24.07.2013 Views

Comparative analysis of causal diagnosis methods of ... - circe

Comparative analysis of causal diagnosis methods of ... - circe

Comparative analysis of causal diagnosis methods of ... - circe

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

PhD Thesis<br />

<strong>Comparative</strong> <strong>analysis</strong> <strong>of</strong> <strong>causal</strong><br />

<strong>diagnosis</strong> <strong>methods</strong> <strong>of</strong> malfunctions<br />

in power cycles<br />

Sergio Usón Gil<br />

Directed by: Antonio Valero Capilla, Luis Carlos Correas Usón<br />

Department <strong>of</strong> Mechanical Engineering.<br />

University <strong>of</strong> Zaragoza.<br />

January, 2008


Antonio Valero Capilla, PhD, pr<strong>of</strong>essor <strong>of</strong> the Department <strong>of</strong> Mechanical<br />

Engineering <strong>of</strong> the University <strong>of</strong> Zaragoza and Luis Carlos Correas Usón, PhD,<br />

acknowledge that the PhD thesis ‘<strong>Comparative</strong> <strong>analysis</strong> <strong>of</strong> <strong>causal</strong> <strong>diagnosis</strong> <strong>methods</strong> <strong>of</strong><br />

malfunctions in power cycles’ written by Sergio Usón Gil, has been developed under<br />

their supervision.<br />

Zaragoza, January 2008<br />

Antonio Valero Capilla Luis Carlos Correas Usón


I would like to thank a lot <strong>of</strong> people and institutions for their help in the development<br />

<strong>of</strong> this thesis.<br />

My supervisors, Antonio Valero and Luis Correas, for their valuable comments,<br />

their words to encourage me to start, develop and finish this thesis, and their interest in<br />

my pr<strong>of</strong>essional development. Their complementary vision <strong>of</strong> the <strong>diagnosis</strong> problem has<br />

become essential in the development <strong>of</strong> this work.<br />

My workmates <strong>of</strong> CIRCE, for their help in a lot <strong>of</strong> tasks and, above all, for their<br />

friendship. The list <strong>of</strong> names would be endless; so, please feel all <strong>of</strong> you included.<br />

Part <strong>of</strong> this work was developed at Politecnico di Torino, with the group <strong>of</strong> Pr<strong>of</strong>.<br />

Vittorio Verda. All people I met in those three months made me feel like at home.<br />

The Regional Government <strong>of</strong> Aragón and European Social Found for the grant<br />

B120/2003, which provided me the financial support to develop this work; and the<br />

personnel <strong>of</strong> Results Office at Teruel Power plant (Endesa Generación) for the<br />

information and valuable advises about the ‘practical side’ <strong>of</strong> the <strong>diagnosis</strong> problem.<br />

My first pupils and the personnel <strong>of</strong> the Department <strong>of</strong> Mechanical Engineering, for<br />

these nice first steps with the chalk.<br />

Last but not least, my friends and, above all, my family.<br />

Thanks very much to all <strong>of</strong> you.


To my parents, Tomás and<br />

Conchita, and my sister, Natalia.


Table <strong>of</strong> contents<br />

Table <strong>of</strong> contents............................................................................................................ ix<br />

List <strong>of</strong> tables ................................................................................................................. xiii<br />

List <strong>of</strong> figures ................................................................................................................ xv<br />

Nomenclature............................................................................................................. xxiii<br />

1. INTRODUCTION ...................................................................................................... 1<br />

1.1Justification, objectives and content <strong>of</strong> the thesis ............................................... 2<br />

2. STATE OF THE ART OF THERMOECONOMIC DIAGNOSIS. ...................... 5<br />

2.1 Thermoeconomic approach to <strong>diagnosis</strong>. ............................................................ 5<br />

2.1.1 Thermoeconomic model.................................................................................... 5<br />

2.1.2 Exergetic cost. ................................................................................................... 7<br />

2.1.3 The fuel impact formula .................................................................................... 8<br />

2.1.4. Cost formation process <strong>of</strong> wastes (or residues).............................................. 10<br />

2.1.5. Other issues. ................................................................................................... 11<br />

2.2 Thermoeconomic <strong>diagnosis</strong> methodologies. ...................................................... 13<br />

2.2.1 The filtration <strong>of</strong> the effects induced by the control system. ............................ 14<br />

2.2.2 Other <strong>methods</strong> based on the filtration <strong>of</strong> induced effects................................ 17<br />

2.2.3 Representation <strong>of</strong> the malfunctions in the h-s plane........................................ 19<br />

2.2.4 Methods directly based on the thermodynamic representation <strong>of</strong> the system 22<br />

2.2.5. Concept <strong>of</strong> the <strong>diagnosis</strong> algorithm and mathemathical formulation............. 23<br />

2.2.6. Other examples............................................................................................... 25<br />

2.3. State <strong>of</strong> the art <strong>of</strong> systems for <strong>diagnosis</strong> and other related issues. ................. 26<br />

2.3.1. Diagnosis systems in general.......................................................................... 27<br />

2.3.2. Diagnosis centres and manufacturer-operator collaboration.......................... 28<br />

2.3.3. Monitoring, <strong>diagnosis</strong> and optimization <strong>of</strong> boilers......................................... 29<br />

2.3.4. Other issues. ................................................................................................... 30<br />

2.4. Conclusion ........................................................................................................... 31<br />

3. QUANTITATIVE CAUSALITY ANALYSIS ....................................................... 33<br />

3.1. Re-formulation <strong>of</strong> the <strong>diagnosis</strong> problem......................................................... 33<br />

3.1.1. Example.......................................................................................................... 35<br />

3.2. Influence <strong>of</strong> measurement errors. ..................................................................... 40<br />

3.2.1 Example........................................................................................................... 42<br />

3.3. Analysis <strong>of</strong> non-linearities.................................................................................. 45<br />

3.3.1. Evaluation <strong>of</strong> the influence <strong>of</strong> linearization errors in the <strong>diagnosis</strong> result ..... 45<br />

3.3.2. Strategies to reduce the errors <strong>of</strong> the methodology due to non-linearities ..... 47<br />

3.4. Free <strong>diagnosis</strong> variables ..................................................................................... 51<br />

3.4.1 Mathematical conditions required to the set <strong>of</strong> free <strong>diagnosis</strong> variables......... 51<br />

3.4.2. Influence <strong>of</strong> a modification in the free <strong>diagnosis</strong> variable choice on the<br />

<strong>diagnosis</strong> result. .............................................................................................................. 52<br />

3.