrogen shows an increase in laminar flame
speed of CH4 as was predicted in the Introduction, this effect was stronger for CO than for CH4.
Around 2.2 times at the same level of H2 content (50% H2 in a volume basis). This establishes
a basis for analyzing the behavior of fuels in the engine combustion chamber, in which other
parameters such as turbulence, temperature and homogeneity of the mixture play a role in the
combustion process.

Table 5 - Fuel composition in the percentage of total volume used for study the effect
of air-fuel ratio in syngas combustion process.
Fuel CH4 H2 CO CO2 N2
[%] [%] [%] [%] [%]
M 100 0 0 0 0
M25 75 25 0 0 0
M50 50 50 0 0 0
S50 0 50 50 0 0
S75 0 75 25 0 0

Table 6 - Properties of basic components present in the fuels used for study the effect of air-
fuel ratio in syngas combustion process.

Properties H2 CO CH4
LHV [MJ/Kg] 121.0 10.2 50.2
LHV [MJ/Nm3] 10.8 12.7 35.8
AFRst mass 34.4 2.46 17.2
Peak Flame Temp [K@1 atm] 2378 2384 2223
Laminar Flame Speed 2.7 0.45 0.35
[m/s @1Atm, 293K, λ=1.0]
48

Table 7 – Main properties of the fuels used for study the effect of air-fuel ratio in syngas
combustion process.

Properties Methane M25 M50 S50 S75


LHV [MJ/kg] 50.2 53.0 58.0 17.5 29.7
AFRst mass 17.2 17.9 19.1 4.6 8.1
Peak flame temp [K @1 atm] 2223 2239 2260 2371 2373
Laminar Flame Speed
0.36 0.43 0.55 1.23 1.77
[m/s @1Atm, 293K, λ=1.0]
CH4/H2 - 3.0 1.0 0 0
CO/H2 - 0 0 1.0 0.3
H2mass [%] 0 4.0 11.2 6.7 17.8
DOD [%] 0 0 0 0 0

The gas filling system consists of individual cylinders of pure gases, i.e. CH4 (99.5 %),
H2 (99.9 %), CO (99.9 %), CO2 (99.9 %) and N2 (99.9 %). The mixture, stored in the 20 liters
volume auxiliary cylinder, is obtained with the partial pressures method of the pure gases and
then fed to the PFI injector (Bosch ML082G). A Perkin Elmer Clarus 580 gas chromatograph
was used for verifying the mixture compositions after preparation. For each mixture three gas
chromatography was performed, and compare with reference calibration gases. An error of ±1%
in volume composition was found. A scheme of the system could be seen in Figure 5.

Figure 5 – Fuel system with representation of: gas storage, panel to prepare the mixture, gas
chromatographer and PFI fuel injector.
49

3.1.3 Test Procedure

The experiments were performed at 900 RPM and maximum load as a representative
point of electric power generation applications in low crankshaft rotational speed, with the
engine operated at full load for extended periods. To achieve the operative point the active
dynamometer bench was set to impose the require crankshaft speed (called motored mode) and
the throttle was total opening.
The engine was equipped with two direct injection (DI) systems optional to the port-
fuel injection (PFI) configuration. In this work, the PFI setup was used, as the most
representative choice for SI engines with gaseous fueling. The pressure was set at 7.0 bar.
Injection pressure was maintained for all conditions, as the rated pressure for the Bosch
ML082G injector designed to operate with gaseous fuels in port fuel configuration.
The duration of injection (DOI), period of time in crank angle degree (CAD) that the
injector is opened, was set to reach the desired relative air-fuel ratio (λ or AFRrel). Values of λ
above 1.0 means lean combustion operation and values below the stoichiometric point indicates
rich combustion process. This parameter was measure using a wideband exhaust gas oxygen
sensor (Bosch LSU 4.9), with an accuracy of ±1%. The oxygen sensor was located in the
exhaust port and an ETAS ES630 module was used for convert the voltage measurements in λ
value. The ETAS module was set to compensate the measurements according to the fuel
characteristics (AFR, C/H and O/H ratio) with respect to pure gasoline (fuel reference of the
oxygen sensor).
The other important parameter of the gas injection is the position in the engine cycle
that the gas is injected. For this work, the end of injection (EOI) was set at 330 crank angle
degree after top dead center (CAD ATDC). This means that the gas injection was during the
exhaust stroke (Figure 6), to provide enough time to obtain a homogeneity air-fuel mixture in
the intake port before the intake valves are opened. Generally, this injection strategy is called
close valve injection [34].
Temperature and pressure of the exhaust gases were measured with the use of a
thermocouple and pressure sensor located in the exhaust runner (30 cm and 40 cm downstream
the exhaust valve, respectively). Coolant and lubricant temperature were maintained at 330-
335 K using a thermal conditioning unit; intake air temperature was in the range of 295-303 K
and the ambient pressure was closer to 950 mbar. For each engine operative condition, tests
were performed according to a procedure of 1 minute of warm-up in motored mode, followed
50

by firing until a stable λ value was obtained (around 15 seconds), after which 200 consecutive
cycles were recorded.
In-cylinder pressure was acquired with an accuracy of ±1% by using a quartz pressure
transducer flush-installed in the region between the intake and exhaust valves (Figure 4-a);
crank angle resolution was 0.1 CAD. Based on this data, the rate of heat release and related
parameters were evaluated. Optical data were detected in the last 25 cycles of the sets of 200 in
order to retrieve information from more stable combustion conditions (see Figure 6). Pressure
values, as well as the image sequences, were related to piston movement by triggering the data
acquisition system in gated mode.
In this work, all timings given in crank angle degree (CAD) with 1 CAD corresponding
to 185 μs (at 900 RPM). Figure 6 shows a scheme of the injection pulse (duration 61 CAD and
end of the injection 330 CAD ATDC), the in-cylinder pressure signal for the baseline case
(methane with stoichiometric AFR), the pressure before the fired cycles (motored pressure), the
spark plug signal (set for this example the start of the ignition at -7 CAD ATDC) and the camera
trigger signal for the images acquisition (start of acquisition -9 CAD ATDC and with 40 CAD
of duration).

Figure 6 – In-cylinder pressure and auxiliary measurements (motored pressure, spark plug
signal, camera trigger and PFI injection pulse) for methane at baseline conditions (λ=1.0,
SA = -7 CAD ATDC).
51

3.1.4 Thermodynamic analysis

In this work, direct comparison of in-cylinder pressure trace was used to study the
behavior of the different fuels and operative condition. The comparison is done by the average
of pressure traces of 200 consecutive cycles for each operative condition (see Figure 7-a). Also,
statistic parameters as coefficient of variation (Cov) is used with a statistical population of 200
fired cycles (see Figure 7-b).

(a) (b)

Figure 7- a) In-cylinder pressure trace for the average of 200 cycles and several single
pressure traces. b) Indicate mean effective pressure (IMEP) and Maximum In-cylinder
pressure (Pmax) values for 200 consecutive cycles for methane at baseline conditions
(900 RPM, λ=1.0, SA = -7 CAD ATDC).

In addition, special calculations are performed to expand the information available to


the analysis. Several methods based on the analysis of recorded in-cylinder pressure [87] are
used to determining combustion parameters such as initiation, duration or optimum phasing.
An improved method that directly relates the amount of fuel chemical energy released by
combustion to the changes in pressure is the heat released approach [34], based on a first law
analysis is shown in Equation 1:

𝑑(𝑀𝐹𝐵) 𝑄𝑓 = 𝑑(𝑈) + 𝑑(𝑄ℎ𝑡 ) + 𝑝 𝑑(𝑉) + ℎ 𝑑(𝑚) (1)


52

where 𝑴𝑭𝑩 is the mass fraction burned in the combustion process, 𝑸𝒇 is the heat released by
fuel oxidation (J), 𝑼 the internal energy of the working fluid measured (J), 𝑸𝒉𝒕 the heat loss to
the cylinder walls (J), 𝒉 the fluid specific enthalpy (J/kg) and 𝒅𝒎 the variation of cylinder
charge (kg).
Therefore, with the resolution of equation 1 in a custom numeric code model developed
by Irimescu et al. [94], the duration of the combustion process can be quantified by the MFB
curve. In general, the burning rate is expressed in crank angle duration from the start of the
spark for a fixed percentage of mass fraction burned (MFB). In this work, three values were
taken to the analysis. The duration between 0-5% MFB that correspond to flame kernel
formation angle, the duration between 0-10% MFB associated to the flame development and
the position of 50% MFB named as the rapid burning angle [34].
Heat transfer represents an important parameter in Equation 1 and needs to be studied
in detailed. The importance of turbulence in heat transfer has been long recognized and several
ways to account for its generation and dissipation have been investigated [88]. Also, previous
works [89, 90] showed that the combustion of pure hydrogen and blends with a high content of
hydrogen exhibits a higher cooling loss to the combustion chamber wall of an internal
combustion engine compared to hydrocarbon combustion. This was identified due to its higher
burning velocity and shorter quenching distance. The result shows that traditional models as
Woschni [91], Annand [92] and Eichelberg [93] equations calculate a lower cooling loss than
experimental values, and the use of correction coefficients does not accurately define the actual
cooling rate. The effects of flame development were also found to be important for heat transfer,
which hinders the application of simple convective models.
Therefore, a new correlation is required when H2 is fuelled such as the convective heat
transfer correlation developed by Irimescu et al. [94]. This study shows an empirically defined
equation for calculating the convective heat transfer coefficient combined with a more
fundamental view of fluid velocity. The latter parameter was calculated by applying a cascade
model of kinetic energy transfer, with the effect of burned gas expansion included during
combustion. The heat transfer correlation is shown in Equation 2:
𝑇 0.11
𝑁𝑢 = 0.037 𝑅𝑒 0.8 𝑃𝑟 0.33 (𝑇 ) (2)
𝑤

where 𝑻 is fluid temperature (K) and 𝑻𝒘 is the correction accounting for the temperature
difference of the fluid within the boundary layer. Prandt (Pr), Reynolds (Re), and Nulset (Nu)
are non-dimensional numbers.
53

The last parameter that has to be calculated is 𝒅(𝒎) , associated to the variation of
cylinder charge. Calculations without considering this parameters, usually ensures good
accuracy when performing heat release analysis in commercial engines [57]. This is due to the
fact that top-land, head gasket, spark plug thread and other crevices volumes are a small
percentage of the clearance volume. However, engines with optical access have special
geometry design (Figure 4-b) with relatively high crevices volume and losses for the crankcase
due to the special non-metal rings. This means high blow-by rates that have a significant impact
on the first-law analysis (Equations 1) [34]. Therefore, a model to predict blow by losses must
be applied in optical engines. For this work, a simplified compressible flow model was
employed for calculating blow-by flow [57]. Mass flow rates were obtained via Equation 3:

𝜌𝑢 𝑤𝑏𝑏 𝐴𝑏𝑏 𝑑𝜃
𝑑(𝑚) = (3)
𝑁

with 𝑨𝒃𝒃 is considered as an equivalent flow area (m2). This parameter was tuned to reach the
end of the combustion just before the exhaust valve open. This hypothesis is also used in other
mass fraction burned calculation [87]. The parameter 𝝆𝒖 is the density of unburned gases
(kg/m3), due to the assumption that all the lost mass is associated to the unburned mixture. The
𝑵 is the crankshaft rotational velocity. Finally, 𝒘𝒃𝒃 is the blow-by flow velocity (m/s). A
simplified compressible flow model was employed for calculating with separation to chocked
flow (Equation 4) and non-chocked flow (Equation 5).

𝛾−1
𝛾 𝑅 𝑃0 𝛾
𝑤𝑏𝑏 = √2 (𝛾−1) 𝑀 𝑇 [1 − ( 𝑃 ) ] (4)

𝛾 𝑅
𝑤𝑏𝑏 = √2 (𝛾+1) 𝑀 𝑇 (5)

with 𝛾 being the heat capacity ratio, 𝑅 the constant of gases (J/K mol), 𝑀 the mach number
and 𝑃 the pressure of the fluid with respect to the reference 𝑃0 (Pa).
54

The performance was measured with Indicated Mean Effective Pressure (IMEP) defined
[34] with Equation 6.
2 𝑃𝑜𝑤
IMEP = (6)
𝑉𝑑 𝑁

with 𝑷𝒐𝒘 as indicated power generated by the engine (w) and 𝑽𝒅 displacement volume (m3).
In general, IMEP (bar) is used as performance value because allows to compare engine of
different size as parameters like power or torque. Therefore, it is a suitable coefficient to
compare the performance of the different fuels with others authors.
Combustion stability is represented by the cyclic variability derived from pressure data.
Several authors agree that the coefficient of variation from indicated mean effective pressure
(CovIMEP) is the most accurate parameter to measure this phenomenon (Equation 7). Another
parameter that quantifies the stability of a combustion process in SI engines is the coefficient
of variation from maximum pressure (CovPmax) that is shown in Equation 8. In this work, both
parameters are analyzed. However, CovIMEP was followed to determine the stable regimen
[34].
𝜎𝐼𝑀𝐸𝑃
CovIMEP = 100 (7)
IMEP
𝜎𝑃𝑚𝑎𝑥
CovPmax = 100 (8)
Pmax

with 𝝈𝑰𝑴𝑬𝑷 , 𝝈𝑷𝒎𝒂𝒙 the standard deviation of IMEP and Pmax, respectively. Finally, fuel
conversion efficiency [34] was calculated by using Equation 9:

IMEP 𝑉𝑑
𝜂𝑓 = (9)
𝑚𝑓 LHV

where LHV is shown in Table 7 for each fuel and the mass of fuel injected per cycle 𝒎𝒇 (kg)
was calculated with injector calibration table and duration of injection fixed in each test with
an accuracy of ±1%.

3.1.5 Exhaust measurements

Finally, pollutant species concentrations (CO, CO2, CH4, and NOx) were measured in
the exhaust gas stream using a Multigas 2030 spectrometer analyzer (MKS Instruments,
Massachusetts, USA). The equipment works with the Fourier Transform Infrared (FTIR)
55

principle, capable of ppb to ppm sensitivity for multiple gas species with one sample for each
second of the test. Gas line heater maintains temperature before the sample enters the gas cell
and the equipment provides automatic temperature and pressure compensation to ensure
accurate analysis. The resolution was 1 ppm for CO, 0.001% for CO2, and 0.5 ppm for the other
two chemical species, all within 5% accuracy. The emission of NOx was calculated by the
summation of individually calculated brake-specific emissions of nitric oxide (NO) and
nitrogen dioxide (NO2). Because of the short period that the experiment is in fire operation
(below 30 seconds), only the last measured point was taken as representative of the combustion
process for each chemical species.

3.1.6 Optical Investigations

Flame front propagation was investigated by cycle resolved digital imaging. A high-
speed 12 bit CMOS (PCO Dimax S1) camera was coupled with a double intensifier (Video
Scope VS4-1845HS). It is widely recognized that taking combustion images inside an engine
with optical access to its combustion chamber is a powerful method to investigate the
combustion quality, particularly in the early flame development stage where the use of pressure
sensor may not provide reliable combustion data [95]. Thanks to the recent advances in high-
speed imaging hardware and advanced processing algorithms, it is possible to determine more
precisely the key characteristics such as the flame front development, flame structure,
propagation speed and cycle-to-cycle variations of combustion inside a real engine. Even
though the recorded combustion image was the 2D projection image of a 3D flame [61], it still
provided important and useful information such as the flame location and flame size.
The optical assembly in the tests performed for this manuscript detected natural flame
emission from the bottom view. The system allowed a high sensitivity in the spectral range
from 290 nm to 700 nm, with 50% quantum efficiency at 450 nm. The camera could work in
full chip configuration (1008x1008 pixel) with a maximum frame rate of 4467 fps. In order to
improve the acquisition speed, a region of interest of 864x896 pixel was selected; this permitted
to reach a frame rate of 5400 fps, corresponding to 1 image/CAD at 900 RPM
(1 CAD = 185 μs). The detection system was equipped with UV-Nikkon 105 mm.
To improve the signal to noise ratio, the level of intensification was modified to obtain
good quality images. Moreover, the f-number of the objective was set to ensure non-saturation
of the image for camera protection and better image post processing. It was seen that the images
of combustion process with high hydrogen content or with air-fuel ratio near the stoichiometric
56

presents the highest luminosity emission. Therefore, for these cases the exposure time was
decreased. If the image still being saturated, the f-number was increased.
The optical setup allowed detecting image sequences with a spatial resolution of
87 µm/pixel. For all the optical measurements, the synchronization between the cameras and
the engine was achieved through the crank angle encoder signal and the delay unit. Also, this
allows the synchronization with pressure traces.
The application of a custom procedure for the image processing allows a detailed
analysis of flame morphology [58]. For this work, a routine was developed in Vision software
of National Instruments (Vision Assistant 2016, NI ACADEMIC SITE LICENSE, Austin,
Texas, USA). It is important to note that this software not allows to process a image in any
configuration different to 8-bit. Therefore, the first step was to transform the 12-bit images in
8-bit to be treated to retrieve geometrical parameters of the flame front. Following the procedure
sketched in Figure 8, after the extraction of the intensity level, a circular mask was fixed in
order to cut light from reflections at the boundaries of the optical access (Figure 8-a).
Successively, the image processing procedure adjusted the contrast and brightness of the images
with respect to the maximum intensity value in order to optimize the signal to noise ratio (Figure
8-b). Then, a threshold was applied to obtain binary images, with 1 (white) associated to a pixel
belonging to the object (foreground) and 0 (black) was referred to the background (Figure 8-c).
In this work, automatic threshold operation (metric method) based on a locally adaptive
algorithm was used [96]. This methodology for threshold selection reduces the time of post
processing. In addition, in contrast to manual thresholding, this methodology did not require
the set of the minimum and maximum image intensities. Therefore, this avoids overestimation
of the flame area in the kernel formation and have good accuracy in the flame propagation
phase. The metric method consists that for each threshold selected, a value determined by the
surfaces representing the initial grayscale is calculated (Equation 10). Therefore, the selected
threshold value is the pixel value k at which the following expression (10) is minimized:

∑𝑘𝑖=0 ℎ(𝑖)|(𝑖 − 𝜇1 )| + ∑𝑁−1


𝑖=𝑘+1 ℎ(𝑖)|(𝑖 − 𝜇2 ) | (10)

where 𝒊 represents the gray level value, k represents the gray level value chosen as the threshold,
𝒉(𝒊) represents the number of pixels in the image at each gray level value. Moreover, N
represents the total number of gray levels in the image (256 for an 8-bit image). Finally, 𝝁𝟏 is
the mean of all pixel values in the image that lie between 0 and k and 𝝁𝟐 is the mean of all the
pixel values in the image that lie between k+1 and 255.
57

After this step, morphological transformations were applied to fill holes and remove
small objects that were not part of the flame and could bias the evaluation of morphological
parameters (Figure 8-d). Finally, the outline border of the flame was extracted (Figure 8-e). \

Figure 8 – Sketch of the image processing steps; intake, exhaust valves, spark plug, and
optical limit are also shown. The axes x, y mark the positive values for the calculus of the
center movement.

The results of image processing consisted in the flame Area (A), Waddel Disk Diameter
(WDD), Heywood Circularity Factor (HCF) and flame centroid movement. The flame area 𝑨
corresponded to the number of pixels included in the foreground of binary images. The WDD
was the diameter of a disk with the same area as the binary flame (mm), calculated with
Equation 11:
WDD = √4 𝐴⁄π (11)

This could be seen as an equivalent diameter of the flame, in which represent a comparative
parameter to numerical models of engine simulation and be comparable with constant volume
chamber experiments among others (Figure 9). A comparative graph between A and WDD
against crank angle after the start of the spark (CAD ASOS) is shown in Figure 10. The flame
area was graphed by scaling the result in pixels and normalized to the cylinder cross-section.
The x-axes represent the distances in CAD from the moment when the energy of the spark is
released in the chamber to begin the combustion process. This allows comparative studies of
cases with different spark timing (SA).
In general, flame diameter under 8 mm was not possible to detect, this is in line with
other researchers works [66, 68]. The graphs presented in this work are an average value, taken
from 25 consecutive cycles. A complete analysis of the optical parameters as well as post-
processing variables was presented in Appendix A, with also quantification of dispersion due
to cycle-by-cycle variations.
58

Figure 9 – Overlap between the binary image of the flame front and the calculated equivalent
diameter with the WDD method.

a) b)

Figure 10 - Flame size measurements as the average of 25 consecutive cycles of methane for
λ=1.0 and SA= -7 ATDC a) Normalize flame area b) Flame diameter calculated as WDD.

Regarding the propagation speed, no unified definition of turbulent flame speed can be
found in the literature; generally, it is referred as the measured displacement of the flame front,
determined based on the optical data recorded from below the combustion chamber [58].
Known the limitations of chemiluminescence imaging techniques, typically associated with the
line-of-sight integration of flame front structure. Aleiferis et al. [44] results obtained by flame
chemiluminescence and horizontal laser flame tomography (Mie scattering) have shown
comparable values between the two techniques in terms of flame morphology and propagation
speed. This provided further confidence in the validity of the results. Therefore, in this study
the flame propagation speed 𝑺𝒊 (m/s) was calculated as the incremental ratio of the WDD
between two frames with respect to the dwell time (Equation 12).
59

1 𝑊𝐷𝐷𝑖− 𝑊𝐷𝐷𝑖−1
𝑆𝑖 = ( ) (12)
2 ∆𝑡

The dwell time (∆𝒕) between two images was set in 185 µs (1 CAD), this value provided
enough information of the flame development as in accordance with Merola et al. [63] and
Aleiferis et al. [69]. Figure 11 shows a scheme of the calculus of flame propagation speed, it
represent a circumference that have an increase in the radius with the time. Figure 12 shows the
characteristic graph of average flame speed propagation for methane in which start from low
values near 2 m/s, reaching a maximum (this case is 12.6 m/s). The steadily decrease is due to
the different instant that the flame hit the wall in the several cycles and the hypothesis that the
flame is a perfect circle. Therefore, the average flame value calculated from the average flame
diameter is affected after reach the maximum. This need to be taken in account in the analysis
of the flame speed propagation along this work. However, the first phase of the combustion
propagation is well represented with this methodology. Therefore, important information as
peak flame speed and initial speed propagation are being taken.

Figure 11 – Representation of calculus of flame front propagation speed.


60

Figure 12 - Flame propagation speed measured as an average of 25 consecutive cycles for


methane with λ=1.0 and SA=7 BTDC.

Flame front distortion was evaluated in terms of Heywood Circularity Factor (HCF) that
corresponds to the ratio between the perimeter (P) of the image and the circumference of a
circle with the same area, according to Equation 13:

𝑃
HCF = (13)
𝜋∗𝑊𝐷𝐷

where 𝑷 (mm) was obtained by scaling the result measured in pixels. A value of HCF equal to
1.0 means a perfect circumference shape. Figure 13 shows the baseline case, values over 1.0
represents flame distortion. In the last phase of the propagation process the circularity return to
1.0 due to the complete of the optical limit. Therefore, the last phase of the combustion process
registered by this analysis is affected by the size of the optical window.
61

Figure 13 – Distortion of the flame front measured as an average of 25 consecutive cycles


calculated with the HCF equation for methane with λ=1.0 and SA=7 BTDC.

Finally, flame centroid was calculated as the arithmetical center of luminosity evaluated
for a binary image. It was identified by the x and y coordinates with respect to the Cartesian
system fixed in the center of the combustion chamber (Figure 14). In general, for gaseous fuels
as methane in this engine configuration, the average movement begin near the spark plug and
propagate towards the intake valves as could be seen in Figure 15. After the maximum
displacement is reached, the flame returns to the center of the combustion chamber. This is due
to the flame reach the entirely flame window that is centered with respect to the combustion
chamber.

Figure 14 – Representation of flame outline for the calculus of HCF and position of the
geometric centroid with respect to the center of the combustion chamber.
62

Figure 15 - Average flame front movement in the combustion chamber for methane in λ=1.0
and SA=7 BTDC.

As was mentioned in the Introduction, despite the fact that the view from the bottom of
the combustion chamber allows only a line-of-sight evaluation (projected burnt area), other
studies [44] has resulted in a comparable flame front propagation speed when lateral and bottom
view is used.
For the baseline case (methane, λ=1.0, SA= -7 CAD ATDC) a lateral view acquisition
was performed to obtain qualitative information of the combustion process (Figure 16). A
dashed line was added to the image sequence to shows the TDC position (0 CAD ATDC). It
can be seen that the flame hit the piston surface at 8 crank angle degree (CAD) after start of
spark (ASOS), this means 1 CAD ATDC. To put in context, the piston moves up in z axes
0.4 mm from the beginning of the combustion process or start of spark (SOS) to top dead center
(TDC) and then only 0.1 mm when the flame reach the piston. Therefore, the flame front has
free movement only up to 22 mm of flame diameter (7% in terms of normalized area), measured
from the bottom view (Figure 10). Afterward, the flame has the same propagation speed than
the piston in z-axes and the interesting information is in the other two axes (x-axes and y-axes).
Therefore, the methodology used in this study can be considered as representative for the
overall propagation speed up to the optical limit.
63

Figure 16 – Combustion process lateral view of methane at 900 RPM, λ=1.0,


SA = - 7 CAD ATDC and wide open throttle (WOT).

3.2 Stoichiometric to lean conditions

As was described previously, the first analysis was concentrate in CO/H2 blends and the
comparison with pure methane in stoichiometric air-fuel ratio (λ=1.0) and lean air-fuel mixture
(λ=1.2-1.4). Two hydrogen content (50% and 75%) was selected as a representative family of
syngas mixtures (S50 and S75).

3.2.1 Operating Points

The air-fuel ratio was initially set at stoichiometric value (λ=1.0) and then increased
until reaching the flammability limit of methane (λ=1.4). Spark advance (SA) was fixed at
- 7 CAD ATDC, that corresponded to the spark timing for the maximum brake torque (MBT)
of the baseline condition (methane in stoichiometric AFR). The same SA allowed the analysis
64

of flame propagation and combustion behavior at roughly the same fluid dynamics conditions
[55]. Therefore, the fuel properties could be study in detail with minimum influence of other
factors as pressure, tumble and swirl motion, among others.
A resume of the test condition can be seen in Table 8. The duration of injection (DOI)
for syngas blends was higher than methane for reach the same λ. This was attributed to the
difference in lower heating value (LHV) of the fuels (S50 17.5 MJ/kg and S75 29.7 MJ/kg,
methane 50.2 MJ/Kg).

Table 8 – Operating conditions set for study the effect of air-fuel ratio from stoichiometric to
lean conditions in syngas combustion process. Engine speed: 900 RPM; SA: 7 CAD BTDC;
Load: WOT.

Fuel λ DOI
[CAD]
1.0 61
M 1.2 50
1.4 43
1.0 204
S50 1.2 180
1.4 150
1.0 150
S75 1.2 130
1.4 105

3.2.2 Thermodynamic Results

In-cylinder pressure traces were measured for 200 consecutive fired cycles and then
analyzed with the methodology presented in section 3.1.4. Due to the cycle-by-cycle variations
that presents ICEs, in this work all the thermodynamic graphs and values are presented as the
average of the 200 consecutive cycles at stable λ. Therefore, the thermodynamic analysis gives
a global perspective of the combustion process evolution; the main parameters were indicate
mean effective pressure (IMEP) as an indicator of performance, peak in-cylinder pressure value
and its position in terms of crank angle degree (CAD), as well as stability measured by the
coefficient of variation of IMEP (CovIMEP). Figure 17-a shows the pressure traces for the
three gas mixtures tested with λ=1.0 and spark timing optimized for the higher performance
(IMEP) of methane. In addition, the motored pressure (dashed line) are plot as a reference with
respect to the fired operation. The syngas blends depict higher pressure in the first phase of the
combustion, with a peak around 3 CAD ATDC. However, this trend is reverted after methane
65

reach the maximum (22 CAD ATDC) and in almost all the expansion stroke the methane in-
cylinder pressure is higher than pure syngas blends.
For the case of λ=1.4 (Figure 17-b), the difference between fuels is even greater than
the stoichiometric case in terms of peak pressure. The maximum for CH4 decreased up to the
point that is at the same level of the motored trace. For syngas blends, the pressure traces are
lower than the stoichiometric case as expected but maintain a high level of pressure (over 25
bar). The case of λ=1.2 is not shown graphically, but maintains the same trend, with higher
pressure for S75 than the other fuels. Moreover, it could be seen a shift to the right in the
position of the maximum pressure value in lean cases with respect to stoichiometric air-fuel
ratio.
Figure 18 illustrates the variation in the in-cylinder pressure when the λ value is
increased for S50. The spark timing is fixed; therefore, the peak value decrease (5.0 bar) and
the position of the maximum is shifted to the right (5.2 CAD) by the effect of the slower
combustion process in the leanest condition than in stoichiometric AFR. Moreover, in the last
phase of expansion not large difference is found between the three cases.

(a) (b)

Figure 17 – In-cylinder average pressure of 200 consecutives cycles to methane, S50 and S75
at SA = 7 CAD BTDC in: (a) λ=1.0 and (b) λ=1.4.
66

Figure 18 – Average pressure with S50 and SA = 7 CAD BTDC for the three air-fuel ratios.

As reported in Figure 19, the performance (IMEP) was found to be comparable for both
pure syngas blends. In spite of S75 presents higher maximum in-cylinder pressures, S50 has
slightly high IMEP values (around 0.4 bar). This effect is explained in the drop of the pressure
traces after reach the peak, which S75 is more abrupt. When comparing with methane, engine
performance levels for syngas resulted lower than the reference fuel (M) for stoichiometric and
intermediate λ values. This is in line with the results presented in Figure 17, that shows higher
pressure values in the last phase of the expansion stroke for methane than pure syngas blends.
This phase of the engine cycle has the largest variation in terms of volume displacement. In
addition, the spark timing set up to maximize the performance for methane has a strong
influence on the differences. The trend was reverted for λ=1.4, in which S50 presents the highest
IMEP value. This result enhances the necessity to retard the spark timing for pure syngas
blends, and the limitation of apply the ‘feed and run’ concept.
67

8.0
M S50 S75
6.0 5.4

IMEP [bar]
4.8
4.2 3.9 4.0
3.8 3.7 3.9
4.0 3.6

2.0

0.0
λ=1.0 λ=1.2 λ=1.4

Figure 19 – IMEP average values for 200 consecutive cycles, for M, S50, and S75 in λ range
from 1.0 to 1.4 at SA = 7 CAD BTDC.

Figure 20 shows engine stability for the investigated operating points, evaluated through
the coefficient of variation of IMEP (CovIMEP) over 200 consecutive cycles. It was found to
be quite stable for all condition of operation. S50 shows more variability than S75, for all λ
tested. When using pure syngas the CovIMEP was higher compared to methane in the
stoichiometric AFR case, suggesting not optimize spark timing. For the leanest case, the trend
was reverted with methane over the maximum value considered as stable operation (IMEP=3%)
[42]. This is an interesting observation, given the necessity of control thse parameters with
respect to the fuel composition. Also, the high instability for methane in λ=1.4 is associated to
be closer to the flammability limit of the fuel. In this point the air-fuel mixture have an important
decrease in the combustion speed, so the combustion can not be propagate with stable operation.
This also explain the drop in combustion peak pressure saw in Figure 17 for methane. This
point will be better explained in the section of optical measurements.
68

5.0
M S50 S75
4.0

CovIMEP [%]
3.4
3.0 2.5
2.1
1.8
2.0 1.6
1.2 1.2
0.8
1.0 0.6

0.0
λ=1.0 λ=1.2 λ=1.4
Figure 20 – CovIMEP average values for 200 consecutive cycles, for M, S50, and S75 in λ
range from 1.0 to 1.4 at SA = 7 CAD BTDC.

The in-cylinder peak pressure, which is another parameter that can be extracted from
the pressure traces, was also graphed (Figure 21) for three λ ratios. Syngas showed a large
difference in this parameter with respect to methane for all the cases. However, both syngas
mixtures values are comparable with only an increase of 3% for a difference of 25% of H2
content between S50 and S75.
The coefficient of variation of peak in-cylinder pressure (CovPmax) is also a parameter
that is used to analyze cycle by cycle variation. The difference between CovPmax and
CovIMEP is that the first represents the local variation of pressure trace and the second one
take into account all the cycle behavior. Sun [97] demonstrated that CovPmax could be used
only to evaluate the cycle variations in ICEs fueled with a high quantity of hydrogen under
medium to high-speed conditions. Therefore, IMEP is possibly the most appropriate parameter
for evaluating the cycle-by-cycle variations. However, as maximum pressure is a local value,
CovPmax could be taken as a measurement of variability in the first stage of the flame
propagation. Therefore, generally it is used to compare with flame front morphology variations.
In Figure 21 the CovPmax depicts an opposite trend with respect to the CovIMEP. Methane
shows the highest values and S75 the lowest for all test condition.
69

M S50 S75
40 35
34 32 33
31
Pmax [bar] 30 27
29
23
20 16

10

0
λ=1.0 λ=1.2 λ=1.4
8
7
CovPmax [%]

6 5.5
5 4.7
4 3.3
3 2.1
2 1.4 1.7
0.9 1.2
1 0.7
0
λ=1.0 λ=1.2 λ=1.4

Figure 21 – Pmax (top) and CovPmax (bottom) average values for 200 consecutive cycles, for
M, S50, and S75 in λ range from 1.0 to 1.4 at SA = 7 CAD BTDC.

A more detailed analysis of the averaged pressure traces was performed using the heat
release approach described in the previous section. The mass fraction burned traces are shown
in Figure 22-a, which emphasize the fact that large difference exist in combustion speed
between pure syngas mixtures and methane. For 10% of MFB, a difference of 12 CAD exists
between S75 and methane. However, the two syngas present comparable values with a
difference of 2 CAD. At middle level of MFB (50%) the difference increase to 15 CAD and 5
CAD respectively. Therefore, the great difference existing between syngas and methane in
combustion velocity can be associated to the first stage of flame propagation.
Figure 22-b shows the MFB at λ=1.4, the same trend is illustrated with respect to λ=1.0
but an increased difference in the crank angle at the same MFB is registered. Methane at 10%
MFB and 50% MFB increase 5 CAD and 11 CAD in λ=1.4 with respect to the stoichiometric,
meanwhile, the syngas blends only suffer a change of 3 CAD for 10% MFB and 2 CAD for
70

50% MFB approximately. This value shows that for methane the lean condition affects more
the combustion time than for both syngas mixtures. The high hydrogen enhances the first stage
of combustion, which is considered the most important phenomena in the flame propagation
[72]. These values are verified and study in detail with the cycle resolved images analyzed in
the next section.
For S50 with different air-fuel ratios, the MFB traces are presented in Figure 23. The
thermodynamics analysis suggest a higher combustion speed for stoichiometric blend than lean-
burn cases. The crank angle of 10% MFB for λ=1.0 is 1 CAD BTDC, meanwhile with λ=1.4 is
approximately 2 CAD ATDC. This suggests an initiation of the flame faster for the
stoichiometric case than lean combustion process. This result is expected and in agreement with
others results presented in the bibliography, although it is noticeable that not large difference is
registered with λ=1.4 with respect to the stoichiometric case. Therefore, this type of syngas
allows working in more extreme lean condition without losing performance.

(a) (b)
Figure 22 – MFB average values for 200 consecutive cycles, for M, S50, and S75 at
SA=- 7 CAD ATDC and λ: (a) 1.0 and (b) 1.4.
71

Figure 23 – MFB average values for 200 consecutive cycles S50 at SA = 7 CAD BTDC and λ
range from 1.0 to 1.4.

Figure 24 shows the fuel conversion efficiency, calculated as the ratio between work
output and chemical energy content of the injected fuel (see section 3.1.4). For λ=1.0 a diference
of 5% in terms of effiency was found between methane and pure syngas. This is in line to the
in-cylinder pressure traces results and the factor that spark timing is not optimized to syngas
blends [33]. In the case of a lean condition, the trend is reverted and S50 have the maximum in
terms of effiency due to the performance of methane decay for that condition. Therefore, this
suggest that lean condition is suitable operation condition for blends with a high percentage of
H2.
72

30
M S50 S75
25
Fuel Conversion 20 21 21

Efficiency [%]
20 19
20 17
16 16
14
15
10
5
0
λ=1.0 λ=1.2 λ=1.4
Figure 24 – Fuel conversion efficiency for M, S50, and S75 with three fuel ratios and
SA = 7 CAD BTDC.

The analysis of exhaust gas measurements give more information on the chemistry of
combustion for the investigate conditions (Figure 25). It is noticeable that for optical engine
due to the top land region a large amount of fuel reminds without being burned. Therefore, the
concentration that will be registered in the exhaust depends on the mixture burned. For syngas,
this high concentration will be with CO and H2 and for methane blends, CH4 was expected.
Because for this test no measurements of H2 was available, this phenomenon will be only seen
with CO and CH4 channels.
A comparison of the three fuels tested in the leanest case is presented in Figure 25.
Carbon monoxide (CO) concentrations were low, with only syngas mixtures above 0.1%. The
methane concentration (CH4) was also measured in the exhaust. It was observed negligible
values for syngas blends. However, a high concentration was measured (around 1.4%) for
methane blends. This in agreement with other works [55, 58, 59, 68] that found the same
behavior in optical research engines.
Figure 26 shows the NOx concentration in which syngas blends shows the highest
values. Meanwhile, methane has almost negligible values. Because the NOx emissions have
strong depends on temperature (also could be associated with maximum pressure in the
chamber) this graph has a similar trend that was seen in Figure 21 (maximum pressure values).
For λ=1.4, syngas blends have 1.5 times the maximum pressure of methane blends, this causes
an increase of NOx emission from 10 ppm to 3000 ppm approximately.
73

M S50 S75 M S50 S75


2000 5
1555
4
1500
CO [ppm]
1179

CH4 [%]
3
1000
545 2 1.4
500
1
0 0
0 0
λ=1.4 λ=1.4
a) b)
Figure 25 – Emission concentration on the exhaust port for M, S50, and S75 at λ=1.4.
a) CO in ppm and b) CH4 in percentage.

M S50 S75
4000
3076 3115
3000
CO [ppm]

2000

1000
10
0
λ=1.4
Figure 26– NOx emission concentration on the exhaust port for M, S50, and S75 at λ=1.4.

3.2.3 Optical investigations

Even if the in-cylinder pressure measurements allow a comprehensive analysis of the


combustion characteristics, it does not furnish detailed results of the local distribution of the
burned mass, flame behavior inside the combustion chamber and neither the flame propagation
speed. In this sense, cycle resolved visualization represents a powerful tool for quantitative
analysis of flame front propagation.
The camera set up for this test maintain the f-number and the exposure time for the
different λ conditions and fuels to ensure not saturation of the image for high hydrogen content.
74

To improve the signal to noise ratio, the level of intensification was maintained in the minimum
of the linear scale of the intensifier for λ=1.0 and λ=1.2. For these values, the light intensity is
high. However, an increase of intensification value was used to maintain the quality of the
images for the leanest case. An intensifier is a necessary tool because gives the opportunity to
analyze images in low emission regimes as lean combustion process. Table 9 shows a resume
of the camera set up.

Table 9 - Camera set up for imaging acquisition set for study the effect of air-fuel ratio
from stoichiometric to lean conditions in syngas combustion process. F-number: 32, Exposure
Time: 0.5 CAD
Fuel λ Intensification
1.0 55000
M 1.2 55000
1.4 65000
1.0 55000
S50 1.2 55000
1.4 65000
1.0 55000
S75 1.2 55000
1.4 65000

The images of Figure 27 correspond to 5%, 10% 20%, 30%, 40%, and 60% with respect
to the total piston area. The last case is the total optical area that has 64 mm in diameter over
82 mm of the piston. It is noticeable the difference in intensity of emission between S50 with
respect to S75 and methane. In concordance with the bibliography [41, 44], there are two effects
that vary the emission intensity. One can be associated with the heat released and temperature
in the combustion chamber (also associated with IMEP and Pmax respectively) and the other is
the fuel composition (radical contents in the flame). In this case, the dominant factor seems to
be the composition of the mass burned because the trend in IMEP shows an inverse proportion.
The higher IMEP is measured in methane; meanwhile, the lowest intensity in the images for
the entire range is for that fuel. For syngas mixtures, S50 has higher IMEP and lower maximum
pressure than S75 but the light emission measured was lower for S75. Therefore, higher CO
content in the fuel and IMEP for S50 lead to higher values of flame intensity emission.
However, this is not a trivial topic and more analysis should be applied to understand the
behavior of the flame emission. Some of this complement test could be spectroscopy analysis
[58] and image filtration with pass-band filters [41].
75

Figure 27 – Images of M, S50 and S75 combustion process in the position of 5%, 10% 20%,
30%, 40% and 60% with respect to the total piston area at λ 1.0 and SA = 7 CAD BTDC.

A comparison between syngas fuels in two λ ratios can be seen in Figure 28. The images
sequence was selected in the same spatial and temporal position to compare the flame
development for the different syngas mixture in the stoichiometric and poor combustion
process. In the case of λ=1.0, the major presence of H2 increase the flame speed and also the
flame wrinkling. This last parameter can be seen as a smooth flame surface. For the cases of
lean burn condition, the trend is similar with respect to higher flame propagation speed for the
case of syngas S75. It is noticeable that the intensification value set in the camera was higher
to λ=1.4, therefore no comparison can be made in this figure between the two λ ratios. However,
it is evident that a decrease in flame emission intensity was recorded when λ increase. In
addition, more deformation and flame wrinkling can be extracted as a conclusion in Figure 28
between the two λ values.
76

Figure 28 – Images of M, S50 and S75 combustion process in several crank angle position
positions at λ=1.0 and λ=1.4 with SA = 7 CAD BTDC.

The post-processing code allows the study of the flame morphology with parameters as
flame area, propagation speed, deformation, and center position as was explained in the image
processing section. Figure 29 compares the flame development of the three fuel mixtures tested
for λ=1.0. The large difference could be seen between methane (baseline fuel) and syngas
mixtures, in which both CO/H2 blends reach the optical window in 1/3 of time (time and crank
angle duration have a linear relation) compared with baseline fuel. The optical limit was
indicated with a horizontal dashed line, in 61% of total piston area. Flame speed has the
maximum in the first images registered for both syngas mixtures and then decrease steadily,
different for methane that was close 15 CAD ASOS for λ=1.0. The addition of hydrogen
strongly influence the propagation of the flame and the flame expands in the combustion
chamber before the piston been in the expansion stroke. Therefore, this shows the necessity of
spark advance for the pure syngas.
77

It is noticeable that the image processing is only capable to determine the size of the
flame after 3% of the total area or 8 mm in terms of flame equivalent diameter. Because syngas
has rapidly increase of the flame size, in the first angle after spark it was possible to find a
measured value. However, for methane 2 CAD need to be discarded. This problem is associated
with spark plug electrode and spark glow, also described for several authors [66-69].
Nevertheless, this method allows the study of the first phase of the flame growth, between 0-
10% of MFB. In the case of thermodynamic analysis for methane, as an example, only after 8
CAD ASOS was possible to have values of MFB (Figure 22). This enhances the potential of
optical techniques as a tool to analyze flame propagation in SI engines.
Figure 30 and Figure 31 shows the flame propagation measurements in lean conditions.
Although the same trend between gases was found, methane has higher decay in terms of
maximum flame velocity (40% from λ=1.0 to 1.4). In addition, the leanest case showed more
stable speed with the movement of the piston for methane, with not a clear maximum peak with
respect to syngas mixtures. Figure 32 shows the maximum flame propagation speed for all the
operative condition of this section.

(a) (b)

Figure 29 – Flame front evolution for M, S50, and S75 at λ=1.0 and SA = -7 CAD ATDC
obtained by averaged data over 25 consecutive engine cycles: (a) Normalized flame area
(b) Flame propagation speed.
78

(a) (b)

Figure 30 – Flame front evolution for M, S50, and S75 at λ=1.2 and SA = 7 CAD BTDC
obtained by averaged data over 25 consecutive engine cycles: (a) Normalized flame area
(b) Flame propagation speed.

(a) (b)

Figure 31 – Flame front evolution for M, S50, and S75 at λ=1.4 and SA = 7 CAD BTDC
obtained by averaged data over 25 consecutive engine cycles: (a) Normalized flame area
(b) Flame propagation speed.
79

40
M S50 S75

Maximum Flame
31.6
30

Speed [m/s]
25.7 26.3
21.1 22.1
20 16.9
12.6
10.8
10 7.7

0
λ=1.0 λ=1.2 λ=1.4

Figure 32 – Maximum Flame Speed Propagation for M, S50, and S75 with λ range from 1.0
to 1.4 and SA = 7 CAD BTDC.

To obtain more details on flame morphology, the Heywood Factor was evaluated in
each operative condition. Results reported in Figure 33 for λ=1.0 and SA=7 CA BTDC suggest,
that for syngas combustion the flame propagation spread more uniformly in all direction. On
the other hand, in the case of methane, the simultaneous action of the flow field and low flame
speed propagation determined an increase in the flame distortion. Syngas for stoichiometric air-
fuel ratio shows similar values of distortion with a shift from the left when H2 content increase
due to higher flame propagation speed. Figure 34 presents the lean cases, in which methane
maintain the higher flame distortion than syngas mixtures. In the case of S50 and S75, a slight
increase for S75 in λ=1.2 was measured, however for λ=1.4 the addition of H2 improve the
flame shape. Figure 35 shows a resume of all cases tested.
80

Figure 33 – HCF for M, S50 and S75 at λ=1.0 and SA = 7 CAD BTDC obtained by averaged
data over 25 consecutive engine cycles.

(a) (b)

Figure 34 – HCF obtained by averaged data over 25 consecutive engine cycles for λ range
from 1.0 to 1.4 and SA = -7 CAD ATDC fueled with (a) S50 and (b) S75.
81

1.6
M S50 S75
1.45

Max imum HCF


1.39 1.41
1.4

1.25
1.19
1.2 1.16 1.15
1.10 1.11

1.0
λ=1.0 λ=1.2 λ=1.4
Figure 35 – Maximum HCF for M, S50, and S75 with λ range from 1.0 to 1.4 and
SA = - 7 CAD ATDC.

Finally, the flame center position was measured as an average of 25 consecutive cycles
(Figure 36) for the analysis of the three fuels and λ ratio. In all conditions, a displacement of
the intake valves from the spark plug electrode was detected. Preferential propagation is
associated with fluid-dynamic phenomena (swirl and tumble) and different temperature profile
[69]. In general, the in-cylinder large rotating scale are characterized into two categories: the
swirl motion rotating about the cylinder axis and the tumble motion rotating about the diametral
axis. It is becoming clear that generating a significant vorticial flow motion inside the engine
cylinder (swirl and/or tumble motions) during the intake process is one of the more promising
ways to obtain high turbulence intensity eventually in the late stroke of compression and
achieve a fast burning rate . A well-defined swirl and/or tumble flow structure is more stable
than other large scale in-cylinder flows and, therefore, it may break up later in the cycle, giving
higher turbulence during combustion. The tumble effect can be directly related to the fact that
most of the air enters the cylinder via the over-flow area of the intake valves, thus creating a
large-scale motion that tends to displace the flame towards the exhaust valves. The temperature
gradient is explained due to exhaust valves have higher temperature profile than intake valves,
due to the absence of refrigeration by the gases. In this case, the first phenomena seem to be
stronger than the temperature distribution. Figure 37 shows the maximum distance reach in the
vertical direction (Y) in terms of distance from the spark plug electrode. The results were
graphed in absolute values and not relative, therefore the zero is set in the center of the
combustion chamber. The addition of H2 decreases the flame movement with almost a
movement from the spark plug to the center of the combustion chamber for S75.
82

(a) (b)

Figure 36 – Mean flame center movement for M, S50 and S75 at SA = -7 CAD ATDC in (a)
λ=1.0 and (b) λ=1.4.

5.0
M S50 S75
Miaximum distance

3.0
in Y [mm]

1.0

-1.0 -0.1 -0.3 -0.6


-0.7
-1.2
-3.0 -2.1

-5.0 -3.7 -3.9


-5.7
-7.0
λ=1.0 λ=1.2 λ=1.4

Figure 37 – Maximum distance from the center of the combustion chamber for M, S50 and
S75 with λ range from 1.0 to 1.4 and SA = 7 CAD BTDC.
83

3.3 Extreme lean conditions

Among all fuel candidates, hydrogen is generally believed to be a promising alternative,


with significant potential for a wide range of operating conditions. In this study, a comparison
was carried out between CH4, two CH4/H2 blends and two mixtures of CO and H2, the last one
taken as a reference composition representative of syngas. It is imperative to fully understand
and characterize how these fuels behave in various conditions. In particular, a deep knowledge
of how hydrogen concentrations affect the combustion process is necessary, given that it
represents a fundamental issue for the optimization of internal combustion engines. To this aim,
flame morphology and combustion stability were studied in an SI engine under lean and
extreme lean burn conditions.

3.3.1 Operating Points

The same spark timing allows studying the flame front propagation in the same fluid-
dynamic conditions and with this maintain roughly the same fluid dynamic conditions (swirl,
tumble, turbulence intensity, among others). In this section, spark timing was set according to
the maximum brake torque (MBT) of the baseline case. Different from the previous section the
baseline case was set for pure methane but λ=1.4. Therefore, the value measured as optimum
spark timing was 13 CAD BTDC.
The excess air ratio was raised from 1.4 to values close to the flammability limit for
each fuel. In this work, the definition of lean limit was set up to the point that the combustion
process became unstable. This effect was measured with the CovIMEP in MBT conditions. It
is important to note that only for the evaluation of the flammability limit spark timing was
modified. The lean limit was set as the λ value for which the CovIMEP parameter exceeded
3%. Methane presented a limit value of λ=1.4; with the addition of 25% (M25) of hydrogen,
this value increased to λ = 1.45 and with 50% (M50) of H2, it was λ = 1.5. In the case of syngas,
the lean limit for S50 was λ = 1.7 and for S75 it reached λ = 1.75. Therefore, the increase of
hydrogen content allowed higher dilution rates, as expected; also, the effect was more evident
when mixed with CO compared to CH4.
Table 10 shows the operative condition for the five blends tested. One interesting point
is that the addition of hydrogen to the methane increase the time duration of the injection.
84

However, this trend is reverted for syngas. Also, the increase in H2 content in CO decrease the
duration of injection as can be seen in Table 10.

Table 10 – Operating conditions set to study the effect of relative air-fuel ratio from lean to
extremely lean conditions in syngas combustion process. Engine speed: 900 RPM; SA: 13
CAD BTDC; Load: WOT.

Fuel λ DOI [CAD]


M 1.40 41
1.40 42
M25
1.45 38
1.40 45
M50
1.50 36
1.40 155
S50 1.60 110
1.70 90
1.40 115
S75 1.60 85
1.75 55

3.3.2 Thermodynamic Results

An initial study of the combustion process was performed through the thermodynamic
approach. In-cylinder pressure was analyzed as average traces of 200 consecutive acquisitions
for each case; results are reported in Figure 38 and Figure 39; motored pressure signals (dashed
line) are also shown as a reference. Figure 38 shows the in-cylinder pressure for the five fuels
for λ =1.4; syngas mixtures shows high peak pressure values in the initial phase and then a rapid
decrease in both hydrogen concentrations. Instead, methane and its blends with hydrogen show
low peaks, but higher-pressure values in late phase of the combustion process than pure syngas
blends. The addition of hydrogen to methane increased peak pressure, as well as advancing its
position with respect to the TDC; the effect was stronger in the case of 50% of addition.
Figure 39 shows the pressure curves for extreme lean condition conditions; it is
noticeable that syngas presents an extended range of operation with respect to methane. Also,
the addition of hydrogen extended the operative range when blended with CO. This can be
directly related to the actual mass participation of hydrogen that is roughly double for syngas
compared to the methane blends and the higher flammability range of CO with respect to CH4.
All fuels near the flammability limit presented low peak pressure, with S50 showing the highest
value; this is due to lower air dilution of the mixture with respect to S75. One interesting point
85

is that methane blends presented higher-pressure values in the last combustion phase, similar to
the cases of λ=1.4.
Table 11 shows the main thermodynamic parameters; it can be noted that methane and
methane-hydrogen mixtures showed similar IMEP values for equal air-fuel ratio, and the effect
of higher-pressure values during the expansion stroke resulted in slightly higher engine output
compared to syngas mixtures. Another important observation from Table 11 is the effect of
hydrogen addition in the cycle-by-cycle variation; with improvements for methane blends, even
if not spark optimization was set. For syngas, the CovIMEP was low with similar values for
both mixtures, and under methane and methane-H2.

a) b)
Figure 38 – In-cylinder pressure traces averaged over 200 consecutive cycles for λ=1.4
a) methane-H2 blends and b) pure syngas blends compare with methane.
86

(a) (b)
Figure 39 – In-cylinder pressure traces averaged over 200 consecutive cycles for extreme lean
condition a) methane-H2 blends and b) pure syngas blends compare with methane.

Table 11 – Main thermodynamic parameters extracted from pressure traces for lean to
extremely lean conditions at fixed spark timing. Engine speed: 900 RPM; SA: 13 CAD
BTDC; Load: WOT.

DOI IMEP CovIMEP Pmax mfuel ηfuel


Fuel λ
[CAD] [bar] [%] [bar] [mg/cycle] [%]

M 1.40 41 3.9 3.3 16.4 18.6 19.9


1.40 42 3.7 2.4 18.9 17.6 18.9
M25
1.45 38 3.3 4.9 15.3 17.2 17.2
1.40 45 3.8 2.5 22.6 16.2 19.3
M50
1.50 36 2.9 2.8 16.6 15.2 15.7
1.40 155 3.5 0.7 29.3 57.0 16.7
S50
1.70 90 2.9 2.5 17.1 40.3 19.5
1.40 115 3.2 0.7 30.8 32.6 15.7
S75
1.75 55 1.7 2.3 15.9 27.6 9.9

A more detailed analysis of the averaged pressure traces was performed using the heat
release approach explain in the previous section. The MFB traces at the same relative air-fuel
ratio (λ=1.4) shows in Figure 40 emphasize the fact that syngas mixtures present a faster initial
combustion phase than methane and CH4-H2 blends. As was expected, with the increase of
hydrogen addition this effect was more prominent. In addition, for the same H2 concentration,
CO presented shorter duration of the initial kernel phase and main combustion phase with
87

respect to CH4. All cases featured quite close evolutions during the final stages of the process,
the mass transfer from the top-land region to the cylinder controlled this part of the process.
In the case of λ near lean flammability limit, the propagation speed was closer for all
fuel types. The observed trends are also reflected in the crank angle durations (Table 12),
calculated at given mass fraction burned levels, i.e. 5%, 10%, and 50%, as representative for
the kernel stage and flame propagation phase.

(a) (b)
Figure 40 – Mass fraction burned traces averaged over 200 consecutive cycles for a) λ=1.4
and b) lean extreme case.

Table 12 – Crank angle durations from the start of spark to fixed mass fraction burned
thresholds calculated from pressure traces for lean to extremely lean conditions at fixed spark
timing. Engine speed: 900 RPM; SA: 13 CAD BTDC; Load: WOT.

5% 10% 50%
Fuel λ
[CAD ASOS] [CAD ASOS] [CAD ASOS]
M 1.40 18 23 47
1.40 16 21 39
M25
1.45 18 24 50
1.40 13 17 32
M50
1.50 17 21 47
1.40 8 9 23
S50
1.70 14 19 43
1.40 6 8 22
S75
1.75 14 20 46

These results emphasize the much stronger effect of hydrogen addition to carbon
monoxide compared to the case of methane; e.g. it resulted in a shorter 0-5% MFB interval,
88

with a difference of 5 CAD at the same concentration of 50% hydrogen. When looking at these
findings, it should be noted that, as previously mentioned, mass concentration in the air-fuel
mixture is roughly double in the case of syngas; given that a close-to-linear correlation was
found between the mass concentration of hydrogen and laminar flame speed of its blends with
methane [98], it can be stated that the noted effect is directly linked to this basic fuel property.
The MFB position also gives an idea of re-calibration procedures that can be considered for
adapting ignition settings for the alternative fuels [99] and feedback strategies for real-time
engine control.
The analysis of exhaust gas measurements gave more information on the chemistry of
combustion for the investigated conditions (Figure 41 and Figure 42). Carbon monoxide (CO)
concentrations were low for λ=1.4, with only syngas mixtures above 0.1%. In the case of
extreme lean condition, methane blends maintain the same behavior. However, syngas increases
the CO emissions due to instabilities that increase the amount of fuel not burned, with a
maximum in S50 of 0.9%. This same effect was also seen in methane blends but with a CH4
concentration in the exhaust emission. It is noticeable that for optical engine due to the top land
region a large amount of fuel reminds without being burned. Therefore, the concentration that
will be obtained in the exhaust depends on the burned mixture. For syngas, this high
concentration was for CO and for methane blends CH4 as was expected.
Figure 42 shows the NOx concentration in which syngas blends for λ=1.4 shows the
highest values. Meanwhile, methane blends for both air/fuel ratios depict almost negligible
values. Because the NOx emissions have a strong dependence on temperature (what also could
be associated with maximum pressure in the chamber), this graph has a similar trend that was
seen in Table 11 with respect to maximum pressure values. For λ=1.4 syngas blends have 1.5
times the maximum pressure of methane blends, this causes an increase of emission from 15
ppm to 4000 ppm approximately. When syngas reaches similar pressures in the case of extreme
lean condition, the NOx values became similar with almost negligible values.
89

25000

20000
17043

CO [ppm] 15000

10000 7502

5000
1943 1311
665 690 882 798 884
0
M M25 M50 S50 S75 M25 M50 S50 S75
λ=1.4 Extreme lean condition

25000
20044
20000

15000 13611 13807


CH4 [ppm]

10310
10000 8818

5000
18 21 13 16
0
M M25 M50 S50 S75 M25 M50 S50 S75
λ=1.4 Extreme lean condition

Figure 41 – CO (top) and CH4 (bottom) concentration in ppm on the exhaust port for λ=1.4
and extreme lean condition.

500
450
400 378

350
300
NOX [ppm]

250
201
200
150
100
50 14 26
3 5 3 1 2
0
M M25 M50 S50 S75 M25 M50 S50 S75
λ=1.4 Extreme lean condition

Figure 42 – NOx concentration in ppm on the exhaust port for λ=1.4 and extreme lean
conditions.
90

3.3.3 Optical Investigations

In order to obtain detailed results on the time evolution and spatial distribution of the
burned mass and flame front, in-cylinder cycle resolved visualizations were carried out. Table
13 shows the main camera set up used for the comparison of the five fuels. Figure 43 and Figure
44 shows selections of images detected during engine cycles recorded in lean burn conditions
(λ=1.4) and at the flammability limit respectively, for all fuels. The images were selected at the
equal flame area with respect to the piston cross-section; due to the different burning speeds,
the delay from ignition resulted differently. In agreement with the literature [61], the flame
kernel was well resolvable only around 3 CAD ASOS in all conditions, due to the very high
luminosity of spark-induced plasma.

Table 13 - Camera set up for imaging acquisition set for study the effect of air-fuel
ratio from lean to extremely lean conditions in syngas combustion process at fixed spark
timing. F-number: 5.6, Intensification: 55000.
Fuel λ Exposure time
[CAD]
M 1.40 1.0
1.40
M25 1.0
1.45
1.40
M50 1.0
1.50
1.40
S50 0.5
1.70
1.40
S75 0.5
1.75
91

Figure 43 – Flame image sequence for all fuels with λ=1.4 at 5%, 10%, 20%, 30%, 40% and
60% area of the piston cross-section with exposure time 1.0 CAD for methane blends
and 0.5 CAD for syngas.

Figure 44 – Flame image sequence for all fuels in extreme lean conditions, at 5%, 10%, 20%,
30%, 40% and 60% area of the piston cross-section with exposure time 1.0 CAD for
methane blends and 0.5 CAD for syngas.
92

A preliminary analysis of results shows in Figure 43 permits to observe that the overall
luminous intensity of syngas flames was stronger than those with methane, in spite of the lower
exposure time. Emission intensity increased with the H2 content in the fuel mixture (at fixed
air-fuel ratio) with higher values for syngas. Moreover, different shape and structures of the
flame front were observed between methane and the CO-containing blends. Figure 44 shows
flame sequences for one selected cycle in the extreme lean cases for the fuels with hydrogen
addition. As it can be observed, the luminous thickness decreased at increasing λ values,
because of the reduction of the reaction zone width.
It should be noted that the presence of CO compared to methane in the 50% hydrogen
mixture not only allowed reaching more diluted conditions (λ=1.7 instead of 1.5), as previously
discussed but determined comparable luminosity, in spite of the reduced exposure time fixed
for syngas acquisitions. This was due to the stronger radiative efficiency of exothermal
reactions that featured carbon monoxide oxidation, with respect to methane in the UV-visible
spectral range. The effect was already evident for richer condition (λ=1.4), as was shown in
Figure 43.
By applying the image processing procedure, the trends of the mean area and
propagation speed of the flame were obtained. Results reported in Figure 45 and Figure 46 are
related to the averaged values over 25 consecutive engine cycles. To put these graphs in the
context of earlier discussions, for λ=1.4, the time it took the flame to reach the optical limit
decreased at higher hydrogen percentage. Due to the higher mass content of H 2 in the syngas-
air mixture, the related flames resulted faster than those induced by methane blends. The
expected improvement in flame propagation speed when switching from S50 to S75 was in part
reduced by the slight increase in the cyclic variability. For extreme lean conditions, the effect
of hydrogen was quite evident in improving flame propagation; of course, no direct comparison
can be performed, given the different air-fuel ratios.
Nonetheless, the influence of H2 addition was important for both types of basis fuel (i.e.
methane and CO); the influence of hydrogen on combustion duration was more significant for
syngas fueling, especially during the initial burning phase.
93

(a) (b)
Figure 45 – Evolution of a) flame front area and b) flame propagation speed for λ=1.4
obtained by averaged data over 25 consecutive engine cycles.

(a) (b)
Figure 46 – Evolution of a) flame front area and b) flame propagation speed for extreme lean
cases obtained by averaged data over 25 consecutive engine cycles.

The effect of hydrogen addition and air dilution in terms of flame deformation was
evaluated for syngas and CH4 through the evolution HCF. At each CAD, the average value over
25 consecutive engine cycles was considered. Results reported in Figure 47-a with the same
air-fuel ratio clearly demonstrate that syngas showed lower flame deformation compared to the
methane blends. The latter resulted less ‘circular’; however, the increase in hydrogen content
decreased the overall deformation of the combustion process. In extreme lean burn conditions
94

(Figure 47-b), the simultaneous action of the flow field and fuel charge distribution determined
a strong increase in the flame distortion in the early combustion stages.

(a) (b)
Figure 47 – Heywood circularity factor of the flame for (a) λ=1.4 and (b) extreme lean cases
obtained by averaged data over 25 consecutive engine cycles.

In order to obtain an estimation of the effect of selected operative conditions on the


macroscopic flame front deviation from the ideal circular and spark plug centered propagation,
the trajectory of the luminous flame centroid was followed. Results referred to averaged data
over 25 consecutive engine cycles, are shown in Figure 48. In all conditions, flame
displacement towards the intake valves region was observed. This reached comparable
maximum distance in the lean burn conditions for the methane-containing mixtures.
Flame displacement was partially due to fluid motion, induced by tumble [69], and to
the different thermal regime between the intake and exhaust side of the combustion chamber
[100]. The reduced effect of tumble motion on the syngas flame centroid was due to the faster
flame propagation in the early stage of combustion determined by the higher hydrogen content.
The effect along the x-direction was negligible, as a consequence of low swirl motion.
For better understanding flame displacement, Figure 49 shows the maximum distance
from the combustion center in the Y direction. It can be noted that syngas mixtures presented a
position closer to the combustion center and spark plug for both conditions, compared to
methane and methane blends. However, the results demonstrated that the hydrogen content
95

increasing determined a decreased flame displacement for both fuels (CH4 and CO). Moreover,
an increase in flame displacement at high air dilution was observed for all fuels.

(a) (b)
Figure 48 – Flame centroid position for (a) λ=1.4 and (b) extreme lean cases, obtained by
averaged data over 25 consecutive engine cycles.

4.0 M M25 M50 S50 S75


Max Center Displacement in Y [mm]

2.0

0.0

-2.0 -1.3
-1.8
-4.0

-6.0 -4.8 -4.8 -5.1


-5.9
-8.0 -7.1 -6.9 -7.0

-10.0
Lambda 1.4 Flamability Limit

Figure 49 – Maximum flame displacement in vertical axes in intake direction.


96

3.4 Summary

An experimental study was undertaken in order to evaluate the effects of air-fuel ratio
and hydrogen addition to syngas blends, on combustion in SI power units. The results were
compared with methane and methane-hydrogen blends. To this aim, an optically accessible
engine was running at 900 RPM, WOT, and fixed spark timing, with thermodynamic and
optical analysis of the combustion process. The fuels used was: methane, 25% and 50% of
hydrogen addition to methane, and finally two pure syngas equivalent mixtures, with 50% and
75% hydrogen content. The study was divided into two section, one focused stoichiometric to
lean mixture of syngas and the second section was concentrated on the study of lean and
extreme lean conditions with all the fuels.
As resume extracted from the first section, the highest in-cylinder pressure was
measured from syngas mixtures in all operative condition. The low decrease in pressure traces
for S50 and S75 compare to pure methane was seen in λ=1.4. The syngas presented higher fuel
conversion efficiency and IMEP value than methane in a lean condition. However, the baseline
case presents better performance and combustion stability for the stoichiometric air-fuel ratio.
This can be directly associated to not spark timing optimization for syngas cases.
The optical analysis provides interesting information in terms of flame development and
morphology. Due to maintaining spark timing fix, the fuels can be analyzed with similar fluid-
motion (tumble, swirl and piston position). A large difference was seen in terms of flame
propagation speed with S75 32 m/s, S50 26 m/s, and M 13m/s, for λ=1.0. In the case of lean
mixtures, pure methane suffers higher decrease than syngas with values in turn of S75 22 m/s,
S50 17 m/s, and M 8m/s. Flame distortion was found higher for methane than syngas for all
cases. This is directly associated with flame propagation speed and hydrogen addition. Finally,
center movement from spark plug electrode was calculated for all test condition. Syngas present
lower displacement than methane (maximum of 6.0 mm) with a peak value 0.5 mm for S75 and
2.0 mm for S50.
An increase of fuels and operative condition was performed in the second section. The
overall thermodynamic analysis revealed that hydrogen addition improved the flammability
range of CH4, with a maximum λ of 1.50 for 50% H2. For syngas mixtures, this range was larger
still, with a λ of 1.70 for 50% H2 and λ 1.75 for 75% H2. For λ settings that were tested (i.e.
λ=1.4 and lean burn limit), the methane blends presented higher pressure values in the last
combustion phase, while syngas featured higher peak pressure. The five fuels that were tested
97

ensured acceptable engine performance, even if the spark timing setting was kept at the MBT
point for methane (with λ=1.4). One major conclusion of the analysis was that spark timing re-
calibration is required in order to fully take advantage of fuel properties such as higher laminar
flame speed and increased stability.
Cycle-resolved flame imaging allowed a more detailed insight into the fuel oxidation
processes with regard to its evolution within the combustion chamber. In line with the
thermodynamic results, the mixture with the highest concentration of hydrogen S75 featured
the fastest flame propagation. Propagation was found to be quite symmetric in all directions,
and similar for all fuels, suggesting that fluid motion had the most important effect in this sense.
These results were also confirmed by the evolution of flame centroid displacement, more
extensive for extreme lean air-fuel ratios, but comparable nonetheless.
98

4 Effect of Spark Timing in Syngas Combustion Process

Increased research efforts are aimed at developing control strategies in order to increase
thermal efficiency and mitigate the tendency toward knock by using cooled exhaust gas
recirculation [101] or water injection [102], reduce cyclic variability [103], provide flexibility
in the choice of fuel [98] among others. The spark timing is one of the basic parameters that
control SI engine, and with a correct selection, the performance and efficiency could be
improved. This selection depends on air-fuel ratio, fuel composition as well as crankshaft
rotational velocity. The study of optimal set up of this parameter also plays an important part,
especially in improving combustion stability, extending lean limits of operation [101] and using
specific features to provide additional control on the process [104]. The optimization of the
early stages of ignition processes represents a substantial part of guiding the development of
new generation SI engines [55]. In fact, the ignition phase and flame kernel formation strongly
influence the in-cylinder flame propagation [68].
Recently, advanced ignition systems with higher spark energies have been developed in
order to reduce cyclic variability and thus enhance combustion efficiency. Among these
systems, plasma-assisted ignition (PAI) has proven to improve ignition characteristics and
flame stability under several conditions [55]. However, the conventional spark plug is the most
used by the manufactured to the equipped internal combustion engine. Therefore, for syngas
application is crucial to understand the effect of different moments that combustion process
starts in the flame growth. The principal factors are the piston position, the pressure at the point
of the energy discharge and variation in the fluid-motion in the chamber (tumble and swirl).
Nowadays an elevated number of engines are equipped with Port Fuel Injection Spark
Ignition (PFI SI) engines. Their technological level is high, but a further optimization is still
possible, especially at low engine speed and high load. To this purpose, the scientific
community is focused on deepening the understanding of thermo-fluid dynamic phenomena
that takes place in this kind of engine: the final purpose is to find key points for the reduction
in engine specific fuel consumption and exhaust emissions without a decrease in performance.
The traditional approach to engine optimization based on the trial and error method has
proved to be expensive and inefficient, the new approach is strictly related to the basic
knowledge of the complex thermo-fluid dynamic phenomena processes occurring in engines
[67]. A better understanding of phenomena connected to the combustion process that can
99

contribute to the reduction in specific fuel consumption and exhaust emissions is necessary.
Gas composition can have a significant impact on the performance and emission characteristics
of an engine by affecting the flame speed, ignition, and knock characteristics and therefore, the
spark timing is required to be optimized carefully for improved engine performance [52].
The objective of the current chapter is to investigate the effect of spark timing in the
performance and their relationship with the flame front propagation in terms of flame speed and
distortion when the engine is fueled with equivalent syngas mixtures. Therefore, this study
expands the understanding of different initiation of the flame front in terms of time and piston
position, which induces different fluid-dynamic condition for the flame growth. The proposed
study can provide fundamental data to validate CFD models for predictive analysis and further
optimization.

4.1 Stoichiometric condition

4.1.1 Operating Points

In this chapter, the combustion process changing spark timing was investigated in the
same optically accessible single cylinder PFI SI engine that was presented in chapter 3. In
addition, the same techniques (section 3.1) was used with thermodynamic and optical analysis.
As fuel was used syngas with 50% H2 and 50% CO in volume (S50) as the most representing
condition of the biomass gasification process [14]. The main fuel properties were presented in
Table 7. The other engine parameters were fixed in the same values of the last chapter as could
be seen in Table 14, with low engine speed and maximum load as a characteristic operative
point of electric generator engines. In addition, the λ value was fixed at the stoichiometric point
to be able to work only with the influence of ignition time.
The purpose of this study is to verify the combustion behavior of pure syngas when the
ignition time is varied. Therefore, a mapping of the engine was performed with the study of
several values. The condition tested contains SA from 7 CAD BTDC (optimum condition for
methane in λ=1.0) up to positives values of spark timing (22 CAD ATDC). Moreover, the MBT
condition with a value of 8 CAD ATDC was reached, as could be seen in Figure 50. It is
noticeable that 15 CAD of SA difference exists between methane and S50. This observation is
in agreement with others investigations, in which large difference was found between syngas
and methane in terms of crank angle to reach maximum performance. Hagos et al. [33] found a
100

difference of 20 CAD between S50 and methane at 1500 RPM for lean mixtures. Mustafi et al.
[52] found that the MBT at 1000 RPM was 4 CAD BTDC for S50 and 24 CAD BTDC for
methane.
From Figure 50, it can be stated that the maximum efficiency (17.4%) was achieved in
the MBT condition (due to same fuel injection in all cases). As was commented in chapter 3,
the efficiency in this type of research engine has to be taken as comparative because the low
compression ratio and combined with large top land region represent a decrease in terms of
performance. The engine stability was found in accepted values below 3% for all test
conditions. A similar observation was reported with other types of syngas in SI engines by
various authors [42, 55]. This was attributed to the combined effect of the higher flame-
propagation nature of H2 and the flame suppressive nature of CO; the latter played a
combustion-controlling role [33].

Figure 50 – Performance parameters (IMEP, CovIMEP, and fuel conversion efficiency) at


stoichiometric air-fuel ratio for S50 and several spark timing.

4.1.2 Thermodynamic Results

Six different spark timing, firstly tested in the mapping study, were analyzed in detail
with thermodynamic and optical tools (Table 14). This range of study includes several fluid-
101

dynamic conditions in which the flame is exposed, with two of them in the compression stroke
and the others four in the expansion stroke.

Table 14 – Operating conditions set for study the effect of spark timing in stoichiometric air-
fuel ratio fueled with S50. Engine speed: 900 RPM; λ=1.0; Load: WOT; DOI= 204 CAD.
Fuel SA
[CAD BTDC]
-7
-4
+0
S50
+4
+8
+18

The mean in-cylinder pressure traces versus crank angle are shown in Figure 51-a. The
motored pressure was also illustrated (point line) as a reference. As could be seen, with the
retard of spark advance (positive SA) a decrease of maximum pressure was registered (as
expected due to the piston moving down). However, advance spark timing cases shows a rapid
decrease in the latest phase of the expansion stroke. This behavior explains the IMEP trend, in
which the graph increase with the retard of spark timing up to reach the MBT (Figure 50). These
results are against mapping seen for traditional hydrocarbon fuels, in which MBT is always
seen in spark timing set at compression stroke. One possible explanation could be the high
laminar flame speed of S50 (see Table 7), and effects of engine geometry (low compression
ratio and top land region). The extreme retard cases (SA above 8 CAD ATDC) shows a decrease
in IMEP due to the low maximum peak pressure.
A complementary analysis was performed with in-cylinder pressure against normalized
cylinder volume (Figure 51-b), to have more information of the combustion process with
respect to the piston movement. Is evident the difference between the different spark timing, in
which negative values shows a rapid increase of the pressure in TDC (zero volume). Moreover,
MBT condition showed a shift to the right in the peak value and a slightly lower decrease in the
last combustion phase. Therefore, the optimization of spark timing set the best pressure
configuration to reach the highest performance.
102

a) b)

Figure 51 – In-cylinder pressure traces for stoichiometric fueling of syngas 50% H2 and
several spark timing versus a) crank angle degree and b) normalized cylinder volume

Figure 52 shows a comparison between peak in-cylinder pressure and NOx emission for
several spark timing. The aforementioned decrease in peak pressure is evident when retarding
the spark timing. In addition, the NOx emission presented a similar behavior, decreasing along
with the pressure. This is due to the dependence of NOx emission with the in-cylinder
temperature profile. As could be seen in the bibliography [34], the temperature increase with
the pressure in SI engines; therefore, the NOx also have a direct relationship to pressure in this
type of applications. For spark timing in the compression stroke, this effect is higher with 4300
ppm of peak in 7 CAD BTDC (Pmax 34 bar). The other pollutant emission presented stable
behavior with variation below 5% between each case and without any clear tendency.
103

Figure 52 - Comparison of maximum in-cylinder pressure and NOx emission for several spark
timings.

A more detailed analysis of the averaged pressure traces was performed using the heat
release approach, based on the first law of thermodynamics (see section 2.1), with sub-models
for calculating heat transfer rates and blow-by losses [76, 87]. The MFB traces at the same
relative air-fuel ratio (λ=1.0) are shown in Figure 53, the left part shows in terms of position
from TDC and right-side with respect to the start of the combustion process. From Figure 53-
a, could be seen that the combustion process was growth in total different piston position (CAD)
as was expected due to the different ignition time ignition (SA). As an example, for spark
advance 7 CAD BTDC the 40% of MFB have the same position that the start of the combustion
process for 4 CAD ATDC. This graph was depicted to readers have an idea of the differences
existing in the combustion process study in this section.
Figure 53-b shows that when changing the horizontal axes from absolute to a relative
one a better comparison can be performed. The initial phase of the combustion process shows
an increase in burn speed when spark timing was advanced with the faster in 7 CAD BTDC.
All cases featured quite close evolutions during the final stages of the process, as this part of
the process was controlled by the mass transfer from the top-land region to the cylinder.
104

(a) (b)
Figure 53 – Mass fraction burned to stoichiometric air-fuel ratio with S50 and several sparks
timing with respect to: (a) top dead center (TDC) and (b) start of spark (SA).

Figure 54 – Zoom view of mass fraction burned for stoichiometric air-fuel ratio with S50 and

several sparks timing with respect to the start of spark (SOS).

To have a better analysis of the MFB variation between cases, a zoom view was
illustrated in Figure 54. This graph emphasizes the fact that the advanced of spark timing
produces a faster initial combustion phase. However, an important decrease of MFB was seen
only for 18 CAD ATDC. For MFB 5%, all cases were close to 5 CAD ASOS except the last
105

one with 6 CAD. For 10% of MFB, the difference increase between -7 CAD ATDC and MBT
condition (+8 CAD ATDC) with 7 CAD. This enhances the rapid burning properties of
hydrogen that reduce the SA effect in the propagation of the flame.

4.1.3 Optical Investigation

As a complement of thermodynamic analysis, images from 25 consecutive cycles was


taken in order to study the flame front propagation in the combustion chamber for the six spark
timings showed in Table 14. The procedure for post-processing was explained in chapter 3;
therefore, flame area, propagation speed, distortion, and displacement from the spark plug was
calculated as an average value in each CAD. The main camera set up parameters were fixed
for all cases, due to similar intensity emission. The F-number was set at 32, exposure time equal
to 0.5 CAD and intensification in the minimum recommended by the intensifier fabricant
(55000).
Image of the combustion process at the same CAD after spark timing from one
representative cycle was depicted in Figure 55. It could be observed a slight decrease in terms
of flame size and light emission when the spark timing is retarded (SA increase). Both effects
are more visible in the cases of positive SA. The emission intensity variation could be associated
with differences in combustion chamber temperature and pressure. Generally, emission
increase with pressure and temperature. However, a spectroscopy analysis could give more
information about chemical species present in the flame and the wavelength of emission.
106

Figure 55 – Images of the flame at the same CAD after the start of the spark of Syngas with
50% of H2 in λ=1.0 for different spark timing. F-number: 32, exposure time: 0.5 CAD and
intensification: 55000.

The mean area and flame propagation speed were calculated and plotted in Figure 56.
For all cases, the flame reaches the entire optical window before 10 CAD ASOS. Also, this
phenomenon could be seen in Figure 56-b in which the maximum flame speed suffer a decay
from 7 CAD ASOS due to interaction with the walls. As was seen directly in the images (Figure
55), effectively the flame area growth decrease with the retard of the spark timing. This is in
agreement with MFB graph (Figure 54), that extract the flame burn duration from pressure
107

traces. For the optimum condition of S50 (8 CAD BTDC) with respect to extreme advances
case (7 CAD BTDC), a variation of 1 CAD was measured in 30% of the piston area. This
difference was maintained up to the optical limit. Flame front propagation speed was depicted
in Figure 56-b, the curve starts from relatively high values, reach the peak and then decrease
steadily due to wall interception for all spark timings. In this case (S50, λ=1.0, SAMBT = 8 CAD
ATDC) the initial measured speed is over 15 m/s.

(a) (b)

Figure 56 – Evolution of the flame front area and flame propagation speed for S50 and several
spark timings by averaged data over 30 consecutive engine cycles.

Figure 57 details the maximum flame speed propagation for each case. As was
commented, a decrease of speed with the retard of spark timing was measured with a variation
of 15% between the extremely advanced case (MBT of methane) and MBT condition for S50.
When compared with Figure 52, it could be extracted that the three parameters (pressure, NOx
and flame propagation speed) decrease with SA.
When looks to the laminar flame speed of gaseous fuels, it is well known that increase
with the temperature and decay with the pressure. As an example, Fanelli et al. [105] performed
a laminar flame study with S50, in several temperature and pressure condition. In this study,
was found an increase from 100 cm/s to 500 cm/s when the temperature increase from 300K to
700K at 1 atm. For pressure tests, the results showed a decrease from 100 cm/s to 35 cm/s when
the pressure changes from 1 atm to 30 atm.
In the emission test results, was found with the advance of SA a pressure and
temperature (NOx) increase. Therefore, could be concluded that temperature profile has more
108

influence on peak flame speed than pressure increase for this case of study. However, need to
be taken into account the pressure increase as a suppressor of flame development speeds.

35.0
Maximum Flame Speed [m/s]

30.0 25.7
23.5 23.3 22.7
25.0 21.8
19.5
20.0
15.0
10.0
5.0
0.0
SA -7 SA -4 SA 0 SA +4 SA +8 SA +18

Figure 57 – Maximum flame speed S50 and several spark timings by averaged data over 25
consecutive engine cycles.

Complementary analysis of the flame front evolution in terms of the morphology could
be done with the Heywood Circular Factor (HCF) and the flame front displacement. For all
cases, the deformation is small (Figure 58), in turn of 1.05. For detailed analysis, the maximum
HCF for each SA was graphed in Figure 59. In general, HCF is associated with the influence
of the medium in the flame front. This parameter shows a decrease in the retard of the spark
timing. Therefore, in SA set in the expansion stroke seem to be less distorted than advanced
cases due to piston position in the first phase of the combustion process. To put in context, the
last four cases of study the flame propagate entirely in the expansion stroke (piston going
down), different for the first two cases (piston going up and then down). A computational or
experimental fluid-dynamics study could help to better conclusions in this field. However, the
difference founded was less than 3% for extreme cases.
Moreover, the flame displacement was also calculated. For this case of study, the flame
center goes from the spark electrode to the center of the combustion chamber almost directly
(Figure 60). Therefore, the influence of swirl and tumble motion are negligible in the flame
front for these cases. The maximum distance from the center of the combustion chamber shows
a stable behavior, with almost no difference between the cases (in turns of 0.5 mm to the intake
side). This is due to high concentration of H2 and stoichiometric air-fuel ratio condition that
produce fast burning rates (Figure 53 and Figure 57).
109

Figure 58 – Evolution of average Heywood circularity factor for S50 at several spark timings.

1.20
1.18
1.16
Maximum HCF

1.14
1.12 1.10 1.10
1.10 1.09 1.09
1.08 1.07
1.08
1.06
1.04
1.02
1.00
SA -7 SA -4 SA 0 SA +4 SA +8 SA +18

Figure 59 – Maximum HCF for S50 at several spark timings.

Figure 60 – The average flame center movement for S50 at several spark timings.
110

4.2 Extreme lean conditions

Spark ignition IC engines fueled with gaseous fuels have been shown to be
advantageous over traditional gasoline engines in terms of fuel efficiencies and pollutant
emissions [106]. Further taking into account the vast availability of natural gas, the advances in
gasification process and the renewability of hydrogen, fuels with a high content of H2 are
considered to be a promising alternative fuel for large-scale applications in spark ignition
engines.
The spark timing has an important effect on the performance and the combustion process
of spark ignition engine. The spark is given generally near the end of the compression stroke in
a four-stroke engine at an optimum time that produces the maximum work output from the
engine. If the spark is advanced too much, the pressure builds up too early inside the cylinder.
Therefore, the work transfer from the piston to the gas in the compression stroke increases. On
the other hand, if the ignition timing is retarded beyond the optimum point, the flame
propagation phase proceeds well into the expansion stroke. Therefore, there is a considerable
decrease in the peak cylinder pressure and the peak pressure shifts away from the TDC location.
The optimization of the spark timing is done by balancing these two mutually opposite effects
of advancing and retarding the spark timing [41]. Thus, flame propagation in the cylinder and
the resulting rate of pressure rise has important effects in optimizing the spark location.
The past decade has witnessed significant progress in the research and development of
HCNG (CH4/H2) and syngas (CO/H2) engines. In this section, the main objective is to provide
experimental data of this two type of fuels operating in a PFI-SI engine with optimum spark
timing. Additionally, correlate fuel properties with experimentally observed engine
performances, and optical measurements of the combustion process providing fundamental
insights into the effects of hydrogen addition on engine working processes. To reach this goal,
the study of blends with an intermediate and high fraction of H2 content in the optimum spark
timing for lean and extreme lean condition was performed. As was seen in chapter 1, work in
the lean condition provides advantages in terms of efficiency and pollutant emissions.

4.2.1 Operating points

For that purpose as was studied in section 2.4, an experimental test was performed with
methane, methane-H2 and CO-H2 blends at λ=1.4 and extremely lean. Variation with the
mentioned section was in optimize the spark timing to reach the maximum IMEP (MBT). This
111

parameter was changed in a previous analysis where the MBT condition was found between
several spark timing tries and the results are presented in Table 15. As was presented in the
previous section syngas with 50% H2 shows an MBT condition on the expansion stroke being
away from common use fuels as methane. However, the behavior of others gaseous fuels, as
studied in section 2.4, did not study yet.
The extreme lean condition, as was explained, was set when the combustion process
stability is over 3% of CovIMEP. The hydrogen addition decreases the cycle by cycle variations
when comparing the same air-fuel ratio. Even in small quantities, hydrogen addition can be
used to improve combustion stability, such as 25% (i.e. equivalent to a mass participation of
only around 4%). The effect is higher for syngas, with CovIMEP below 1% in lean condition
(λ=1.4). As was mentioned a similar observation was reported with other types of syngas and
was attributed to two factors: higher flame-propagation nature of H2 blends and flame
suppressive nature of CO.

Table 15 – Operating conditions set for study the effect of spark timing from lean to
extremely lean conditions in syngas combustion process. Engine speed: 900 RPM; Load:
WOT.

Fuel λ SA DOI IMEP CovIMEP


[CAD BTDC] [CAD] [bar] [%]
M 1.40 -13 41 3.9 3.3
[%]
1.40 -11 42 3.8 2.4
M25
1.45 -15 38 3.3 4.9
1.40 -7 45 3.9 2.3
M50
1.50 -11 38 3.1 2.9
1.40 0 155 4.3 0.7
S50 1.60 -4 110 3.5 1.0
1.70 -13 90 2.9 2.5
1.40 +8 115 4.0 0.7
S75 1.60 +4 85 3.4
[bar] 1.0
1.75 -15 55 1.9 2.3

Another interesting point from Table 15 is that syngas mixtures (S50 and S75) have non
negative angles for spark timing when use λ=1.4 as was seen in stoichiometric air-fuel ratio to
S50. This type of behavior was explained in the previous section and for S75 due to lower burn
duration than conventional fuels, this set up is seen up to λ=1.6. However, for the extreme lean
condition the SA change radically, and need to be set in -15 CAD ATDC. For methane blends,
the increase in hydrogen content retards the spark timing in 2 CAD for 25% H2 and 6 CAD for
112

50% H2 with respect to pure methane. In addition, as was commented in section 2.4 the H2
content extend the operative condition.

4.2.2 Thermodynamic Results

The thermodynamic analysis give a first approach of the combustion behavior with
results as maximum in-cylinder pressure, burning duration, pollutant emissions among others.
A custom code was applied to extract more information of the combustion traces like MFB and
fuel conversion efficiency. IC engines, by definition, are machines that combust fuels to achieve
energy conversion and thus mechanical power output. Therefore, it is natural to deduce that fuel
type is an essential factor affecting engine performances [106]. Figure 61 shows the pressure
inside the combustion chamber for the same air-fuel ratio and in extreme lean condition. Due
to optimization of the spark timing, is not clear the tendency in terms of peak pressure as was
seen for the same spark advance (Figure 38).

a) b)
Figure 61 – In-cylinder average pressure curve for 200 cycles at λ=1.4 for: a) methane-
hydrogen blends and b) pure syngas blends compare with pure methane.
113

a) b)
Figure 62 – In-cylinder average pressure curve for 200 cycles at extreme lean condition for: a)
methane-hydrogen blends and b) pure syngas blends compare with pure methane.

S50 shows the higher peak pressure and IMEP with an SA in TDC position for λ=1.4.
Therefore, 50% H2 in CO seems to be the best blend in terms of performance over pure methane
for the lean range. A decrease in 7% of IMEP was seen for 75% H2 in carbon monoxide and the
spark timing need to be retarded in 8 CAD. To put in context, a reduction of 15% in the brake
torque was observed with syngas (S50) compared to NG by Hagos et al [33]. In addition,
reductions of 23% at a compression ratio of 8:1 and 18% at a compression ratio of 11:1 were
reported for 50-50% CO/H2, as compared to NG at 1500 rev/min by Mustafi et al. [52].
However, both studies compare the fuels in different air-fuel ratio due to limitations in the
injection pulse in where syngas was leaner than CNG. In spite of this, could be extracted from
these studies that for the same load value, syngas can operate at leaner values and obtain higher
fuel conversion efficiencies. Therefore, the present work increases the available bibliography
in terms of pure syngas, which shows a higher load for the same air-fuel ratio than pure methane.
For methane blends, a flat trend was seen in terms of performance (variations below
3%). This is in agreement with the review of Yan et al. [106], they have concluded that HCNG
engines generally have comparable power output level with CNG engines. Several effects of
hydrogen addition explained this behavior. Some of these factors are that hydrogen addition
could increase the rate of heat release inside the cylinder and therefore is beneficial to the engine
thermal efficiency and power output. In addition, H2 content may lead to increased heat loss
due to high cylinder temperature, contributing negatively to power output. Finally, hydrogen
114

substitution reduce the energy of the air fuel mixture due to hydrogen's low volumetric energy
density, which in turn tend to reduce the power output.
A comparison between methane and hydrogen shows that although the volumetric LHV
of hydrogen is only around the third part that methane, hydrogen requires 50% less of air, on
volume basis, to complete combustion. Therefore, the addition of hydrogen in methane will
certainly reduce the LHV of the fuel per unit volume; its effect on the volumetric heating values
of the air-fuel mixture exhibit different trends. It is noticed that for the stoichiometric mixture
(λ = 1.0), the LHV decreases as H2 content increases. However, for relatively lean mixtures (1.4
< λ < 1.6), the LHV becomes insensitive to H2 addition. For λ > 1.6 the trend is get reverted
and the LHV starts to increase as more hydrogen is added.
These trends were studied by Ma et al. [107] and are relevant to understand the behavior
of the combustion process in CH4-H2 blends. Therefore, many benefits of engines come from
their operations with extreme lean fuel/air mixtures, as opposed to the stoichiometric mixture
strategy for conventional gasoline engines. As a result, the effects of hydrogen addition on CNG
engine power performance are low in most cases and may change depending on different engine
operating conditions. Seems in section 2.4 was seen the opposite trend, this effect could be
associated to the retard in the spark timing.

a) b)
Figure 63– MFB curve for 200 cycles of methane, M25, M50, S50 and S75 at spark timing
for MBT a) λ=1.4 and b) extreme lean condition.

As a complement of the pressure traces, the MFB curves are depicted in Figure 63. The
fast flame propagation nature of the H2 species results in a short delay period, enhance their
115

lean burn capabilities and leading to the retard of the ignition onset. The effect was greater for
syngas due to the fuel properties of CO versus CH4 (Table 6). As can be observed in Figure 63
for methane, the addition of 50% H2 has a strong influence reducing in 3 CAD the first phase
of the combustion process compare to pure methane. For 25% of H2, the variation was lower,
1 CAD with respect to the baseline. For the extreme lean condition, the difference in terms of
burn duration was lower than λ=1.4, in which mixtures with 50% H2 shows rapidly increase in
the phase 0-10% MFB.
Fuel conversion efficiency was also calculated, as a direct relationship of power output
and energy content in the air-fuel mixture. Figure 64-a shows a flat trend with difference below
5% between all cases, half-content of H2 in volume basis shows the higher efficiency for both
fuels (methane and syngas). For extreme lean condition, a steadily decrease was seen with the
addition of H2. In addition, this trend is complemented with the decrease in terms of air-fuel
ratio from left to right. Methane and syngas for the same proportion of H2 shows almost the
same fuel conversion efficiency (below 5%). It is important to note that low efficiency is due
to effects of low effective compression ratio, captured unburned gas in the top land region and
higher heat transfer intrinsic to optical research engines with respect to commercial ones.

a) b)
Figure 64 – Fuel conversion efficiency for M, M25, M50, S50 and S75 at spark timing MBT
for a) λ=1.4 and b) extreme lean condition.

The last thermodynamic parameter analyzed was pollutant emission (CO, CO2, CH4,
HC, and NOx) measured in the exhaust gas stream, using a Multigas 2030 spectrometer analyzer
116

in the last 200 cycles of operation. It is well known the advantage of gaseous fuels compared to
traditional hydrocarbon fuels (gasoline or diesel) to reduce the level of pollutant emissions
[106]. In addition, hydrogen is characterized as a clean and renewable energy carrier. In general,
hydrogen addition could reduce CO2, CO and HC emissions due to reduction of carbon content
in the fuel mixture and enhanced combustion process [106]. However, NOx emissions in most
cases increase with the increase of hydrogen content, mainly through the increase in combustion
temperature [43]. In spite of this general trend being stated, the influence of hydrogen
enrichment on SI engine emissions could still vary depending on engine design and operating
parameters such as compression ratio, ignition timing, excess air ratio, engine load among
others [106]. To accomplish this, it is necessary to investigate the emission characteristics
methane-H2 and syngas engines and its relationship with hydrogen fraction and others
combustion parameters.
Figure 65 shows the CO emission measured in ppm against H2 content in the fuel (CH4
and CO), for lean and extreme lean condition. Could be seen that from M to M50 a steadily
increase of CO content was observed for both air-fuel ratios. Several experimental
investigations [108, 109] have confirmed that hydrogen enrichment helps to reduce CO
emissions. It is noticeable that generally these experiments are performed near stoichiometric
air-fuel ratio. Ma et al. [107] shows that for lean condition (over λ=1.3) this trend could be
reverted and high CO content is emitted. This is caused by the reduced combustion temperature
and thus decreased oxidation rates (transformation of CO in CO2). For this work could be seen
that the explanation of possible decrease in temperature profile correlates with the decrease in
peak pressure in the combustion chamber (Figure 61); therefore is a possible justification of
increasing CO emission.
This behavior for the extreme lean condition is enhanced (Figure 65-b) and an increase
in 36% for M25 and 62% for M50 was observed with respect to the baseline case. For syngas
blends, the trend is totally opposite, CO emission decrease with the H2 content for both air-fuel
ratios. In addition, the content is drastically increased for CO/H2 than methane blends (50% for
λ=1.4 and 1000% for extreme lean condition). This is directly related to the content of CO in
the fuel and the combustion efficiency. The addition of H2 to the fuel reduces the carbon
monoxide presence and as can be seen in Figure 65-a the emission is reduced. Moreover, in
extreme lean condition the combustion efficiency is reduced as was seen in Figure 64.
Therefore, more unburned gases are trapped in the top land region and crevices, increasing CO
emissions.
117

a) b)

Figure 65 - CO emission in [ppm] for the five blends study at (a) lean and (b) extreme lean
condition for MBT spark advance timing.

The second parameter analyzed was methane content in the exhaust gas. For CH4 and
syngas blends, the methane emission is also called methanic HC due to the direct proportion
between CH4 and HC. A large emission of methanic HC was measured for all the methane
blends and it decrease with the hydrogen content in the blend. This result can be explained due
to a higher homogeneity reached in the mixture and more stability of combustion, allowing a
complete combustion. It is also important to take in to account the decrease in C/H ratio. The
replacement of some natural gas by hydrogen decreases, in fact, the carbon fraction in the fuel
blends. For extreme lean conditions, the emission of CH4 increase due to instabilities of the
combustion process.
Syngas blends have almost negligible content of CH4 in the exhaust gases. This result
is in agreement with other works [33], in which was attributed to the absence of heavy
hydrocarbons in the syngas. Additionally, syngas is an oxygenated fuel because of its CO
species. This might contribute to facilitating combustion.
118

a) b)
Figure 66 - CH4 emission in [ppm] for the five blends study at (a) lean and (b) extreme lean
condition for MBT spark advance timing.

The most harmful product of hydrogen combustion is NOx. Lean burn technology is
well recognized for inhibiting in-cylinder NOx formation by combusting fuel at lower
temperatures, as was seen in chapter 1. Indeed, the EURO-V emission standard can be met
satisfactorily by lean-burn natural gas engines without the use of exhaust gas after-treatment
devices [110]. To meet the stricter NOx emission limit set by the EURO-VI standard, even
leaner NG-air mixtures need to be used for a further reduction of flame temperature, if the
expensive De-NOx devices are to be avoided [110]. However, due to NG's relatively slow flame
speed, it is typically impossible for traditional SI NG engines to run at such lean conditions
without significantly compromising engine efficiencies. The enrichment of NG with a fast-
burning fuel, i.e. hydrogen, which has a laminar burning velocity seven times higher than NG,
has been shown to be an effective method to extend the lean operation limit. Although for
syngas the studies are more limited, is well recognize that is an imperative necessity to ran at
extreme lean conditions to reach the emissions standards [1, 14, 49].
Figure 67 shows the NOx content in the exhaust gases for the five blends at two lean
operative conditions. For λ=1.4, methane blends shows almost negligible values. However, S50
and S75 was around 2000 ppm. These values shows that natural gas enrichment with hydrogen
at lean operative condition could be a solution in terms of low emissions, comparable
performance values and acceptable combustion stability (Table 15). In the case of Figure 67-b,
the extreme lean condition shows radically decrease for syngas blends and similar negligible
119

values for methane blends. Therefore, syngas with a high content of H2 (50% and 75%) need to
operate at lean condition above λ=1.4 to have acceptable NOx emission.

a) b)

Figure 67- NOx emission in [ppm] for the five blends study at (a) lean and (b) extreme lean
condition for MBT spark advance timing.

4.2.3 Optical Investigation

Optical techniques were applied in order to perform a local analysis of the combustion
process with high spatial and temporal resolution. In particular, high-speed 2D-digital flame
chemiluminescence imaging was performed. Figure 68 displays the combustion propagation
under different hydrogen content for CH4 and CO at λ=1.4. The image sequence shows the raw
chemiluminescence signal. It could be appreciated that for the same flame size, the CAD
position decreased with the increase of hydrogen for both fuels. The addition of hydrogen
stimulates the formation of H, O and OH radicals, as was shown by Di Lorio et al. [43]. These
radicals are effective in accelerating the chain reactions and consequently improving the local
laminar flame speed. Moreover, the flame of the carbon monoxide/hydrogen blends propagates
faster than methane blends highlighting that the CO accelerates the combustion process for the
same H2 content.
120

Figure 68 – Evolution of the flame front area and flame propagation speed for λ=1.4 obtained
by averaged data over 25 consecutive engine cycles in MBT condition.

Figure 69 shows the evolution of the flame front area and the flame propagation speed
for λ=1.4 in the maximum brake torque condition for each fuel. Methane-H2 blends shows
propagation improvements with the addition of H2. However, the differences between the three
blends are not as much as tests employing the same spark timing (SA -13, see Figure 45). This
behavior is due to the effect of retarding the SA, which was explained in the previous section.
Pure syngas shows similar effects than methane blends when changing spark timing. To put in
context, Table 16 shows a comparison between fixed SA and MBT condition for the same λ
value (λ=1.4). The difference in optimized SA between M50 and M25 is 1 m/s, while in the
same SA was 3.8 m/s. In the same way, pure syngas shows differences between S75 and S50
of 1.5 m/s for optimized cases and 5.1 m/s in fixed SA. This enhances the importance of the
effects of spark timing in the combustion behavior. Moreover, for the study of fuel properties
is better to use fix SA. On the other hand, for study real combustion operation condition is
preferred optimized setup.
121

(a) (b)
Figure 69 – Evolution of the flame front area and flame propagation speed for λ=1.4 obtained
by averaged data over 25 consecutive engine cycles in MBT condition.

Table 16 – Maximum flame speed comparison between fixed spark advanced and MBT
condition for M, M25, M50, S50, and S75 at λ=1.4.

Fuel SA Flame Propagation Speed


[CAD BTDC] [m/s]
M -13 8.7
[%]
-11 8.4
M25
-13 9.8
-7 9.4
M50
-13 13.6
0 18.4
S50
-13 23.2
+8 19.9
S75
-13 28.3

The extreme lean condition was also analyzed in the MBT point and the data is plotted
in Figure 70. The S50 blend was the fastest fuel in the first phase of the combustion process (up
to 15 CAD ASOS). However, the baseline case (methane) reach the peak flame speed with 8.7
m/s at 16 CAD ASOS. The flame propagation speed decreases rapidly after the peak value for
both fuels (Figure 70-b). For the other three fuels, the maximum flame propagation speed
presents similar values (7 m/s) and suffer a shift to the right (20 CAD ASOS) in relation to S50
and methane.
122

(a) (b)
Figure 70 – Evolution of the flame front area and flame propagation speed for extreme lean
condition obtained by averaged data over 25 consecutive engine cycles in MBT condition.

The flame distortion and center displacement for the two operative condition and the
five blend are plotted in Figure 71 and Figure 72. For the same λ value, the addition of hydrogen
reduces the flame circularity (near to a circumference). The effect was greater for CO than CH4.
This trend was also seen when fixed SA was set (Figure 47). Table 17 shows a comparison
between the optimized case and fixed spark timing in terms of maximum HCF. In general, the
retard of the spark reduces flame distortion.
In terms of flame displacement, as was seen in previous chapters the fuels have
preferential propagation for the intakes valves. The hydrogen content reduces this preferential
propagation and as an example for pure syngas, the movement was almost directly from the
spark plug to the center of the chamber. Table 17 shows that for methane blends the retard of
SA increases the maximum distance from the center of the combustion chamber. This behavior
could be explained due to reducing of flame propagation speed. For pure syngas, the trend was
reverted, and the advanced cases (MBT point) shows lower flame displacement than fixed SA.
In spite of flame propagation speed decrease with the retard of SA, these cases (MBT point of
pure syngas) propagates in the expansion stroke in which less fluid motion effect is present.
123

(a) (b)
Figure 71 - Heywood circularity of the flame for (a) λ=1.4 and (b) extreme lean condition
obtained by averaged data over 25 consecutive engine cycles in MBT condition.

For the extreme lean condition, the trends were not so clear due to different air-fuel
ratios and spark timings settings. However, the fastest fuels (M and S50) presents higher flame
distortion and then decrease rapidly in the initial phase. For M25, M50 the distortion was lower
than pure methane, showing the importance of H2 content to reduce flame deformation. The
leanest case (S75) presents a higher flame distortion due to low flame speed and advanced spark
timing. This enhances the effect of low flame propagation speed in the shape of the flame.
Center displacement for the extreme lean condition is shown in Figure 72-b. The fuels
presented maximum distances in y-axes from the combustion chamber center of: M 6.8 mm,
M25 7.8 mm, M50 5.7 mm, S50 5.0 mm, and S75 6.9 mm. In spite of air dilution, high H 2
content in methane reduce flame displacement. Pure syngas in extreme lean condition shows
similar values than methane at extreme lean condition.
124

(a) (b)
Figure 72 – Flame centroid position for (a) λ=1.4 and (b) extreme lean condition, obtained by
averaged data over 25 consecutive engine cycles in MBT condition.

Table 17 - Maximum HCF and distances from the center of the combustion chamber
comparison between fixed spark advanced and MBT condition for M, M25, M50, S50, and
S75 at λ=1.4.

Fuel SA HCF Maximum flame displacement


[CAD BTDC] [mm]
M -13 [%]
1.44 7.1
-11 1.40 6.6
M25
-13 1.34 5.9
-7 1.32 5.2
M50
-13 1.33 4.8
0 1.17 1.7
S50
-13 1.21 1.8
+8 1.10 0.8
S75
-13 1.16 1.3
125

4.3 Summary

The present chapter resumes the results obtained by an experimental activity performed
on a port fuel injection (PFI) spark ignition (SI) engine fueled with methane-H2 and equivalent
syngas blends. The objective of the proposed research was to improve the knowledge of the
combustion process when alternative bio-derived fuels are used to replace fossil energy sources
in different spark timing configurations. The experiments were carried out at low speed and
wide open throttle, the typical working point of SI engines for stationary applications. The effect
of different spark timing for the same fuel (S50) and air-fuel ratio (λ=1.0) was studied in
section 4.1. Finally, in section 4.2 was included a comparison of five different blends in
maximum brake torque condition (optimum spark timing) and lean conditions (λ=1.4 to
λ=1.75).
Pure syngas with 50% of hydrogen content was studied in detail over several spark
timing configuration in the compression and expansion stroke. The optimum set up for
stoichiometric air-fuel ratio was seen in 8 CAD ATDC. The behavior is different from
conventional fossil fuels, in which generally engine mapping works in values before the top
dead center. This is due to the short burning duration (0-50% MFB at 17 CAD) with respect to
methane (0-50% MFB at 32 CAD). The retard of the SA was also seen for S50 by other
researchers in several engine conditions [33, 52] with respect to methane.
The peak in-cylinder pressure and NOx emissions decrease with the retard of the spark
timing. The pressure decrease is due to the piston position and the effect in NOx emission could
be associated with a decrease in the combustion temperature. Optical investigations
demonstrated that flame propagation speed decrease with the retard of SA, with the extreme
cases speed of 25.7 m/s for SA -7 CAD ATDC and 21.8 m/s to +8 CAD ATDC. The flame
distortion also presents a decrease with the retard of the SA, with values around 1.09. Finally,
the flame center displacement was similar for the six spark timing condition with a maximum
distance from the center of the chamber at 0.5 mm. This is due to the fast propagation speed of
S50 at λ=1.0, with an average of 22 m/s.
An experimental study was undertaken in order to evaluate the effects of hydrogen
addition to methane, and pure syngas (CO/H2) blends, on combustion in an SI power unit at
optimum spark advance set up. To this aim, methane was used as baseline fuel and blended
(25% and 50% of hydrogen addition). Finally, two syngas equivalent mixtures, with 50% and
75% hydrogen content was tested. The study was focused on lean and extreme lean conditions,
126

with a thermodynamics and optical analysis of the combustion process. The overall
thermodynamic analysis revealed that hydrogen addition improved the flammability range of
CH4, with a maximum λ of 1.5 for 50% H2. For pure syngas, this range was still larger with λ
of 1.7 for 50% H2 and λ of 1.75 for 75% H2.
Thermodynamic analysis reveal that pure syngas featured higher peak pressure and
similar pressure curves in the last phase of the combustion process than methane blends.
Therefore, higher IMEP values were obtained (9% over) than methane blend for the same
hydrogen content. The five fuels that were tested ensured acceptable engine performance and
stability for λ=1.4. One major conclusion of the analysis was that spark timing re-calibration
allows to fully taking advantage of fuel properties such as lower burning duration and increased
stability.
Cycle-resolved flame imaging allowed a more detailed insight into the fuel oxidation
processes with regard to its evolution within the combustion chamber. In line with the
thermodynamic results, S75 presents the fastest flame propagation due to the high H2 content.
However, the flame speed decrease with respect to the advanced case (study in section 3.2) for
all fuels. Propagation was found to be quite symmetric in all directions for syngas (lower than
1.2), and similar for all methane blends (around 1.4), suggesting that flame speed had the most
important effect. In terms of flame centroid displacement, the higher effect of spark timing was
seen for the extreme lean condition, in which the advanced of the SA increase the distance with
respect to the combustion chamber.
Hydrogen in engines has recently received some interests [111-113], however, its
widespread use is still likely to be hindered by many practical difficulties which include, but
are not limited to large-scale hydrogen production, storage, fueling infrastructures as well as
engine abnormal combustion. At the current stage, being an additive for fuels seems to be a
more reasonable approach to promote the application of hydrogen as transport fuels. Even if
further investigations are required to obtain a deeper understanding of the effects of spark
timing on the combustion process and SI engine performance in a wide range of operative
conditions, the current work represents a contribution to the detailed database to validate
numerical models and optimization codes.
127

5 Effect of Diluent Content in Syngas Combustion Process

Syngas is ideally a mixture of hydrogen and carbon monoxide [11]. However, in its real
composition, large amounts of nitrogen (N2) and carbon dioxide (CO2) can be present. These
components depend on several factors as the oxidizer used or the type of gasification process
applied to obtain the syngas [2]. As already mentioned, the syngas is characterized by high
variability in composition depending on the process conditions and feedstock composition [33].
In this way, its use can imply some practical difficulties, being of fundamental importance a
more detailed study so that the syngas can be used in the next generation of SI engines.
Downdraft fixed bed gasifier (DFBG) can handle all types of dry small sized biomass
wastes. Agroindustry and municipal waste can be used for power generation with this type of
gasifier technology. These locations are generally rural areas in which low cost of installation
and simple operation are necessary. Therefore, DFBG coupled with SI engines is an interesting
option and is being studied by several authors to apply in real applications [12, 13, 17, 45].
Generally, the content of inert gases in the fuel at the outlet of DFBG is around 50% when the
air is used as an oxidizer (Table 2). Since the energy content of syngas is lower than the
conventional hydrocarbon fuels on a volume basis, the volumetric efficiency, and as a
consequence engine output could decrease when syngas is used as fuel in SI engines. In
addition, the AFRst is between 5.0-1.0 compared to 17.2 for methane and thus long injection
timing are needed. Therefore, the use of syngas in spark ignition (SI) engines need to sort out
certain related combustion issues before being used in real applications.
Furthermore, several authors [12, 13, 48, 50, 53] study a large proportion of diluents in
SI engines and compared with natural gas (CH4) performance. However, up to the knowledge
of the author, not optical techniques were applied to follow the combustion process with these
type of mixtures. For that reason, this chapter adds important information to continue the
fundamental development of syngas engines.
Starting from these considerations, in this study two different equivalent syngas
mixtures S50 and S75 (pure syngas) with the addition of 50% diluents (named as S5050 and
S7550) were tested in a PFI-SI engine equipped with a piston optical access. Experiments were
performed at fixed crankshaft rotational speed and maximum load, representative of stationary
applications. The λ was changed from stoichiometric (λ=1.0) to values close to the extreme lean
condition of methane (λ=1.4). The analysis of the combustion process was done through
combined thermodynamic and optical investigations as was explained in chapter 3. In
128

particular, high spatial resolution cycle resolved digital imaging was used to characterize the
flame front propagation.

5.1 Operating Points

Two equivalent syngas mixtures were tested (Table 18) in order to analyze the effect of
the degree of dilution proportions (DOD) on the combustion process. The CO/H2 ratio selected
(S50 and S75) are representative of the family of pure syngas and was used in the last chapters.
Table 19 shows the main properties of the blends used in this study. As could be seen, both
diluted syngas has an LHV 9 times lower than pure methane and 3 times lower than pure syngas
with the same H2 content in fuel basis. Therefore, this chapter covers a wide range of low
heating values (from 1.5 MJ/kg to 17 MJ/kg) operating in real SI engine geometries.
Moreover, laminar flame speed was calculated by CHEMKIN-PRO [86] for the five blends to
add the fundamental basis for the analysis. Both diluent mixture speed was between methane
and pure syngas. For the same proportion of H2 in the fuel basis, a reduction of 60% was seen
in the laminar flame speed when the diluent content was added. In spite of a large amount of
inert gases, the laminar flame speed for S7550 and S5050 was 2.0 times and 1.5 times faster
than methane, respectively. This behavior is due to the combined effect of H2 and CO as
observed in the previous chapters.

Table 18 – Fuel composition in the percentage of total volume used for study the effect of
diluent content in pure syngas combustion process.

Fuel CH4 H2 CO CO2 N2


[%] [%] [%] [%] [%]
M 100 0 0 0 0
S50 0 50 50 0 0
S5050 0 25 25 15 35
S75 0 75 25 0 0
S7550 0 37.5 12.5 15 35
129

Table 19 – Main properties of the fuels used for study the effect of diluent content in pure
syngas combustion process.

Properties M S50 S5050 S75 S7550


LHV [MJ/kg] 50.2 17.5 5.5 29.7 6.1
AFRst [kgair/kgfuel] 17.2 4.59 1.44 8.11 1.73
Peak flame temp [K@1 atm] 2223 2371 2005 2373 1981
Laminar Flame Speed
0.35 1.23 0.52 1.77 0.73
[m/s@1Atm, 293K, λ=1.0]
H2/CO - 1.0 1.0 3.0 3.0
H2mass [%] 0 6.7 2.1 17.8 3.7
DODvol [%] 0 0 50 0 50

The λ was initially set at stoichiometric and then increased until reaching the
flammability limit of methane (λ=1.4). Spark advance (SA) was fixed in order to analyze the
combustion process with a focus on the fuel composition effect. As was seen in chapter 4, the
different set of SA affects the flame behavior in terms of flame propagation speed and other
morphology parameters. Therefore, the same SA allowed the analysis of the flame behavior in
the combustion chamber at roughly the same fluid-dynamics conditions. In addition, this
strategy is commonly used when is necessary to apply the ‘feed and run’ concept. This means
not change the engine control parameters when fuel composition varies. The results of diluted
mixtures were compared with the results of section 3.2.
A resume of the test condition can be seen in Table 20. It can be seen that the increase
in duration of injection (DOI) is proportional to the AFRst presented in Table 19, with the
difference between extreme cases of 360 CAD (methane and S5050) for λ=1.0. This shows the
importance of study the behavior of the combustion process with conventional injection and air
inlet technology to be able to propose a new system for the future.
130

Table 20 – Operating conditions used for study the effect of diluent content in pure syngas
combustion process. Engine speed: 900 RPM; SA: 7 CAD BTDC; Load: WOT.

Fuel λ DOI [CAD]


1.0 60
M 1.2 50
1.4 43
4
1.0 204
S50 1.2 180
1.4 155
1.0 420
S5050 1.2 360
1.4 290
1.0 150
S75 1.2 130
1.4 105
1.0 390
S7550 1.2 340
1.4 290
4

5.2 Results

5.2.1 Thermodynamic results

A preliminary global analysis of the combustion process was performed through the
thermodynamic approach. In-cylinder pressure was analyzed as the average value of 200
consecutive acquisitions; results are reported in Figure 73. The motored pressure signal (dashed
line) is also shown as a reference. For both λ conditions, pure syngas shows the highest peak
pressure, followed by diluted syngas blends and lastly pure methane. As all the fuels were
ignited in the same spark timing (same piston position), it could be seen large differences in
terms of burning duration between fuels. This also causes a shift to the right of the maximum
pressure value.
In stoichiometric air-fuel ratio, methane shows the highest value of pressure in the last
phase of the combustion process (Figure 73-a), because of optimized SA in MBT point. Due to
its large content of inert species, S5050 and S7550 resulted in a decrease of 22% of the
maximum in-cylinder pressure. Additionally, there was observed 8 CAD shift to the right in
terms of peak position with respect to pure syngas (compared the same H2 proportion in the
131

fuel basis). Another interesting point is that the same dilution (N2 + CO2) causes a similar
reduction in both pure syngas in terms of pressure.
The lean condition shows an important decrease in terms of maximum in-cylinder
pressure for methane and both diluted syngas (Figure 73-b); even though pure syngas
maintained the high-pressure development in the initial phase. In addition, syngas with 50% of
dilution content presents higher-pressure rates than methane in lean conditions.

a) b)

Figure 73 – In-cylinder pressure traces averaged over 200 consecutive cycles at λ=1.0 for: (a)
50% H2 blends and (b) 75% H2 blends, compare with methane.
132

a) b)

Figure 74 – In-cylinder pressure traces averaged over 200 consecutive cycles at λ=1.4 for: (a)
50% H2 blends and (b) 75% H2 blends, compare with methane.

Performance of the fuel mixture was analyzed by IMEP (Figure 75), an independent
parameter of the cylinder volume. Due to spark timing was optimized for methane in λ=1.0,
higher IMEP value was seen for this fuel than syngas. For λ=1.0, S50 and S5050 presents
similar performance as well as S7550 with respect to S75 (slightly increase, around 5%). This
is directly related to the “fed and run concept”, in which the performance will be strongly
affected when existing an important difference between optimized SA and the fixed SA chose.
For intermediate values (λ=1.2), similar trends were seen that for the stoichiometric
case. Pure and diluted syngas shows identical values of IMEP for the same H2 content, as was
seen in λ=1.0. This suggests that the decrease in volumetric efficiency due to fuel dilution is
less important than the optimization of SA in this engine condition (900 RPM, WOT and λ=1.0-
1.2)
Finally, the leanest case (λ=1.4) shows an important decrease for diluted syngas and
methane; meanwhile, S50 and S75 conserve the performance. This is in line with the behavior
of the peak in-cylinder pressure previously described. In this operative condition, the effect of
high air dilution and fuel dilution were added to S50 and S75. Therefore, syngas (S7550 and
S5050) performance presents important reduction compare to pure one.
133

6.0
M S50 S5050 S75 S7550
5.0

4.0
IMEP [bar]
3.0

2.0

1.0

0.0
λ=1.0 λ=1.2 λ=1.4
Figure 75 – IMEP evaluated average on 200 consecutive engine cycles for different air-fuel
ratios (λ=1.0-1.4) for M, S50, S5050 and S7550 in fixed SA (7 CAD BTDC).

The cycle-by-cycle variation is an important point when developing alternative fuels,


due to the stability of the machine in long-term operation. The most common parameter to
measure this phenomenon is CovIMEP (Figure 76), a limit for good drivability is commonly
set at 3% [55]. For stoichiometric air-fuel ratio, pure syngas shows high values near the limit
imposed. Diluted syngas and pure methane were below 1%, showing that instabilities are not a
limitation for use syngas in SI engines in spite of high dilution proportion.
For intermediate lean case, the trend was similar with a decrease in CovIMEP for pure
syngas with respect to the stoichiometric case. In addition, S5050 shows the lowest value. This
could be explained due to a reduction between optimized and fixed SA of all syngas blends.
The same behavior was seen for S7550 in λ=1.4. Methane shows CovIMEP value over the
maximum acceptable (3%) and syngas resulted in low CovIMEP values (around 1.3%). This is
most likely related to the presence of hydrogen in its composition, a component that is known
to ensure better combustion stability in high air dilution conditions [114]. These analysis
enhance the importance of hydrogen content and the limitations of the “fed and run” concept.
134

4.0
M S50 S5050 S75 S7550

Cov IMEP [%]


3.0

2.0

1.0

0.0
λ=1.0 λ=1.2 λ=1.4

Figure 76 – CovIMEP evaluated on 200 consecutive engine cycles for different air-fuel ratios.

The observed trends are also reflected in the calculated crank angle durations at given
mass fraction burned levels (Table 21). These results emphasize the much stronger effect of λ
on methane and diluted syngas combustion compared to pure syngas, especially in the kernel
development phase (0-5% of MFB). They also give an idea into re-calibration procedures that
can be considered for adapting ignition settings for the alternative fuels [11] and feedback
strategies for real-time engine control. The higher proportion of H2 in the fuel S75 over S50
(and in S7550 over S5050) results in lower burning times for all λ values tested. Furthermore,
syngas blends presents a higher rate of mass fraction burned when compared with the baseline
fuel in the same λ value.
135

Table 21 – Crank angle durations from the start of spark to fixed mass fraction burned
thresholds calculated from pressure traces for study the effect of diluent content in pure
syngas combustion process. Engine speed: 900 RPM; SA: 7 CAD BTDC; Load: WOT.

Fuels λ 5% 10% 50%


[CAD ASOS] [CAD ASOS] [CAD ASOS]
1.0 13 16 32
M 1.2 13 17 33
1.4 16 21 43
1.0 5 6 21
S50 1.2 6 8 25
1.4 7 9 27
1.0 9 11 28
S5050 1.2 10 13 28
1.4 13 17 37
1.0 3 4 17
S75 1.2 4 5 17
1.4 5 6 18
1.0 7 9 25
S7550 1.2 8 11 24
1.4 12 16 36

The power reduction due to non-optimization of spark timing, when using different
syngas, is most evident when looking at fuel conversion efficiency data in the stoichiometric
and intermediate high λ values (Figure 77). Even if it is not directly comparable to the number
usually found in the literature since optical engines have higher blow-by losses and lower
compression ratio [115], the relative tendency is used to verify the behavior of the combustion
related to the composition of the fuels. In general, efficiency levels were found to be around
20%. Nonetheless, as an initial evaluation, the ‘feed-in and start’ scenario ensures acceptable
engine operation in a relatively wide range of air-fuel ratios.
In the leanest case, diluted syngas presents a loss of efficiency. This may be associated
with reduced volumetric efficiency and relatively high values of CovIMEP, thus confirming
that increasing inert gas levels result in poor combustion development. This suggests that turbo-
charged applications (that are the majority of cases in stationary applications over a certain level
[116]) would be more suited for syngas fueling, given that the drop in volumetric efficiency
can be compensated for with increased boosting pressure. On the other hand, higher overall in-
cylinder pressure negatively affects laminar flame speed, thus reducing combustion speed.
136

25
M S50 S5050 S75 S7550

Fuel Conversion
Efficiency [%]
20

15

10
λ=1.0 λ=1.2 λ=1.4
Figure 77 – Fuel conversion efficiency for all the selected conditions and fuels.

The last parameter analyzed in this section was the exhaust emissions measured with
Multigas 2030 spectrometer analyzer in the last 200 cycles of operation.
Table 22 shows pollutant emission (CO, CO2, CH4, and NOx) for λ=1.4. The emission
of CO increases with the original content in the fuel and with the combustion efficiency, due to
unburned gas trapped in the crevice volume. For pure syngas, S50 presented the highest value
because of more presence of CO in the fuel. The same trend was seen with diluted syngas
(S5050 with respect to S7550). In spite of S50 have more CO content than S5050 in volume
basis, the last one shows an increase of 5 times the CO emission. This is explained comparing
the fuel conversion efficiency, in which for the leanest case a loss in this parameter was seen.
Therefore, more unburned gases were in the exhaust port compared to pure syngas. The same
occurrences were observed with S7550 with respect to S75.
NOx emission shows an important decrease for diluted syngas and methane with respect
to pure syngas. As was seen in previous chapters an important dependence on temperature
(associated to maximum in-cylinder pressure) exist in NO2 and NO generation in the
combustion process. Therefore, the same trend as was seen in Figure 73-b could be appreciated
in NOx emission values. The other gases measured shows expected trend, with CH4 almost
negligible for syngas and higher concentration of pure methane (unburned gases) and CO2 for
fuels with 50% of H2 in fuel basis. This is due to more concentration of CO in the fuel than
syngas with 75% of H2.
137

Table 22 – Exhaust gas emissions measured at engine speed: 900 RPM; SA: 7 CAD BTDC;
Load: WOT, and λ=1.4.
CO CO2 NOx CH4
Fuels
[ppm] [%] [ppm] [ppm]
M 545 4.5 10 13824
S50 1555 11.2 3076 5
S5050 7419 9.7 2 5
S75 1179 6.2 3115 5
S7550 3603 6.5 2 5

5.2.2 Optical Investigation

Even if the in-cylinder pressure measurements allow a comprehensive analysis of


combustion characteristics, they do not furnish detailed results of the local distribution of the
burned mass, flame behavior inside the combustion chamber and in the flame speed. In this
sense, cycle resolved visualization represents a powerful tool for quantitative analysis of flame
front propagation. Table 23 presents the camera setup used in this chapter, for methane and
pure syngas mixtures at λ=1.0 increase of f-number need to be done due to saturation of the
images. A re-calibration of the scale and focus was performed because of difference between
these two f-number setups. More information about the image processing was added in
Appendix A.

Table 23 – Camera set up for study the effect of diluent content in pure syngas mixtures at
fixed spark timing (7 CAD BTDC). Exposure Time: 0.5 CAD; Intensification: 55000.
Fuel λ F-number
1.0 32
M 1.2 5.6
1.4 5.6
1.0 32
S50 1.2 5.6
1.4 5.6
1.0 5.6
S5050 1.2 5.6
1.4 5.6
1.0 32
S75 1.2 5.6
1.4 5.6
1.0 5.6
S7550 1.2 5.6
1.4 5.6
138

Figure 78 and Figure 79 shows a sequence of representative images detected during one
engine cycle in the stoichiometric air-fuel ratio for the five fuels. In all investigated cases, the
flame kernel was well resolved only around 1–2 CAD after ignition for pure syngas and 3-4 for
diluted syngas and methane. This phenomenon was also seen by other researchers and justified
due to the very high luminosity of spark-induced plasma [44] and problems with spark plug
electrode [58].
The high difference in terms of luminosity between fuels is due to changes in the f-
number. Therefore, no direct comparison in terms of luminosity could be done. However, the
fuels can be compared in terms of morphology (flame speed, distortion, and preferential
displacement). For both figures, diluted syngas and methane flames presents higher distortion
and increase the preferential displacement to intake valves (downside). Another interesting
point is the position that the same flame area percentage is reached. In the first combustion
phase, between S50 and S5050 there was a difference of 3 CAD, meanwhile in the last phase
of the optical measurements increased to 6 CAD. Similar behavior was found with S75 and
S7550.

Figure 78 – Images of flame evolution for stoichiometric air-fuel ratio with M, S50, and
S5050.
139

Figure 79 – Images of flame evolution for stoichiometric air-fuel ratio with M, S75, and
S7550.

The same camera set up between fuels was used for lean air-fuel ratio cases. This allows
both comparisons in terms of morphology and in terms of flame emission. Raw data was
depicted in Figure 80 and Figure 81. It is observed that pure syngas has the highest emission
intensity, followed by diluted syngas and methane. In addition, the flame thickness was greater
for S50 and S75 than for the diluted gases. Flame position for the same area percentage resulted
in pure syngas was faster with 5 CAD in the initial phase and 15 CAD in the last images
recorded than diluted mixtures for both H2 contents.
140

Figure 80 – Images of flame evolution for λ=1.4 with M, S50, and S5050.

Figure 81 – Images of flame evolution for λ=1.4 with M, S75, and S7550.

By applying the image processing described in chapter 3, the trends of mean diameter
and propagation speed of the flame were obtained. Results reported in Figure 82, Figure 83,
Figure 84 are related to the averaged values over 25 consecutive engine cycles. It is quite
evident that pure syngas was always faster than diluted syngas and pure methane (also seen in
chapter 2), starting with the initial combustion phase. These results are in line with the
thermodynamic analysis based on in-cylinder pressure evaluation. It is noticeable that optical
investigation allows the study of the combustion process from 0% to 25% MFB approximately
(see Table 21), phase in which thermodynamic analysis are not accurate.
141

Moreover, these findings follow the same trend of the calculated laminar flame speed at
atmospheric condition (see Table 19). In general, when same fluid-dynamic conditions are set
in the ignition of the flame, the correlation between laminar and engine like condition
propagation could be observed [43]. However, Table 24 shows that engine like condition
reduces the effect of fuel in the propagation of the flame. Other chemical and fluid-dynamic
processes are involved in the combustion process; so reduce the difference observed in laminar
tests. As was expected stoichiometric air-fuel ratio shows the fastest propagation speed with a
reduction of 40% from pure to diluted syngas in both H2 proportions. In addition, S7550 and
S5050 shows an increase of 45% and 15% with respect to methane, respectively. As a general
consideration on the use of syngas mixtures, it is possible to observe that the presence of
hydrogen improved the propagation of the flame, compensating the slowdown effect of dilution
through inert gases. These results are also in line with those of the thermodynamic analysis.
The lean cases (λ=1.2 and λ=1.4) showed the same trend than stoichiometric air-fuel
ratio. For the leanest operation point, S7550 suffer a decrease of 8.7 m/s with respect to λ=1.0
and in the case of S5050, a decrease of 5.5 m/s (Figure 84) was seen. The direct comparison
between fuels and air-fuel ratio could be done because spark timing was fixed; as was seen in
chapter 3, these trends would change if optimize condition for each fuel are set. This point was
not part of the objectives of this chapter.

a) b)
Figure 82 - Evolution of the flame front area and propagation speed for selected
stoichiometric air-fuel ratio obtained by averaged data over 25 consecutive engine cycles.
142

a) b)

Figure 83 - Evolution of the flame front area and propagation speed for λ=1.2 obtained by
averaged data over 25 consecutive engine cycles.

a) b)

Figure 84 - Evolution of the flame front area and propagation speed for λ=1.4 obtained by
averaged data over 25 consecutive engine cycles.
143

40
M S50 S5050 S75 S7550

Maximum Flame
30

Speed [m/s]
20

10

0
λ=1.0 λ=1.2 λ=1.4
Figure 85 – Maximum flame propagation speed obtained by averaged data over 25
consecutive engine cycles.

Table 24 - Comparison between Laminar Flame Speed (LFS) and Engine like Condition
Flame Speed (EFS) for the five fuels study in stoichiometric condition. Engine speed: 900
RPM; SA: 7 CAD BTDC; Load: WOT.
LFS EFS
LFSf/ LFSm EFSf/ EFSm
Fuel [m/s@1Atm, [m/s@Engine
293K, λ=1.0] condition, λ=1.0]
M 0.35 12.6 1.0 1.0
S50 1.23 25.7 3.5 2.0
S5050 0.52 14.3 1.5 1.1
S75 1.77 31.6 5.1 2.5
S7550 0.73 18.1 2.1 1.4

The effect of the fuels dilution and air dilution on the flame front shape were evaluated
in terms of distortion by the Heywood circularity factor. Results reported in Figure 86 clearly
demonstrated that during stoichiometric combustion the flame propagated quite uniformly in
all directions, even if the flame front of methane resulted less “circular”. Figure 87 shows a
comparison between the five fuels in the three air dilution ratios tested. In lean burn conditions,
the simultaneous action of the flow field and fuel charge distribution determined a strong
increase in the flame distortion, even if the effect was feeble for syngas mixtures. Due to the
high presence of H2 in the syngas blends the effect was weak and similar HCF values were
measured for λ=1.4.
144

Figure 86 – Heywood Circularity Factor for stoichiometric air-fuel ratio obtained by average
data over 25 consecutive engine cycles.

1.6
M S50 S5050 S75 S7550
1.5
Maximum HCF

1.4

1.3

1.2

1.1

1.0
λ=1.0 λ=1.2 λ=1.4
Figure 87 – Maximum Heywood Circularity Factor obtained by average data over 25
consecutive engine cycles for all air-fuel ratios.

The last flame morphology parameter analyzed in this chapter was the average center
movement. Figure 88 shows the trends for stoichiometric air-fuel ratio, as was seen in previous
chapters an initial movement to the intake valves and then a return to the center of the
combustion chamber was measured. Almost negligible movements in the horizontal axes were
seen; therefore, in Figure 89 maximum vertical displacement was depicted. The addition of H2
reduces the movement to intake valves (S75 with respect to S50 and S7550 with respect to
S5050). However, the increase in fuel dilution enhances the preferential displacement for intake
valves (S7550 and S5050). Air dilution has a similar effect than fuel dilution. This can be seen
145

when pure syngas is diluted in the air (S50 and S75 at λ=1.4) and when inert gas was added to
pure syngas (S5050 and S7550 at λ=1.0). When both effects were tested (S5050 and S7550 at
λ=1.4) similar values in terms of flame speed, distortion and center displacement than methane
with air dilution (methane at λ=1.4) was found.

Figure 88 – Evolution of the luminous centroid with respect to the geometrical center of the
combustion chamber; each step of the path is obtained by averaging data over 25 consecutive
engine cycles for the stoichiometric air-fuel ratio.

5.0
M S50 S5050 S75 S7550
displacement [mm]
Maximum Y

0.0

-5.0

-10.0
λ=1.0 λ=1.2 λ=1.4
Figure 89 – Maximum centroid displacement with respect to the geometrical center of the
combustion chamber in the y-direction; each step of the path is obtained by averaging data
over 25 consecutive engine cycles.
146

5.3 Summary

An experimental study was undertaken in order to evaluate the effects of syngas


composition on combustion in SI power units. To this aim, an optically accessible engine was
ran at 900 RPM and WOT specific for power generation applications, with pure and diluted
syngas in two H2/CO ratios.
The overall thermodynamic analysis revealed that the four syngas mixtures ensured
acceptable engine stability and similar engine performance for the three air dilution tested, even
if the spark timing setting was kept at the MBT point for methane. In the “feed and run” concept,
a reduction of 25% in IMEP was seen for syngas with respect to methane in stoichiometric
condition. However, this difference was reduced for pure synthetic gas (almost similar
performance) for lean mixtures. Diluted syngas mixtures due to non-optimization of spark
timing presents lower values of IMEP than pure methane.
A brief analysis of fuel conversion efficiency shows that for stoichiometric air-fuel ratio,
diluted mixtures presents better results than pure syngas due to the large difference between set
SA and MBT point. In addition, diluted syngas shows similar value in terms of efficiency than
methane. These results reinforce the importance of new studies related to syngas technology
for SI engines. The addition of air dilution shows a greater effect in diluted syngas with
efficiency below 18% due to non-optimized spark timing and a significant reduction of
volumetric efficiency as was mentioned in the analysis of IMEP. One major conclusion of the
analysis was that spark timing re-calibration is required in order to fully take advantage of fuel
properties such as higher laminar flame speed and increased stability, especially during lean
operation.
Cycle-resolved flame imaging allowed a more detailed insight into the fuel oxidation
processes with regard to its evolution within the combustion chamber. As expected, a clear
distinction between the flame front and reaction zone could be observed for the lean cases,
while stoichiometric fueling featured a more even distribution of the excited species specific
for exothermic reactions. In line with the thermodynamic results, the pure syngas mixtures
shows the faster flame propagation follow by the diluted syngas and lastly pure methane.
Propagation was found to be quite symmetric in all directions and similar for all fuels except to
methane, suggesting that H2 content and fluid motion had the most important effect in this
sense. These results were also confirmed by the evolution of flame centroid displacement, more
extensive for lean air-fuel ratios, but comparable nonetheless.
147

6 Quasi-dimensional Modelling in an Optically Accessible


SI Engine with Syngas

In order to obtain high levels of specific power output and efficiency, modern SI units
need to operate as close as possible to limit conditions. These set up depends on the type of fuel
used, injection and ignition system, load and engine speed among others. To reach the best
condition of operation, fundamental knowledge of the combustion process is necessary in order
to reduce time and cost. More difficulties arise considering the stochastic nature of engine
flows, causing the cycle-to-cycle variability [85]. In this scenario, numerical simulation is
becoming more and more important, since it can help to reduce expensive testing and
prototyping enabling a quick comparison among different solutions.
The great progress that has been achieved so far in the field of spark ignition engines
technology is to a significant extent attributed to the development and use of cycle simulation
models [117]. Nevertheless, to properly face the current engine development trends, models are
continuously evolving and becoming more and more complex. Therefore, tools able to describe
and predict combustion evolution are fundamental in the design of modern SI units with
alternative fuels [55].
Prediction models used to simulate the working cycle of SI engines generally fall into
three types: zero-dimensional, quasi-dimensional and multi-dimensional ones. The first one is
basically a pure thermodynamic model based on mass and energy conservation which totally
neglects the spatial distribution of flow parameters and treats combustion simply as a heat
addition process [118]. The rate of heat addition is usually obtained by empirical formulas as
Wiebe functions, which approximate the shape of experimentally observed burn rate curve by
calibration. The main disadvantage of this model is that since it considers the whole cylinder to
be uniform without spatial distribution, it cannot simulate the flame propagation process and
thus cannot predict burn rate. Moreover, the zero-dimensional model is critically constrained
by the use of a predefined heat release profile which renders it virtually nonresponsive to
operating condition changes. This kind of model is usually used to examine the engine overall
performance parametrically [34].
The second type commonly used and focus of the current chapter is quasi-dimensional
model. This is an advancement of the thermodynamic zero-dimensional model where in-
cylinder heat release is modeled as turbulent combustion process. Therefore, involves
148

parameters associated with the propagation of the flame like laminar flame speed and
turbulence intensity [85].
The governing equations for such models are based on conservation of mass and energy.
One assumption commonly used, is the two-zone formulation that separates the burned from
the unburned gases by an infinitely thin, spherically propagating flame front, called entrainment
zone. Quasi-dimensional modeling of combustion in SI engine has been approved as a
computationally low-cost dynamic model, suitable for engine modeling, control, and basic
design [85]. The main advantage, with correct empirical models and previous validation studies,
is that allows testing several operating conditions and fuels with a low computational cost [99].
Multiple-dimensional model, on the contrary, can provide detailed information about
the cylinder flow field and species concentration. Fundamentally, it coupled combustion
chemistry and species transport with the 3D CFD model [122], which by itself is already
overwhelmingly complicated. Although it is the most robust model, it is also the most
complicated one in terms of validation and computational cost [124]. In addition, some studies
like cycle-to-cycle variability need to compute long time series, with CFD tools this topic is
almost impossible nowadays. The multi-dimensional scheme is more suitable for combustion
chamber modeling, as it is more accurate based on the numerical solution of conservation of
mass, energy, momentum, and species in three dimensions [123]. Therefore, the best choice in
terms of the model to simulate SI engines depends on the results that are necessary, in order to
optimize cost and time.
Comparing quasi-dimensional models against other computational schemes, its strength
relies on its relatively easy implementation and reduced computational cost. With ordinary
computers is possible to obtain 200 simulated cycles in less than 10 minutes. This enables to
check the influence of a wide variety of parameters and thus to perform optimization studies
[125]. With respect to cyclic variability studies, in addition to its capability for reproducing
experimental phenomenology, allows obtaining long time series and statistically representative
data [126]. Therefore, several authors selected for their simulation quasi-dimensional models
due to the capabilities, with a few changes in the model, to study alternatives fuels [127].
Although numerical models of the combustion of conventional fuels in ICEs have been
developed, few modeling and simulation studies have been performed on the syngas
combustion process. Moreover, limited literature could be identified that holistically addresses
the model formulation, validation and tuning for syngas blends [125].
Syngas as part of in development fuels and with a low degree of maturity with respect
to others (hydrogen, ethanol, and butanol) represents a challenge in terms of numerical study.
149

A few authors get interested in the application of numerical models in the last years to analyze
spark ignition engines fueled with syngas [13, 21, 45, 99, 117, 124, 130, 131]. However, the
most of them prefer simple models as zero-dimensional [13, 21, 117, 124, 130].
Shivapuji et al. [130] apply a quasi-dimensional model for the study of syngas and
establish that operating a conventional fuel engine with bio-derived, represents a typical
scenario of thermo-chemistry alteration with significant influence on engine operational
response. Thus, require fuel specific tuning and validation, based on experimental results, of
the various empirical coefficient. Also, found an increase in convective cooling that is reflected
in about 32% cooling load against 25% for conventional fuels, attributed to the presence of
hydrogen. Moreover, the simulation results are in accordance within 5% of experimental results
and the versatility and robustness of quasi-dimensional approach to simulate non-regular bio-
derived alternative fuels was established.
Numerical simulations rely on model calibrations and validations which can only be
performed through comparison with high-level data acquired on research engines. In general,
the numerical analysis are compared against thermodynamic data. However, with the
development of optical techniques, more information is available to compare with numerical
analysis. A study by Irimescu et al. [76] validates a quasi-dimensional model in an optical
research engine. In this work was confirmed by analyzing optical data obtained with two flame
visualization methods that the entrainment assumption seems to be valid for SI engines. The
study also identified the scale at which mass transfer takes place as an essential factor for correct
turbulent combustion modeling. Moreover, the fact that the calibration coefficients require
different values for accurate prediction of the pressure traces during flame propagation
emphasizes the complexity of the combustion process.
Numerical approaches minimize the number of necessary experimental tests that have
to be conducted that are usually costly and time-consuming. Thus, the aim of this chapter is to
validate a quasi-dimensional model for application in a research optical engine fueled with pure
syngas (S50) and methane (M). Moreover, a comparative analysis between experimental and
simulated measured flame area is presented in the results section. The operative point was taken
as the baseline case presented in chapter 3, with 900 RPM and spark timing optimized for
methane (-7 CAD ATDC) in stoichiometric air-fuel ratio.
150

6.1 Simulation Model

The model used in this work, for the prediction of combustion behavior in SI engines,
is a quasi-dimensional two-zone model. The numerical code was developed by P.L.Curto Risso
et al. [125, 126, 131-133] and is based on the first principle of thermodynamics for open systems
coupled with several sub-model for the resolution of the pressure and temperature differential
equations. The language used for programming was FORTRAN. The model was validated for
several fuels, as gasoline [132] and others blends (gasoline-hydrogen [85] and gasoline-ethanol
[134]) in commercial engine geometries. The aim of this chapter is to implement the code for
the simulation of gaseous fuel (methane and pure syngas) in an optical SI research engine.
Therefore, is necessary to insert the fuel properties and changes in the sub-models to simulate
methane and pure syngas. In addition, geometric parameters as well as change in the sub-models
for being able to simulate the optical research engine presented in the previous chapter was
done. More details about these new parameters are explain in the next sections and Appendix B.
Focusing on the model, the cylinder inner wall delimits the control volume considered.
The working fluid is an adiabatic mixture of unburned (u) and burned (b) gases except during
combustion, where a two-zone scheme is followed. All gases are considered as ideal for
temperature dependent specific heats. The adiabatic constant for the burned gases is calculated
from the equilibrium composition at any time. Enthalpy changes are only associated with
temperature changes except during combustion, where species redistribution according to
chemical reactions are taken into account [34]. A system of differential equations for pressure
and temperature inside the combustion chamber is built, after taking particular values for the
initial conditions. These equations are valid during the evolution of the system (including the
overlapping period where intake and exhaust valves are simultaneously open) except for
combustion. A differential equation for either burned or unburned gases temperature inside the
cylinder can be written as Equation 14.

𝑑𝑇 1 𝑑𝑝
= [𝑄𝑢 ̇ + 𝑄𝑏̇ + 𝑚𝑖𝑛
̇ ℎ𝑖𝑛 + 𝑚𝑒𝑥
̇ ℎ𝑒𝑥 − 𝑚̇𝑢 ℎ𝑢 − 𝑚̇ 𝑏 ℎ𝑏 + 𝑉 ] (14)
𝑑𝑡 𝑚𝑢 𝐶𝑝,𝑢 + 𝑚𝑏 𝐶𝑝,𝑏 𝑑𝑡

In the above equation the term 𝒎̇ 𝒖 𝒉𝒖 + 𝒎̇ 𝒃 𝒉𝒃 corresponds to enthalpy changes of the


̇ 𝒉𝒊𝒏 + 𝒎𝒆𝒙
gas mixture inside the cylinder, while 𝒎𝒊𝒏 ̇ 𝒉𝒆𝒙 correspond to the enthalpy changes
̇ and 𝒎𝒆𝒙
associated with intake or exhaust processes. 𝒎𝒊𝒏 ̇ can be positive or negative depending
on the relative pressures between the interior of the cylinder and the intake or exhaust pressures.
151

Moreover, 𝑸𝒖̇ and 𝑸𝒃̇ corresponds, respectively, to the heat transferred for unburned or burned
gases. In the processes where there is no presence of unburned gases inside the cylinder, the
𝒎𝒖 = 𝟎 and then 𝑸𝒖̇ = 𝟎, and the same applies for the burned gases. Therefore, each term for
u or b can appear or not in the equation depending on the considered stroke. The initial values
are given by the external conditions. For pressure, with the same arguments, the corresponding
equation can be written as Equation 15 and Equation 16:

𝑑𝑝 𝑚̇ 𝑢 𝑚̇ 𝑏 𝑑𝑉
= [𝑝 ( + − )
𝑑𝑡 𝜌𝑢 𝜌𝑏 𝑑𝑡
(15)
1
+ 𝛿(𝑄𝑢 ̇ + 𝑄𝑏̇ +𝑚̇ 𝑖𝑛 ℎ𝑖𝑛 + 𝑚𝑒𝑥
̇ ℎ𝑒𝑥 − 𝑚̇ 𝑢 ℎ𝑢 − 𝑚̇ 𝑏 ℎ𝑏 )]
[𝑉(1 − 𝛿)]

𝑉
𝛿=
𝑉𝑢 𝐶𝑝,𝑢 𝑉𝑏 𝐶𝑝,𝑏 (16)
𝑅𝑢 + 𝑅𝑏

The model is applicable to stationary (constant crankshaft rotational velocity) or non-


stationary conditions, if it is added the mechanical equation. In this chapter the first-mentioned
case was applied, 900 RPM as constant crankshaft speed. Therefore, from the mechanical
viewpoint, the only relationship required is the variation of the combustion chamber volume at
any given instant.
For all the stages of the cycle, particular models are required as combustion, friction,
heat, and mass transfer models among others. The mass flow rates through the valves are
quantified by means of the standard equations for isentropic flow of a compressible fluid
through an orifice [34]. The intake valve discharge coefficient was taken from R. Rodrigues
da Costa et al. [135] that studied the same cylinder head geometry of the present work by the
experimental and numerical test. The discharge coefficient for exhaust valves was fixed in 0.45,
as the usual value defined in the bibliography [34] for the diameter/max-lift ratio of this engine.
The valve lift in each crankshaft position was modeled by custom equation from fabricant
camshaft data. More details about the sub-models of mass flow rates are presented in
Appendix B.
For friction components, the empirical correlation by Barnes-Moss is considered
including both linear and quadratic terms of the engine speed [136]. The calibration was done
in motored mode by measurements of torque in different crankshaft rotational velocity.
152

The heat transfer between the gas inside the cylinder and the cylinder inner wall is
modeled with the Woschni correlation [91]. It is usually used mainly because it is less intensive
with regard to computational effort, as compared to more complex models. Moreover, shows
better approximation than Annand [92] and Eichelberg models [93].
In the combustion process, two different control volumes separated by the flame front
are considered. The initial temperature of burned gases is taken as the adiabatic flame
temperature at constant pressure. Moreover, the pressure is considered as uniform during all
this period. To simulate combustion we use the turbulent model proposed by Blizard and Keck
[137, 138] and afterward improved by Beretta et al. [139]. Keck's model for turbulent
combustion starts from the laminar flame speed as a fundamental variable. However, it
incorporates turbulence effects by two characteristic parameters, 𝒖𝑻 and 𝒍𝑻 . These parameters
can be related to turbulence intensity by using Damkohler's approximation [140]. During flame
propagation, the flame front is supposed approximately spherical but not all the mass inside is
burned, there are eddies of unburned gases. They have a characteristic length, 𝒍𝑻 .
The set of coupled differential equations that give the time evolution of the total mass
inside the flame front, 𝒎𝒆 (this is the total mass inside the flame front, unburned and burned
gases) and the mass of burned gases, 𝒎𝒃 , is written as:

(𝑚𝑒 − 𝑚𝑏 )
𝑚̇𝑏 = 𝐴𝑓 𝜌𝑢 𝑆𝐿 + (17)
𝜏𝑏

𝑚̇𝑒 = 𝐴𝑓 𝜌𝑢 [𝑢𝑡 + 𝑆𝐿 ] (18)

in Equation 17, 𝒕 represents the time elapsed from the beginning of combustion, 𝑨𝒇 , is the area
of the spherical flame front, calculated from its radius, and this in turn from the volume of gases
inside the flame front 𝑽𝒇 [132]. This volume is given by 𝑽𝒇 = 𝑽 − (𝒎 − 𝒎𝒆 )/𝝆𝒖 where 𝑽 is
the total cylinder volume, 𝝆𝒖 , the density of unburned gases and 𝒎, the total mass in the
cylinder. The first term at the right of Equation 18 represents the laminar propagation of the
flame front, and the rightmost the combustion of the fresh mixture inside the flame front. The
time parameter 𝝉𝒃 = 𝒍𝒕 /𝑺𝑳 , is the characteristic time required for the combustion of a eddie of
characteristic length, 𝒍𝒕 , at an speed 𝑺𝑳 , which is the laminar flame speed. Moreover, 𝝉𝒃 also
represents the time the flame front needs to develop into a turbulent flame from the initial
conditions: laminar flame and spherical symmetry [34]. Equation 18 describes the total mass
inside the flame front rate that evolves at a velocity 𝒖𝒕 + 𝑺𝑳 (𝝉 ≫ 𝝉𝒃 ), where 𝒖𝒕 is the
153

characteristic speed of the mixture when crossing the flame front. During combustion, these
equations are coupled to the thermodynamic ones.
In this work, two fuels were used for numerical simulation, methane as baseline fuel and
S50 as the most representative fuel to the family of pure syngas, both working in stoichiometric
condition. For methane, the laminar speed 𝑺𝑳 from several in-cylinder conditions was taken
from Gottengs et al. [141], that is based in an asymptotic analysis with follow approximation
formula:

𝑛
−𝐺 𝑇𝑢 𝑇𝑏 − 𝑇 0
𝑆𝐿,𝑚𝑒𝑡ℎ𝑎𝑛𝑒 = F (𝑌 𝑚 )𝑒 𝑇 0 ( 0) ( ) (1 − 2.1𝑦𝑟 0.77 ) (19)
𝑇 𝑇𝑏 − 𝑇𝑢

where F, G, m, n are constant values extracted from Muller et al. [142]. Moreover, 𝑻𝟎 is the
inner layer temperature, representing the crossover temperature between chain-branching and
chain-breaking reactions. Inner layer temperature is a function of pressure in the combustion
chamber [141]. Subsequent, Y is the mass fraction of the fuel in the unburnt gas and the
temperatures 𝑻𝒖 and 𝑻𝒃 are those in the unburnt and the burnt gas, respectively. Moreover,
𝒚𝒓 is the mole fraction of residual gases from the previous cycle that are still in the combustion
chamber.
In the case of pure syngas the dependence on temperature and pressure of the laminar
flame speed was calculated from the correlation by Blizard and Keck [138], generally used for
blend fuels:

𝛼 𝛽
𝑇𝑢 𝑝
𝑆𝐿,𝑆50 = 𝑆𝐿𝑂 ( ) ( ) (1 − 2.1𝑦𝑟 0.77 ) (20)
𝑇𝑟𝑒𝑓 𝑝𝑟𝑒𝑓

where 𝒑 is the pressure in the combustion chamber, and 𝑻𝒓𝒆𝒇 , and 𝒑𝒓𝒆𝒇 , reference values. For
S50, the reference laminar speed, 𝑺𝑳𝑶 , and the exponents α and β were taken from CFD
simulation results by Fanelli et al. [105] that fitted Lieuwen et al. [143] experimental data. The
equations that fit the results for different λ values are shown in Equation 21 to 23. A resume of
the constant parameters used in Equation 19 and Equation 20 is shown in Table 25.
154

1 2 1
𝑆𝐿𝑂 = 9.554 ( ) + 166.790 ( ) − 62.303 (21)
𝜆 𝜆

1 2 1
α = −0.586 ( ) + 0.327 ( ) + 2.020 (22)
𝜆 𝜆

1 6 1 5 1 4
β = −163.0378 ( ) + 834.277 ( ) − 1778.405 ( )
𝜆 𝜆 𝜆
1 3 1 2 1 (23)
+ 2023.483 ( ) − 1298.163 ( ) + 448.404 ( )
𝜆 𝜆 𝜆
− 64.848

Table 25 – Constant parameters used for the calculus of laminar flame speed of methane and
S50.

Parameter Value
F [m/s] 0.22
G [K] -6444.27
m 0.565175
n 2.5158
𝑇𝑟𝑒𝑓 [K] 298
𝑝𝑟𝑒𝑓 [Pa] 1.0x105

In order to calculate 𝑢𝑡 and 𝑙𝑡 empirical correlation developed by Beretta [139] is used


shows in Equation 24 and Equation 25.

𝜌𝑢 0.5
𝑢𝑡 = 𝐶𝑡 0.08 𝑈𝑖 ( ) (24)
𝜌𝑖

𝜌𝑖 0.75
𝑙𝑡 = 𝐶𝑙 0.8 𝐿𝑉𝑚𝑎𝑥 ( ) (25)
𝜌𝑢

Where 𝑼𝒊 is the average inlet gas speed, 𝝆𝒊 is the average density of the inlet gas and
𝑳𝑽𝒎𝒂𝒙 is the maximum lift of intake valves. Moreover, 𝑪𝒕 and 𝑪𝒍 represents the tuning
coefficient for turbulent combustion, Beretta [139] showed for gasoline in a two valve cylinder
head (one valve for intake and one for exhaust) that these two values are equal to one. However,
its need to be adjusted for each engine geometry and fuel used. In the present study, was
considered the two calibration coefficients as the only parameters to be modified for its
155

application, given that they have a fundamental basis for describing entrainment and fuel
oxidation phenomena [76].
In the validation section, for each average pressure trace that corresponded to a specific
fuel (methane and S50), simulations were performed by a range of values for both coefficients.
The optimum combination of calibrated values was identified by the minimum difference
between experimental and simulated data. Apart from the fact that the approach made the
calibration process very much straightforward, it also emphasizes the correlation between the
two main factors that influence combustion, namely fluid motion and chemical kinetics [144].
Another issue that is important for simulation accuracy is how ignition is handled [145],
given the complexity of kernel formation [76]. No ignition sub-model was applied for this
study; rather, the enflamed volume at spark timing was set as that of a semi-sphere. The
dimensions were fit with the data taken from the flame visualization experiments for each fuel.
Optical engines as was explained in the previous chapter have some limitations with
respect to commercial engines, due to the substitution of metal walls by optical one. The most
influence variations are the position and material of the piston rings that allows the passage of
un-burned mass to the top land region and crankcase. For the top land region, a simplification
was used with a reduction of the geometrical compression ratio (CR) to fit the motored pressure
curve. This assumption means that top land region is assumed as part of the dead volume. This
reduce the range of application of the computational model, because the effects of mass loss in
the top land, which it does not take in account, will affect the thermodynamic state in the final
stages of combustion.
For the chemical reaction associated with the engine combustion, chemical equilibrium
conditions with dissociation were considered for a mixture of fuel, dry air, and previous cycle
reaction products. While this assumption can be viewed as a source of errors in the simulations,
several works shows that the main influence on combustion modeling is through laminar flame
speed (SL) and temperature of unburned gases (Tu) as pointed out by several authors [94, 99].
Moreover, note that the stated chemical reaction do not consider explicitly a detailed
composition of combustion products with respect to hydrogen. This is because the main focus
of this work is placed on the energetics components. Furthermore, the explicit estimation of
NOx emissions would require non-equilibrium reaction models that are out of the scope of this
work. Therefore, the effect of the hypothesis of chemical equilibrium is negligible on the
calculations hereafter shows in the work. We use a method presented by Ferguson [36] to solve
chemical equilibrium for ten species, but considering residual gases.
156

6.2 Validation

The simulation scheme previously described was validated by comparing with the
experimental results showed in previous chapters. The engine speed for simulations was taken
at 900 RPM, the spark advance at 7 CAD BTDC and the relative air-fuel ratio at stoichiometric
value (λ=1.0). The validation process was performed with methane as baseline fuel, to calibrate
CR among other general parameters. Finally, for both fuels (methane and S50) turbulent
coefficients were tuned to reach similar pressure curves. Several researchers used this approach
in optical [76] and commercial engines [127, 134] due to variations in turbulent condition
depends on the fuel and operating condition used. Table 26 shows complementary information
to Table 4 for engine and experimental condition used in the numerical code.

Table 26 – Complementary specifications of the set up used in the PFI-SI single-cylinder


research engine AVL 5406.

Component Size Unit


Geometric Compression Ratio 9.7:1
Fit Compression Ratio 8.3:1
Spark distance 5.0 mm
Ambient pressure 0.950 bar
Intake pressure 0.945 bar
Exhaust pressure 1.021 bar
Ambient temperature 298 K
Wall temperature 333 K
Exhaust temperature 673 K

Figure 90 shows the simulated pressure traces compared to the measured ones for
methane and syngas with 50% H2 in stoichiometric air-fuel ratio. Practically all numerical traces
featured very good accuracy during flame propagation (i.e. from spark timing to the point of
peak pressure of experimental curve), while during the last expansion pressure levels were over-
predicted. This is due to the procedure used for calibration but allows a more direct comparison
between simulated flame area and the values obtained through optical measurements.
157

(a) (b)

Figure 90- Experimental and simulated pressure traces for stoichiometric air-fuel ratio
and SA -7 CAD ATDC for: a) methane and b) S50.

In addition, mass fraction burned was compared between experimental and simulated
[76], similar behavior than pressure curves were seen. The model predicted the ratio of burned
mass up to 40%, this corresponds to 10 CAD ASOS for syngas and 25 CAD ASOS for methane.
Another interesting point is that when compare to Figure 29 (flame area for stoichiometric air-
fuel ratio and spark timing 7 CAD BTDC) both fuels reach the optical window before the
position that the model not predicted well the MFB. In spite of the model not predict with
accuracy the last combustion phase, the first part is enough to compare with optical data.
To improve the results of the expansion stroke, it is suggested by some authors to apply
a model to account blow-by losses. A simplified compressible flow model was employed for
Irimescu et al. [57]. For a good prediction of thermodynamic parameters as IMEP or efficiency
is necessary to take in account the effects of unburned mass trapped in the top land region and
the increase in heat transfer. However, for this analysis was preferred to compare only in the
first phase of the combustion for simplicity.
The model presented in the current chapter was used in commercial engine chamber
geometries with good accuracy of the results in all the combustion process for several fuels.
Generally, this type of engine has mass transfer for the crankcase below 2% of the total fuel
mass. However, optical engines was found over 20% [57]. In addition, could be seen that any
author found in the bibliography has an accuracy pressure prediction when quasi-dimensional
158

model is used in optical engines as could be seen for commercial engines. This information
confirms that improves models need to be developed to quasi-dimensional models be capable
to correct simulate the all combustion process in this engine geometries.

(a) (b)

Figure 91- Experimental and simulated mass fraction burned traces for stoichiometric
air-fuel ratio and SA -7 CAD ATDC for: a) methane and b) S50.

6.3 Results

Figure 92 shows the comparison between simulated and flame area measured with the
high-speed camera coupled with intensifier; just for clarity, this parameter refers to the area of
the flame projected onto the piston crown, normalized to the cross-section surface of the
cylinder bore, including both burned gas and the reaction zone. For both fuels, the simulated
traces seem to be higher than the experimental ones. Irimescu et al. [76] also seen the same
trend for gasoline in several crankshaft rotational velocity and air-fuel ratios. In the mentioned
work, this behavior was explained (at least in part) by the fact that wrinkling results in a larger
flame front area.
Another point is that methane presented larger differences than syngas (Figure 93), one
possible explanation is that flame displacement was quite significant for this fuel. Therefore,
for the same rate of mass entrainment (i.e. equivalent flame front area), the actual flame area
needs to be larger, resulting in part of it touching the walls early during the propagation phase.
159

With the assumption of a spherical flame front propagating outward from the central position
of the spark plug, the model predicts the pressure trace with lower flame area [76]. This further
underlines the complexity of flame front propagation and fuel oxidation.

Figure 92 – Comparison between simulated and experimental normalized flame area for
stoichiometric air-fuel ratio and SA -7 CAD ATDC.

The correlation between volume fraction burned and the flame area can be considered
as a good indicator of the model’s accuracy (Figure 94), because compare the thermodynamic
evaluation of flame size (volume fraction burned) with respect to the measured by the optical
apparatus (normalized flame area). As a general conclusion of the comparison between
numerical and experimental results, the accuracy of the calibrated model can be considered as
good during flame propagation for all conditions that were investigated. As was seen for
normalized flame area against CAD, pure syngas shows better results than methane. For this
last fuel, the model accuracy was seen better up to 25% of VFB after that the walls effects
seems to be greater.
160

20.0 1.0
exp sim exp sim

Normalized Flame Area


CAD [ASOS] 16.0 0.8

12 Optical Limit
12.0 0.6
10
0.41
8.0 0.4
0.29
5
4
4.0 0.2
0.08
0.03
0.0 0.0
M S50 M S50
a) b)

Figure 93 - Comparison of the average measured (exp) and simulated (sim) results recorded
during 25 consecutive cycles for a) CAD position of 30% of the normalized flame area and
b) flame diameter in 5 CAD ASOS, at 900 RPM and spark timing of 7 CAD BTDC.

a) b)

Figure 94 – Flame area against volume fraction burned for simulated and experimental
measurements for stoichiometric air-fuel ratio and fixed spark timing for a) methane b) S50.
161

6.4 Summary

A two-zone model was applied for simulating the combustion process in an optical SI
engine for pure syngas and methane in stoichiometric air-fuel ratio and fixed spark timing
settings. In-cylinder pressure measurements were used for calibrating the model. The
entrainment concept was the essential starting point for the turbulent flame propagation study,
with optical data recorded used for analyze the numerical results.
Good accuracy in the first phase of the combustion was ensured through the calibration
procedure based on measured in-cylinder pressure. Turbulent experimental coefficients 𝑪𝒕 and
𝑪𝒍 was tuned to simulate pressure up to the maximum of the experimental case. After this point,
the simulated traces were greater due to the not consideration of mass transfer from the
combustion chamber to the top land region and crankcase. The application of an accuracy
crevice model is necessary for the good prediction of pressure traces in the expansion phase.
However, it is possible to obtain interesting information for this first phase of the combustion
process. It important to notice that optical visualization works in this phase of the combustion
process.
Therefore, an analysis was performed using the flame area recorded in the UV–visible
wavelength range. The variation of the two calibration coefficients suggests that engine
operating parameters have a significant influence on the scale at which burn-up occurs behind
the flame front. A good correlation between the modeled and measured flame area values
confirmed that the spherical propagating kernel is a suitable assumption and that flame front
area increase due to wrinkling plays a minor role in the entrainment process. The scale at which
mass transfer takes place between the unburned-reacting zones and the size of reacting gas
pockets are found to be the main areas of improvement for 0D models, in order to ensure that
they are more predictive with reduced calibration requirements. Another important conclusion
of the study is that the effects of flame displacement need to be better described by the
combustion models, especially in cases of methane in which the spherical hypothesis seems to
be not precise for this case of study.
162

7 Conclusions

An experimental study was undertaken in order to study the combustion process with
pure syngas (CO, H2 blends) and real syngas mixtures (CO, H2, NO2, CO2 blends). For this
reason, an optically PFI-SI accessible engine was used in order to integrate the results from the
thermodynamic analysis, based on in-cylinder pressure measurements, with cycle resolved 2D
flame visualization (high spatial and temporal resolution). The experiments were carried out at
low speed (900 RPM) and wide-open throttle, the typical working point of SI engines for
stationary applications. A numerical code was used to evaluate the in-cylinder pressure data
and obtain other thermodynamics parameters as MFB, fuel conversion efficiency among others.
Moreover, exhaust measurements were taken to obtain comparative results in terms of pollutant
emissions from fuels and operative conditions. On the other hand, optical diagnostic, supported
by a custom image processing methodology, allows characterizing the flame front propagation
in terms of size, velocity, shape, and displacement.
From the study of pure syngas (chapter 3) in several AFRrel, the highest in-cylinder
pressure was measured for pure syngas mixtures in all operative condition. The syngas presents
higher fuel conversion efficiency and IMEP value than methane in λ=1.4. However, the baseline
case presents better performance and combustion stability for λ=1.0 and λ=1.2. This was
directly associated to not spark timing optimization for pure syngas cases. Exhaust emission
shows that for the same AFRrel (λ=1.4) syngas presents higher NOx emissions than methane,
due to higher combustion temperatures present in the combustion chamber. The optical analysis
shows that a large difference exist in terms of flame propagation speed between pure syngas
and methane. Moreover, in the case of λ=1.4, methane suffers higher decrease than syngas due
to methane is closer to the flammability limit. Therefore, hydrogen content ensures high flame
speed, also in high air dilution condition. Flame distortion was found higher for methane than
syngas for all cases. Finally, syngas present lower displacement than methane. All flame
parameters behavior is directly associated with the hydrogen addition, which enhances the
flame propagation speed and therefore a better combustion process.
An increase of fuels (methane-H2 blends) and operative condition (extreme lean
conditions) was performed in the last section of chapter 3. Thermodynamic analysis reveals that
hydrogen addition improves the flammability range of CH4 (λ=1.4), with a maximum AFRrel of
1.50 for 50% H2 in CH4 (M50). For syngas mixtures, this range was even larger, with AFRrel
163

1.75 for 75% H2 in CO (S75). The five fuels that were tested ensured acceptable engine
performance, even if the spark timing setting was kept at the MBT point for methane at λ=1.4.
The analysis of exhaust gas emissions confirmed the potential of extremely lean operation for
reducing NOx emissions. Cycle resolved visualization shows that the mixture with the highest
concentration of hydrogen S75 featured the fastest flame propagation. Moreover, hydrogen
content improves the flame speed of methane for both contents tested. However, the influenced
of hydrogen in CO was higher than CH4 due to the basic properties of the components. In
addition, flame distortion and center movement decrease when H2 was added due to improving
on flame propagation speed. One major conclusion of chapter 3 was that spark timing re-
calibration is required in order to fully take advantage of fuel properties such as higher laminar
flame speed and increased stability.
The target of Chapter 4 was to improve the knowledge of the combustion process when
pure syngas and methane-H2 blends are used to replace fossil energy sources in different spark
timing configurations. The effect of different spark timing for S50 and fixed AFRrel (λ=1.0) was
studied in section 4.1. In section 4.2 was included a comparison of five different blends in
maximum brake torque condition (optimum spark timing) and lean conditions (λ=1.4 to
λ=1.75). The optimum SA for S50 at stoichiometric AFRrel was seen in 8 CAD ATDC. The
behavior is different from conventional fossil fuels, in which generally engine mapping works
in values before the top dead center. This is due to the short burning duration with respect to
methane. The peak in-cylinder pressure and NOx emissions decrease with the retard of the spark
timing. Optical investigations demonstrated that flame propagation speed and flame distortion
decrease with the retard of SA. The flame center displacement was similar for the six spark
timing condition. This is due to the fast propagation speed of S50 at λ=1.0, with an average of
22 m/s.
The last section of chapter 4 reveals that pure syngas featured higher peak pressure and
similar pressure curves in the last phase of the combustion process than methane blends.
Therefore, higher IMEP values were obtained, (around 9%), than methane blend for the same
hydrogen content. The five fuels that were tested ensured acceptable engine performance and
stability for λ=1.4. Therefore, spark timing re-calibration allows to fully taking advantage of
fuel properties such as lower burning duration and increased stability. In line with the
thermodynamic results, cycle resolved images shows that S75 presents the fastest flame
propagation due to the high H2 content. However, the flame speed decrease with respect to the
advanced case for all fuels. Propagation was found to be quite symmetric in all directions for
syngas (lower than 1.2), and similar for all methane blends (around 1.4), suggesting that flame
164

speed had the most important effect. In terms of flame centroid displacement, the higher effect
of spark timing was seen for the extreme lean condition, in which the advanced of the SA
increase the distance with respect to the combustion chamber.
The last experimental study was shown in chapter 5. This study was undertaken in order
to evaluate the effects of fuel dilution in two H2/CO ratios. The overall thermodynamic analysis
revealed that the four syngas mixtures ensured acceptable engine stability and similar engine
performance for the three air dilution tested, even if the spark timing setting was kept at the
MBT point for methane. A reduction of 25% in IMEP was seen for syngas with respect to
methane in stoichiometric condition. However, this difference was reduced for pure syngas
(almost similar performance) for lean mixtures. Diluted syngas mixtures due to not optimization
of spark timing and a decrease in terms of volumetric efficiency maintain the lower values of
IMEP with respect to pure methane. A brief analysis of fuel conversion efficiency showed that
for stoichiometric air-fuel ratio diluted mixtures presented better results than pure syngas due
to the large difference between set SA and MBT point. In addition, diluted syngas shows similar
value in terms of efficiency than methane. The addition of air dilution shows a greater effect in
diluted syngas with efficiency below 18%. The pure syngas mixtures featured the fastest flame
propagation follow by the diluted syngas and lastly pure methane. Propagation was found to be
quite symmetric in all directions and similar for all fuels with respect to methane, suggesting
that H2 content and fluid motion had the most important effect in this sense. These results were
also confirmed by the evolution of flame centroid displacement, more extensive for lean air-
fuel ratios, but nonetheless comparable.
Finally, a numerical investigation in the combustion process of pure syngas was
performed. A two-zone model was applied for simulating the initial phase of the combustion
process in an optical SI engine. Pure syngas and methane in stoichiometric air-fuel ratio and
fixed spark timing was studied. A good correlation between the modeled and measured flame
area values confirmed that the spherical propagating kernel is a suitable assumption and that
flame front area increase due to wrinkling plays a minor role in the entrainment process. The
scale at which mass transfer takes place between the unburned-reacting zones and the size of
reacting gas pockets are found to be the main areas of improvement for 0D models, in order to
ensure that they are more predictive with reduced calibration requirements. Another important
conclusion of the study is that the effects of flame displacement need to be better described by
combustion models, especially in cases of methane.
165

8 Bibliography

[1] AGARWAL, A. K.; PANDEY, S. D. A.; SINGH, A. P. Combustion for Power


Generation and Transportation. Singapore: Springer Singapore, 2017.
[2] MAGALHÃES, E. M. Combustion Study of Mixtures Resulting from a Gasification
Process of Forest Biomass. [s.l.] École Nationale Supérieure de Mécanique et
D’Aerotechnique, 2011.
[3] KUMAR, A.; JONES, D. D.; HANNA, M. A. Thermochemical Biomass Gasification:
A review of the Current Status of the Technology. Energies, v. 2, n. 3, p. 556–581,
2009.
[4] PARIKKA, M. Global biomass fuel resources. Biomass and Bioenergy, v. 27, n. 6, p.
613–620, 2004.
[5] GATES, B. C. et al. Catalysts for emerging energy applications. MRS Bulletin, v. 33,
n. 4, p. 429–435, 2008.
[6] DE FILIPPIS, P. et al. Gasification process of Cuban bagasse in a two-stage reactor.
Biomass and Bioenergy, v. 27, n. 3, p. 247–252, 2004.
[7] SAHOO, A.; RAM, D. K. Gasifier performance and energy analysis for fluidized bed
gasification of sugarcane bagasse. Energy, v. 90, n. October 2017, p. 1420–1425, 2015.
[8] SUSASTRIAWAN, A. A. P.; SAPTOADI, H.; PURNOMO. Small-scale downdraft
gasifiers for biomass gasification: A review. Renewable and Sustainable Energy
Reviews, v. 76, p. 989–1003, set. 2017.
[9] MARTÍNEZ, J. D. et al. Syngas production in downdraft biomass gasifiers and its
application using internal combustion engines. Renewable Energy, v. 38, n. 1, p. 1–9,
2012.
[10] PRZYBYLA, G. et al. Fuelling of spark ignition and homogenous charge compression
ignition engines with low calorific value producer gas. Energy, v. 116, p. 1464–1478,
2016.
[11] SHIVAPUJI, A. M.; DASAPPA, S. In-cylinder investigations and analysis of a SI gas
engine fuelled with H2and CO rich syngas fuel: Sensitivity analysis of combustion
descriptors for engine diagnostics and control. International Journal of Hydrogen
Energy, v. 39, n. 28, p. 15786–15802, 2014
[12] RAMAN, P.; RAM, N. K. Performance analysis of an internal combustion engine
operated on producer gas, in comparison with the performance of the natural gas and
diesel engines. Energy, v. 63, p. 317–333, 2013.
[13] CENTENO, F. et al. Theoretical and experimental investigations of a downdraft
biomass gasifier-spark ignition engine power system. Renewable Energy, v. 37, n. 1,
p. 97–108, 2012.
[14] BIKA, A. S.; FRANKLIN, L.; KITTELSON, D. B. Engine knock and combustion
characteristics of a spark ignition engine operating with varying hydrogen and carbon
monoxide proportions. International Journal of Hydrogen Energy, v. 36, n. 8, p.
5143–5152, 2011.
[15] SAMIRAN, N. A. et al. Progress in biomass gasification technique – With focus on
Malaysian palm biomass for syngas production. Renewable and Sustainable Energy
Reviews, v. 62, p. 1047–1062, set. 2016.
[16] CIFERNO, J. P.; MARANO, J. J. Benchmarking biomass gasification technologies
for fuels, chemicals and hydrogen productionUS Department of Energy. National
166

Energy. [s.l: s.n.]. Disponível em:


<http://seca.doe.gov/technologies/coalpower/gasification/pubs/pdf/BMassGasFinal.pd
f>.
[17] GIL, J. et al. Biomass gasification in atmospheric and bubbling fluidized bed: Effect of
the type of gasifying agent on the product distribution. Biomass and Bioenergy, v. 17,
n. 5, p. 389–403, nov. 1999
[18] GOBBATO, P.; MASI, M.; BENETTI, M. Performance analysis of a producer gas-
fuelled heavy-duty SI engine at full-load operation. Energy Procedia, v. 82, p. 149–
155, 2015.
[19] BANAPURMATH, N. R.; TEWARI, P. G. Comparative performance studies of a 4-
stroke CI engine operated on dual fuel mode with producer gas and Honge oil and its
methyl ester (HOME) with and without carburetor. Renewable Energy, v. 34, n. 4, p.
1009–1015, abr. 2009.
[20] RAMADHAS, A. S.; JAYARAJ, S.; MURALEEDHARAN, C. Dual fuel mode
operation in diesel engines using renewable fuels: Rubber seed oil and coir-pith
producer gas. Renewable Energy, v. 33, n. 9, p. 2077–2083, set. 2008.
[21] TINAUT, F. V. et al. Method for predicting the performance of an internal combustion
engine fuelled by producer gas and other low heating value gases. Fuel Processing
Technology, v. 87, n. 2, p. 135–142, 2006.
[22] BRIDGWATER, A. V.; MEIER, D.; RADLEIN, D. An overview of fast pyrolysis of
biomass. Organic Geochemistry, v. 30, n. 12, p. 1479–1493, 1999.
[23] BEENACKERS, A. A. C. M. Biomass gasification in moving beds, a review of
European technologies. Renewable Energy, v. 16, n. 1–4, p. 1180–1186, jan. 1999.
[24] KIRUBAKARAN, V. et al. A review on gasification of biomass. Renewable and
Sustainable Energy Reviews, v. 13, n. 1, p. 179–186, jan. 2009.
[25] BRIDGWATER, A. V. The technical and economic feasibility of biomass gasification
for power generation. Fuel, v. 74, n. 5, p. 631–653, maio 1995.
[26] MEHRLING, P.; VIERRATH, H. Gasification of lignite and wood in the Lurgi
circulating fluidized-bed gasifier: Final report. United States: [s.n.]. Disponível em:
<https://www.osti.gov/biblio/5955631>.
[27] NARVÁEZ, I. et al. Biomass Gasification with Air in an Atmospheric Bubbling
Fluidized Bed. Effect of Six Operational Variables on the Quality of the Produced Raw
Gas. Industrial & Engineering Chemistry Research, v. 35, n. 7, p. 2110–2120, jan.
1996.
[28] HOFBAUER, H. et al. Two years experience with the FICFB-gasification process. 10th
European Conference and Technology Exhibition, Wurzburg, n. January, p. 3–6,
1998.
[29] ISING, M.; GIL, J.; UNGER, C. Gasification of biomass in a circulating fluidized
bed with special respect to tar reduction. 1st World Conference and exhibition on
biomass for energy and Industry. Anais...Sevilla, Spain: 2000.
[30] GREDINGER, A.; SPÖRL, R.; SCHEFFKNECHT, G. Comparison measurements of
tar content in gasification systems between an online method and the tar protocol.
Biomass and Bioenergy, v. 111, p. 301–307, abr. 2018.
[31] ABE, H. et al. Potential for rural electrification based on biomass gasification in
Cambodia. Biomass and Bioenergy, v. 31, n. 9, p. 656–664, 2007.
[32] SIEMONS, R. V. Identifying a role for biomass gasiÿcation in rural electriÿcation in
developing countries : the economic perspective. Biomass and Bioenergy, v. 20, p.
271–285, 2001.
[33] HAGOS, F. Y.; AZIZ, A. R. A.; SULAIMAN, S. A. Syngas (H2/CO) in a spark-ignition
direct-injection engine. Part 1: Combustion, performance and emissions comparison
167

with CNG. International Journal of Hydrogen Energy, v. 39, n. 31, p. 17884–17895,


2014a.
[34] HEYWOOD, J. B. Internal Combustion Engine Fundamentals. New York, USA:
McGraw-Hill, 1988.
[35] STONE, R. Solutions Manual for Introduction to Internal Combustion Engines.
London: Macmillan Education UK, 1999.
[36] FERGUSON, C. R.; KIRKPATRICK, A. T. Internal Combustion Engines: applied
thermosciences. 3th. ed. Colorado, USA: [s.n.].
[37] KARIM, G. A.; WIERZBA, I. Comparative Studies of Methane and Propane as
Fuels for Spark Ignition and Compression Ignition Engines. SAE Technical Paper
831196. Anais...8 ago. 1983Disponível em: http://papers.sae.org/831196/.
[38] ROUSSEAU, S.; LEMOULT, B.; TAZEROUT, M. Combustion characterization of
natural gas in a lean burn spark-ignition engine. (I. SAGE PUBLICATIONS,
Ed.)Proceedings of the Institution of Mechanical Engineers:Journal of Automobile
Engineering, Part D; London. Anais...United States, London: Copyright Mechanical
Engineering Publications, Ltd. 1999, 1999Disponível em:
https://search.proquest.com/docview/220657837?accountid=26674.
[39] HUANG, Z. et al. Combustion characteristics of a direct-injection engine fueled with
natural gas-hydrogen blends under different ignition timings. Fuel, v. 86, n. 3, p. 381–
387, 2007.
[40] ZHAO, J. et al. Effects of compression ratio on the combustion and emission of a
hydrogen enriched natural gas engine under different excess air ratio. Energy, v. 59, n.
x, p. 658–665, 2013.
[41] DI IORIO, S.; SEMENTA, P.; VAGLIECO, B. M. Experimental investigation on the
combustion process in a spark ignition optically accessible engine fueled with
methane/hydrogen blends. International Journal of Hydrogen Energy, v. 39, n. 18,
p. 9809–9823, 2014.
[42] CATAPANO, F. et al. Characterization of CH4 and CH4/H2 Mixtures Combustion in a
Small Displacement Optical Engine. SAE International Journal of Fuels and
Lubricants, v. 6, n. 1, p. 2013-01–0852, 8 abr. 2013.
[43] DI IORIO, S.; SEMENTA, P.; VAGLIECO, B. M. Analysis of combustion of methane
and hydrogen–methane blends in small DI SI (direct injection spark ignition) engine
using advanced diagnostics. Energy, v. 108, p. 99–107, 2016.
[44] ALEIFERIS, P. G.; BEHRINGER, M. K. Flame front analysis of ethanol, butanol, iso-
octane and gasoline in a spark-ignition engine using laser tomography and integral
length scale measurements. Combustion and Flame, v. 162, n. 12, p. 4533–4552, 2015.
[45] SHIVAPUJI, A. M.; DASAPPA, S. Analysis of thermodynamic scope engine
simulation model empirical coefficients: Suitability assessment and tuning of
conventional hydrocarbon fuel coefficients for bio syngas. International Journal of
Hydrogen Energy, v. 42, n. 26, p. 16834–16854, 2017a.
[46] HAGOS, F. Y. et al. Effect of fuel injection timing of hydrogen rich syngas augmented
with methane in direct-injection spark-ignition engine. International Journal of
Hydrogen Energy, v. 42, n. 37, p. 23846–23855, 2017.
[47] ARROYO, J. et al. Experimental study of ignition timing and supercharging effects on
a gasoline engine fueled with synthetic gases extracted from biogas. Energy
Conversion and Management, v. 97, p. 196–211, 2015.
[48] SHIVAPUJI, A. M.; DASAPPA, S. Influence of fuel hydrogen fraction on syngas
fueled SI engine: Fuel thermo-physical property analysis and in-cylinder experimental
investigations. International Journal of Hydrogen Energy, v. 40, n. 32, p. 10308–
10328, 2015.
168

[49] HAGOS, F. Y.; AZIZ, A. R. A.; SULAIMAN, S. A. Effect of air-fuel ratio on the
combustion characteristics of syngas (H2:CO) in direct-injection spark-ignition engine.
Energy Procedia, v. 61, n. Di, p. 2567–2571, 2014b.
[50] TSIAKMAKIS, S. et al. Experimental study of combustion in a spark ignition engine
operating with producer gas from various biomass feedstocks. Fuel, v. 122, p. 126–139,
2014.
[51] ARROYO, J. et al. Combustion behavior of a spark ignition engine fueled with synthetic
gases derived from biogas. Fuel, v. 117, n. PART A, p. 50–58, 2014.
[52] MUSTAFI, N. et al. Spark-ignition engine performance with “Powergas” fuel (mixture
of CO/H2): A comparison with gasoline and natural gas. Fuel, v. 85, n. 12–13, p. 1605–
1612, set. 2006.
[53] SRIDHAR, G.; PAUL, P. J.; MUKUNDA, H. S. Biomass derived producer gas as a
reciprocating engine fuel - An experimental analysis. Biomass and Bioenergy, v. 21,
n. 1, p. 61–72, 2001.
[54] BIKA, A. S.; FRANKLIN, L.; KITTELSON, D. B. Homogeneous charge compression
ignition engine operating on synthesis gas. International Journal of Hydrogen
Energy, v. 37, n. 11, p. 9402–9411, 2012.
[55] MEROLA, S. S. et al. Optical characterization of combustion processes in a DISI engine
equipped with plasma-assisted ignition system. Applied Thermal Engineering, v. 69,
n. 1–2, p. 177–187, 2014.
[56] WINKLHOFER, E. Optical access and diagnostic techniques for internal combustion
engine development. Journal of Electron Imaging, v. 10, n. 3, p. 134–140, 2001.
[57] IRIMESCU, A. et al. Compression ratio and blow-by rates estimation based on motored
pressure trace analysis for an optical spark ignition engine. Applied Thermal
Engineering, v. 61, n. 2, p. 101–109, 2013.
[58] MARTINEZ, S. et al. Flame front propagation in an optical GDI engine under
stoichiometric and lean burn conditions. Energies, v. 10, n. 9, 2017.
[59] IRIMESCU, A.; MEROLA, S. S. Spark discharge and flame inception analysis through
spectroscopy in a DISI engine fuelled with gasoline and butanol. IOP Conference
Series: Materials Science and Engineering, v. 252, n. 1, 2017.
[60] IRIMESCU, A. et al. Effect of coolant temperature on air–fuel mixture formation and
combustion in an optical direct injection spark ignition engine fueled with gasoline and
butanol. Journal of the Energy Institute, v. 90, n. 3, p. 452–465, 2017a.
[61] MEROLA, S. S. et al. Effect of injection timing on combustion and soot formation in a
direct injection spark ignition engine fueled with butanol. International Journal of
Engine Research, p. 1468087416671017, 2016.
[62] COSTA, M. et al. Mixture preparation and combustion in a GDI engine under
stoichiometric or lean charge: an experimental and numerical study on an optically
accessible engine. Applied Energy, v. 180, p. 86–103, 2016.
[63] MEROLA, S. S.; TORNATORE, C.; IRIMESCU, A. Cycle-resolved visualization of
pre-ignition and abnormal combustion phenomena in a GDI engine. Energy
Conversion and Management, v. 127, p. 380–391, 2016.
[64] CATAPANO, F.; SEMENTA, P.; VAGLIECO, B. M. Air-fuel mixing and combustion
behavior of gasoline-ethanol blends in a GDI wall-guided turbocharged multi-cylinder
optical engine. Renewable Energy, v. 96, p. 319–332, 2016.
[65] IRIMESCU, A. et al. Combustion process investigations in an optically accessible DISI
engine fuelled with n-butanol during part load operation. Renewable Energy, v. 77, p.
363–376, 2015a.
169

[66] MEROLA, S. S. et al. UV-visible Optical Characterization of the Early Combustion


Stage in a DISI Engine Fuelled with Butanol-Gasoline Blend. SAE International
Journal of Engines, v. 6, n. 4, p. 2013-01–2638, 2013.
[67] TORNATORE, C.; MEROLA, S.; SEMENTA, P. Combustion Process Investigation
in a Small SI Engine using Optical Diagnostics. SAE Paper 2010-01-2262, 2010.
Anais...25 out. 2010Disponível em: http://papers.sae.org/2010-01-2262/.
[68] ALEIFERIS, P. G. et al. Flame chemiluminescence studies of cyclic combustion
variations and air-to-fuel ratio of the reacting mixture in a lean-burn stratified-charge
spark-ignition engine. Combustion and Flame, v. 136, n. 1–2, p. 72–90, 2004.
[69] ALEIFERIS, P. G. et al. Insights into Stoichiometric and Lean Combustion Phenomena
of Gasoline–Butanol, Gasoline–Ethanol, Iso-Octane–Butanol, and Iso-Octane–Ethanol
Blends in an Optical Spark-Ignition Engine. Combustion Science and Technology, v.
189, n. 6, p. 1013–1060, 2017.
[70] AUGOYE, A.; ALEIFERIS, P. Characterization of Flame Development with Hydrous
and Anhydrous Ethanol Fuels in a Spark-Ignition Engine with Direct Injection and Port
Injection Systems. 2014.
[71] ALEIFERIS, P. G.; SERRAS-PEREIRA, J.; RICHARDSON, D. Characterisation of
flame development with ethanol, butanol, iso-octane, gasoline and methane in a direct-
injection spark-ignition engine. Fuel, v. 109, p. 256–278, jul. 2013.
[72] ALEIFERIS, P. G. et al. Cyclic Variations of Initial Flame Kernel Growth in a Honda
VTEC-E Lean-Burn Spark-Ignition Engine. SAE Technical Paper, n. 2000-01–1207,
2000.
[73] RATHINAM, B.; RAVET, F.; DELAHAYE, L. Experimental and Numerical
Investigations of Tumble Motion on an Optical Single Cylinder Engine. SAE
International Journal of Engines, v. 1698, 2015.
[74] MEROLA, S. S. et al. Split Injection in a DISI Engine Fuelled with Butanol and
Gasoline Analyzed through Integrated Methodologies. SAE International Journal of
Engines, v. 8, n. 2, p. 2015-01–0748, 14 abr. 2015.
[75] RAHMAN, K. M. et al. Local fuel concentration measurement through spark-induced
breakdown spectroscopy in a direct-injection hydrogen spark-ignition engine.
International Journal of Hydrogen Energy, v. 41, n. 32, p. 14283–14292, ago. 2016.
[76] IRIMESCU, A.; MEROLA, S. S.; VALENTINO, G. Application of an entrainment
turbulent combustion model with validation based on the distribution of chemical
species in an optical spark ignition engine. Applied Energy, v. 162, p. 908–923, 2016.
[77] KOJIMA, J.; IKEDA, Y.; NAKAJIMA, T. Basic aspects of OH(A), CH(A), and C2(d)
chemiluminescence in the reaction zone of laminar methane–air premixed flames.
Combustion and Flame, v. 140, n. 1–2, p. 34–45, jan. 2005.
[78] SAMANIEGO, J.-M.; EGOLFOPOULOS, F. N.; BOWMAN, C. T. CO 2 *
Chemiluminescence in Premixed Flames. Combustion Science and Technology, v.
109, n. 1–6, p. 183–203, nov. 1995.
[79] STEVENS, E.; STEEPER, R. Piston Wetting in an Optical DISI Engine: Fuel Films,
Pool Fires, and Soot Generation. SAE Technical Paper 2001-01-1203. Anais...5 mar.
2001Disponível em: http://papers.sae.org/2001-01-1203/.
[80] MONTEIRO, E. et al. Laminar burning velocities and Markstein numbers of syngas-air
mixtures. Fuel, v. 89, n. 8, p. 1985–1991, 2010.
[81] VERHELST, S. et al. Laminar and unstable burning velocities and Markstein lengths
of hydrogen–air mixtures at engine-like conditions. Proceedings of the Combustion
Institute, v. 30, n. 1, p. 209–216, jan. 2005.
170

[82] ALEIFERIS, P. G.; ROSATI, M. F. Flame chemiluminescence and OH LIF imaging in


a hydrogen-fuelled spark-ignition engine. International Journal of Hydrogen
Energy, v. 37, n. 2, p. 1797–1812, 2012.
[83] HEYWOOD, J. B.; VILCHIS, F. R. Comparison of Flame Development in a Spark-
Ignition Engine Fueled with Propane and Hydrogen. Combustion Science and
Technology, v. 38, n. 5–6, p. 313–324, 20 jul. 1984.
[84] AYALA, F. A.; HEYWOOD, J. B. Lean SI Engines: The role of combustion variability
in defining lean limits. SAE Technical Paper, n. 2007-24–0030, 2007.
[85] MARTÍNEZ-BOGGIO, S. D. et al. Simulation of cycle-to-cycle variations on spark
ignition engines fueled with gasoline-hydrogen blends. International Journal of
Hydrogen Energy, v. 41, n. 21, p. 9087–9099, 2016.
[86] G. SMITH et al. GRI-Mech 3.0, Gas Research InstituteDesignChicago, 2011.
Disponível em: www.me.berkeley.edu/gri_mech.
[87] IRIMESCU, A. et al. Evaluation of different methods for combined thermodynamic and
optical analysis of combustion in spark ignition engines. Energy Conversion and
Management, v. 87, p. 914–927, 2014.
[88] Nefischer, A. Application of a flow field based heat transfer model to hydrogen internal
combustion engines. SAE Int J Engines, v. 2, p. 1251–1264, 2009.
[89] MICHL, J. et al. Derivation and validation of a heat transfer model in a hydrogen
combustion engine. Applied Thermal Engineering, v. 98, p. 502–512, 2016.
[90] SHUDO, T.; SUZUKI, H. Applicability of heat transfer equations to hydrogen
combustion. JSAE Review, v. 23, n. 3, p. 303–308, 2002.
[91] WOSCHNI, G. A Universally Applicable Equation for the Instantaneous Heat
Transfer Coefficient in the Internal Combustion Engine. SAE Technical Paper
670931. Anais...1 fev. 1967Disponível em: http://papers.sae.org/670931/.
[92] ANNAND, W. J. D. Heat Transfer in the Cylinders of Reciprocating Internal
Combustion Engines. Thermodynamics and Fluid Mechanics Group, v. 1, p. 973–
996, 1963.
[93] Eichelberg, G. Some investigations on old combustion-engine problems. Parts I and II.
Engineering, v. 148, p. 463–466 and p. 547–550, 1939.
[94] IRIMESCU, A. et al. Development of a semi-empirical convective heat transfer
correlation based on thermodynamic and optical measurements in a spark ignition
engine. Applied Energy, v. 157, p. 777–788, 2015b.
[95] HUNG, D. L. S. et al. Experimental Investigation of the Variations of Early Flame
Development in a Spark-Ignition Direct-Injection Optical Engine. Journal of
Engineering for Gas Turbines and Power, v. 136, n. 10, p. 101503, 2014.
[96] PARKER, J. Algorithms for image processing and computer vision. 2th Edsiti ed.
Indianapolis, Indiana: Wiley Publishing, Inc, 2010.
[97] SUN, B.; ZHANG, D.; LIU, F. Cycle variations in a hydrogen internal combustion
engine. International Journal of Hydrogen Energy, v. 38, n. 9, p. 3778–3783, 2013.
[98] IRIMESCU, A. et al. Numerical investigation of engine speed and fuel composition
effects on convective heat transfer in a spark ignition engine fueled with methane-
hydrogen blends. (CHT-17-223, Ed.)Proceedings of International Symposium on
Advances in Computational Heat Transfer. Anais...Napoli, Italy: 2017b.
[99] SHIVAPUJI, A. M.; DASAPPA, S. Quasi dimensional numerical investigation of
syngas fuelled engine operation: MBT operation and parametric sensitivity analysis.
Applied Thermal Engineering, v. 124, p. 911–928, 2017b.
[100] ROCCO, G. et al. Curvature effects in turbulent premixed flames of H2/Air: A DNS
study with reduced chemistry. Flow, Turbulence and Combustion, v. 94, n. 2, p. 359–
379, 2015.
171

[101] SU, J. et al. Combined effects of cooled EGR and a higher geometric compression ratio
on thermal efficiency improvement of a downsized boosted spark-ignition direct-
injection engine. Energy Conversion and Management, v. 78, p. 65-73, 2014.
[102] BORETTI, A. Water injection in directly injected turbocharged spark ignition engines.
Applied Thermal Engineering, v. 52, n. 1, p. 62–68, abr. 2013.
[103] PAN, M. et al. Effects of EGR, compression ratio and boost pressure on cyclic variation
of PFI gasoline engine at WOT operation. Applied Thermal Engineering, v. 64, n. 1–
2, p. 491–498, mar. 2014.
[104] BÖKER, D.; BRÜGGEMANN, D. Advancing lean combustion of hydrogen–air
mixtures by laser-induced spark ignition. International Journal of Hydrogen Energy,
v. 36, n. 22, p. 14759–14767, nov. 2011.
[105] FANELLI, E. et al. On laminar flame speed correlations for H2/CO combustion in
premixed spark ignition engines. Applied Energy, v. 130, p. 166–180, 2014.
[106] YAN, F.; XU, L.; WANG, Y. Application of hydrogen enriched natural gas in spark
ignition IC engines: from fundamental fuel properties to engine performances and
emissions. Renewable and Sustainable Energy Reviews, v. 82, p. 1457–1488, fev.
2018.
[107] MA, F. et al. Performance and emission characteristics of a turbocharged spark-ignition
hydrogen-enriched compressed natural gas engine under wide open throttle operating
conditions. International Journal of Hydrogen Energy, v. 35, n. 22, p. 12502–12509,
nov. 2010.
[108] CEPER, B. A.; AKANSU, S. O.; KAHRAMAN, N. Investigation of cylinder pressure
for H2/CH4mixtures at different loads. International Journal of Hydrogen Energy,
v. 34, n. 11, p. 4855–4861, 2009.
[109] MATHAI, R. et al. Comparative evaluation of performance, emission, lubricant and
deposit characteristics of spark ignition engine fueled with CNG and 18% hydrogen-
CNG. International Journal of Hydrogen Energy, v. 37, n. 8, p. 6893–6900, 2012.
[110] TURRIO-BALDASSARRI, L. et al. Evaluation of emission toxicity of urban bus
engines: Compressed natural gas and comparison with liquid fuels. Science of the Total
Environment, v. 355, n. 1–3, p. 64–77, 2006.
[111] KARIM, G. A. Hydrogen as a spark ignition engine fuel. International Journal of
Hydrogen Energy, v. 28, p. 569–577, 2003.
[112] SINGH, S. et al. Hydrogen: A sustainable fuel for future of the transport sector.
Renewable and Sustainable Energy Reviews, v. 51, p. 623–633, 2015.
[113] VERHELST, S.; WALLNER, T. Hydrogen-fueled internal combustion engines.
Progress in Energy and Combustion Science, v. 35, n. 6, p. 490–527, 2009.
[114] PARK, C. et al. A comparative study of lean burn and exhaust gas recirculation in an
HCNG-fueled heavy-duty engine. International Journal of Hydrogen Energy, v. 42,
n. 41, p. 26094–26101, 2017.
[115] ALEIFERIS, P. G.; BEHRINGER, M. K. Modulation of integral length scales of
turbulence in an optical SI engine by direct injection of gasoline, iso-octane, ethanol
and butanol fuels. Fuel, v. 189, p. 238–259, 2017.
[116] PAPAGIANNAKIS, R. G.; ZANNIS, T. C. Thermodynamic analysis of combustion
and pollutants formation in a wood-gas spark-ignited heavy-duty engine. International
Journal of Hydrogen Energy, v. 38, n. 28, p. 12446–12464, 2013.
[117] RAKOPOULOS, C. D.; MICHOS, C. N.; GIAKOUMIS, E. G. Availability analysis of
a syngas fueled spark ignition engine using a multi-zone combustion model. Energy, v.
33, n. 9, p. 1378–1398, 2008.
[118] MA, F. et al. Development and validation of a quasi-dimensional combustion model for
SI engines fuelled by HCNG with variable hydrogen fractions. International Journal
172

of Hydrogen Energy, v. 33, n. 18, p. 4863–4875, 2008.


[119] AGARWAL, A. et al. Assessment of Single- and Two-Zone Turbulence Formulations
for Quasi-Dimensional Modeling of Spark-Ignition Engine Combustion. Combustion
Science and Technology, v. 136, n. 1, p. 13–39, 1 jul. 1998.
[120] PERINI, F.; PALTRINIERI, F.; MATTARELLI, E. A quasi-dimensional combustion
model for performance and emissions of SI engines running on hydrogen-methane
blends. International Journal of Hydrogen Energy, v. 35, n. 10, p. 4687–4701, 2010.
[121] CINNELLA, P. et al. Multi-Zone Quasi-Dimensional Combustion Models for
Spark-Ignition Engines Multi-zone quasi-dimensional combustion models for
Spark-Ignition engines. SAE Technical Papers 2013-24-0025. Anais...2013.
[122] THOBOIS, L. et al. The Analysis of Natural Gas Engine Combustion Specificities in
Comparison with Isooctane Through CFD Computation. SAE Technical, n. 724, 2003.
[123] BATTISTONI, M. et al. Combustion CFD modeling of a spark ignited optical access
engine fueled with gasoline and ethanol. Energy Procedia, v. 82, p. 424–431, 2015.
[124] GAMIÑO, B.; AGUILLÓN, J. Numerical simulation of syngas combustion with a
multi-spark ignition system in a diesel engine adapted to work at the Otto cycle. Fuel,
v. 89, n. 3, p. 581–591, 2010.
[125] CURTO-RISSO, P. L.; MEDINA, A.; CALVO HERNÁNDEZ, A. Optimizing the
geometrical parameters of a spark ignition engine: Simulation and theoretical tools.
Applied Thermal Engineering, v. 31, n. 5, p. 803–810, 2011.
[126] CURTO-RISSO, P. L.; MEDINA, A.; HERNÁNDEZ, A. C. Theoretical and simulated
models for an irreversible Otto cycle. Journal of Applied Physics, v. 104, n. 9, p. 1–
11, 2008.
[127] SCALA, F.; GALLONI, E.; FONTANA, G. Numerical analysis of a downsized spark-
ignition engine fueled by butanol / gasoline blends at part-load operation. Applied
Thermal Engineering, v. 102, p. 383–390, 2016.
[128] VERHELST, S.; SIERENS, R. A quasi-dimensional model for the power cycle of a
hydrogen-fuelled ICE. International Journal of Hydrogen Energy, v. 32, n. 15 SPEC.
ISS., p. 3545–3554, 2007.
[129] JI, C. et al. A quasi-dimensional model for combustion performance prediction of an SI
hydrogen-enriched methanol engine. International Journal of Hydrogen Energy, v.
41, n. 39, p. 17676–17686, 2016.
[130] RAKOPOULOS, C. D.; MICHOS, C. N. Development and validation of a multi-zone
combustion model for performance and nitric oxide formation in syngas fueled spark
ignition engine. Energy Conversion and Management, v. 49, p. 2924–2938, 2008.
[131] KAN, X. et al. An investigation on utilization of biogas and syngas produced from
biomass waste in premixed spark ignition engine. Applied Energy, v. 212, n.
September 2017, p. 210–222, 2018.
[132] CURTO-RISSO, P. L.; MEDINA, A.; HERNÁNDEZ, A. C. Optimizing the operation
of a spark ignition engine : Simulation and theoretical tools Optimizing the operation of
a spark ignition engine : Simulation. Journal of Applied Physics, v. 105, n. 94904,
2009.
[133] MEDINA, A. et al. Quasi-Dimensional Simulation of Spark Ignition Engines.
London: Springer London, 2014.
[134] P. CURTO, A. MEDINA, A. C. Numerical approach of effects of gasoline-ethanol
blends on cycle-to-cycle variability in spark ignition engines. ECOS 2011
Conference. Anais...Novi Sad, Serbia: 2011.
[135] COSTA, R. B. R. DA et al. Experimental Methodology and Numerical Simulation of
Intake Valves Discharge Coefficients for a Single Cylinder Research Engine. SAE
173

Technical Paper 2015-36-0267, n. 2015-36–0267, 2015.


[136] BARNES-MOSS, H. W. A Designer’s Viewpoint in Passenger Car Engines.
Institution of Mechanical Engineers. Anais...London: 1975.
[137] BLIZARD, N. C.; JAMES C. KECK. Experimental and theoretical investigation of
turbulent burning model for internal combustion engines. SAE paper 740191.
Anais...Detroit, M.I.T.: 1974.
[138] KECK, J. C. Turbulent flame structure and speed in spark-ignition engines.
(Elsevier, Ed.) Symposium (International) on Combustion. Anais...Pittsburgh: Elsevier,
1982.
[139] BERETTA, G. P.; RASHIDI, M.; KECK, J. C. Turbulent flame propagation and
combustion in spark ignition engines. Combustion and Flame, v. 52, p. 217–245, 1983.
[140] ANDREWS, G. E.; BRADLEY, D.; LWAKABAMBA, S. B. Turbulence and Turbulent
Flame Propagation-A Critical Appraisal. Combustion and Flame, v. 24, p. 285–304,
1975.
[141] GÖTTGENS, J.; MAUSS, F.; PETERS, N. Analytic approximations of burning
velocities and flame thicknesses of lean hydrogen, methane, ethylene, ethane, acetylene,
and propane flames. Symposium (International) on Combustion, v. 24, n. 1, p. 129–
135, 1992.
[142] MÜLLER, U. C.; BOLLIG, M.; PETERS, N. Approximations for burning velocities
and Markstein numbers for lean hydrocarbon and methanol flames. Combustion and
Flame, v. 108, n. 3, p. 349–356, 1997.
[143] Lieuwen, T.; Yang, V.; Yetter, R. Synthesis Gas Combustion. Boca Raton: CRC Press,
2009.
[144] Irimescu, A. et al. Correlation between Simulated Volume Fraction Burned Using a
Quasi-Dimensional Model and Flame Area Measured in an Optically Accessible SI
Engine. SAE Technical Paper, 2017-01-0545, 2017, doi:10.4271/2017-01-0545.
[145] SALVI, B. L.; SUBRAMANIAN, K. A. Experimental investigation and
phenomenological model development of flame kernel growth rate in a gasoline fuelled
spark ignition engine. Applied Energy, v. 139, p. 93–103, 2015.
174

9 Appendix

9.1 Appendix A

The focus of this section is the analysis of how different factors of image processing
affect the measurement of flame morphology. Especially the area of the flame, measured by the
number of pixels in the binarized area equal to 1. The flame size can reveal details very early
in the combustion process (0–5% MFB), a period that is not typically resolved well by
thermodynamically derived MFB data. Therefore, is very important an accurate method for the
study of flame propagation. The procedure as was explained in chapter 3, was to binarized the
8 bits images of the flame in each crank angle by the metric method. Then after some image
tools (fill holes, remove small objects), the effective flame area was calculated for each CAD
in several consecutive cycles.

9.1.1 Image processing tests

Figure 95-a shows 55 consecutive cycles of the normalized flame area versus crank
angle after the start of spark. In addition, in black was depicted the mean flame area of the 55
cycles. The difference associate to the flame growth of the cycles was direct correlate with cycle
by cycle variation. This phenomena is always present in SI engines and is due to several factors.
Some correspond to fluid-motion, fuel mixture homogeneity, spark plug energy, temperature
profile among others. The mean flame area with the standard deviation for each CAD is shown
in Figure 95-b. Could be seen that for the first CAD the standard deviation is high (over 30%)
then decrease for the interval between 20-50% of the piston area (approximately 15% of the
measured) and the last part of the measurement cycles is below 5%. This values can change
with the stability of the combustion process but can be set as a reference for a stable flame
development (CovIMEP = 0.6%).
175

a) b)
Figure 95 - Flame Area with dispersion bars for methane λ=1.0, fn=11, exp time=0.5 CAD,
Intensification=65000.

The number of cycles that needs to a representative mean flame area graph was studied.
For the same combustion process was take four different cases in which 5, 15, 25 and 55
consecutive cycles is shown in Figure 96. Could be seen that slight difference was found for 5
cycles than the rest of the cases (5% of the variation in medium and high flame areas), and
negligible variation between 25 and 55 cycles (1% in middle flame growth position 16 CAD
ASOS). Therefore, an intermediate level of combustion cycles was taken (25 cycles) to have a
representative curve of the flame growth for this analysis.
176

Figure 96- Flame Area comparison between different amount of cycles for methane λ=1.0,
fn=11, exp time=0.5 CAD, Intensification=55000.

One important aspect of experimental measurements is the repeatability between


experiments. For this reason, three cases were performed with the same fuel (methane), air-fuel
ratio (λ=1.0), SA (7 CAD BTDC) and camera set up (f-number=11, intensification=55000, and
exposure time=0.5 CAD) to detect difference intrinsic in the system and from the analysis
procedure. The first point is due to the difference in the combustion process like temperature
profile, air-fuel mixture, and cycle by cycle variations. The second could be associated with an
error in the post-processing methodology. For all cases study in this section previous
comparison in in-cylinder pressure was done, to ensure that the combustion process is with the
same fuel and engine set up. Figure 97 depicts the three cases, in which no difference was
detected to case 1 and 3, and the small difference was detected between case 2 and 3. For
intermediate flame areas near 12 CAD ASOS mean variations of 6% were calculated, and for
the last phase of the combustion process, this difference was below 1%.
177

Figure 97 - Comparison Flame Area in the same case for three repetitive combustion process
for methane λ=1.0, f-number=11, intensification=55000 and exposure time 0.5 CAD.

9.1.2 Camera set up tests

Finally, the three camera parameters (exposure time, intensification and f-number) was
studied. This parameters change the light intensity of the photographs and with that the matrix
of 8 bits numbers that represent the flame shape. The change in this three parameters is generally
used to improve the flame quality. However, could affect the threshold selection and the flame
size. For this reason is important to know the effect of each parameter on the results. The
exposure time is the time that the camera is open to taking the light of the flame. For the test
the time between two frames was 1CAD, therefore the maximum exposure time that can be
chosen is 185 μs. For this analysis 1 CAD and 0.5 CAD were selected and graph in Figure 98.
The first phase of the combustion process not difference between the two cases was seen (below
2%). In the last phase this difference increase with a maximum of 5%. Therefore, can be
affirmed that is not a critical parameter because is below the difference found for previous cases
and both exposure time can be used in the tests.
178

Figure 98- Comparison Flame Area with different values of exposure time for methane λ=1.0,
f-number=11, intensification=55000.

The intensification value could be a change in the intensifier that is coupled to the high-
speed camera. The relay lens assembly transfers the image from the output window of the
intensifier to the input of the High-Speed camera. The intensifiers are adjusted for maximum
gain, uniform output brightness and linear amplification of the input light. This parameter
changes directly the gain of the image and with that the intensity of the flame front. The
minimum recommendable value to used is 55000 over a maximum of 80000 by the fabricant
(Figure 99-a). In this study, the variation was from 45000 and 75000, in which the first case is
below the minimum recommendable and the last one is the maximum for not image saturation.
The results are plotted in Figure 99-b, the extremes (45000 and 75000) shows higher values
than intermediate intensifier values with a difference around 10% in the initial phase and 6%
for intermediate and high flame size. For intermediate intensifier values (55000 to 65000) the
difference was below 6% for the entire range. Therefore, intensifier value affects the
measurements of the flame area when the intensity is near the saturation and when is outside of
the linear graph of the fabricant. For this study, the selection of the intensification was selected
between 55000 and 65000 with careful to not saturate the image.
179

a)

Figure 99- Comparison Flame Area with different values of intensification for methane λ=1.0,
f-number=32, exposure time=0.5 CAD.

The f-number (𝒇𝒏 ) of an optical system is the ratio of the system's focal length (f) to
the diameter of the effective aperture (D), Equation 26.
180

𝒇𝒏 = 𝒇/𝑫 (26)

Ignoring differences in light transmission efficiency, a lens with greater f-number


projects darker images. Therefore, is a parameter that can help when fuels or test condition
generate a flame with high luminous intensity. In this study, three different f-numbers were
tested to measure the difference in the flame area. For low and intermediate flame size the f-
number=11 and f-number=32 shows similar flame growth (Figure 100). However, an important
difference was seen between the extreme cases with variation around 17% in the range of 20-
50% normalize flame area. Therefore, this parameter shows larger difference than the previous
cases and in this work was not change when comparing cases with similar light intensity.

Figure 100- Comparison Flame Area with different values of f-number for methane λ=1.0,
exposure time=0.5 CAD, intensification=55000.
181

9.2 Appendix B

9.2.1 Gas flow rates model

The mass flow rate of a gas through a poppet valve is usually described by the equation
for a compressible flow through a flow restriction that is derived from a one-dimensional
isentropic flow analysis. For an ideal gas, flow rate depends on the upstream pressure (𝒑𝟎 ), and
temperature (𝑻𝟎 ), the pressure just downstream the restriction (𝒑𝑻 ) assumed equal to the
pressure at the restriction, and the orifice area, 𝑨𝑻 . Two different regimens of the flow could be
presence, one that is called non-choked flow and Equation 27 is used for the calculus of mass
𝜸
𝒑𝑻 ⁄𝜸−𝟏
flow rate (𝒎̇𝑵𝑪 ). The other regime is chocked flow ( ⁄𝒑𝟎 = (𝟐⁄𝜸 + 𝟏) ) and Equation

28 is used (𝒎̇𝑪 ).

1⁄
1⁄ 𝛾−1 2
2𝐶𝐷 𝐴𝑇 𝑝0 𝑝𝑇 𝛾 2𝛾 𝑝𝑇 𝛾
(27)
𝑚̇𝑁𝐶 = ( ) { [1 − ( ) ]}
√𝑅𝑇0 𝑝0 𝛾−1 𝑝0

𝛾+1
2𝐶𝐷 𝐴𝑇 𝑝0 1 2 2(𝛾−1)
𝑚̇𝐶 = 𝛾 ⁄2 ( ) (28)
√𝑅𝑇0 𝛾+1

The coefficient of discharge (𝐂𝐃 ), that is determined experimentally, accounting for real
gas flow effects. In this study 𝐂𝐃 was taken from an experimental and numerical work [135]
that test the same cylinder head geometry than the present work (Figure 101).
182

Figure 101 - Numerical and experimental data of 𝑪𝑫 for the calculus of mass flow rate
through the valves [135].

The valves lift for intake and exhaust valves was taken from engine specification with
the geometry is shown in Figure 102.

Figure 102 – Valve lift for intake and exhaust valve.

9.2.2 Heat Transfer model

In order to solve the thermodynamic set of differential equations, it is necessary to


assume a heat transfer model between the gas mixture inside the cylinder and the wall. Several
models have been proposed in the literature for spark ignition four strokes engines [91-93]. The
formulation developed by Woschni [91] was used in this work (Equation 29), also widely used
in the literature, because its simplicity to calculate the instantaneous heat flow rate (𝑸̇𝒔 ).
183

𝑄̇𝑠 = ℎ𝐴𝑠 (𝑇 − 𝑇𝑤 ) (29)

where 𝑨𝒔 is the instantaneous surface transfer area, 𝑻 is the instantaneous bulk gas temperature,
and 𝑻𝒘 the mean cylinder inner surface temperature. Woschni’s equation assumes that the
working fluid is air and the influence of Prandtl number is reduced, given that its corresponding
term is close to unity, thus reducing the above relation as Equation 30.

ℎ = 129.8 𝐵 −0.2 𝑝0.8 𝑇 −0.53 𝑤 0.8 (30)

where 𝒑 the instantaneous pressure inside the cylinder, and 𝒘 equivalent velocity
(Equation 31). This last parameter is obtained by matching calculated values with
measurements, with the main assumption that fluid flow inside the cylinder is proportional to
the mean piston speed 𝑺𝒑 .

𝑉𝑑 𝑇𝑟
𝑤 = 𝐶1 𝑆𝑝 + 𝐶2 (𝑝 − 𝑝𝑚 )
𝑃𝑟 𝑉𝑟 (31)

where 𝒑𝒎 is the motored pressure and 𝑷𝒓 , 𝑽𝒓 , 𝑻𝒓 are reference values. Constant 𝑪𝟏 equal to 2.28
during the closed valves period of the cycle, while constant C2 is equal to 3.24 x 10-3 for
compression, combustion and expansion.

9.2.3 Friction model

For friction components, the empirical correlation by Barnes-Moss [136] is considered,


including both linear and quadratic terms of the engine speed (Equation 32).

𝐹𝑓𝑟𝑖𝑐 = 𝐴𝑝𝑖𝑠𝑡𝑜𝑛 (𝐶1 + 𝐶2 𝜃 + 𝐶3 𝜃 2 ) (32)

with 𝑪𝟏 = 𝟏𝟖𝟎𝟎𝟗𝟒, 𝑪𝟐 = 𝟓𝟓𝟏, and 𝑪𝟑 = 𝟏. 𝟒𝟔 obtained by experimental tests in motored


mode with calibration by measurements of torque in different crankshaft rotational velocity.

You might also like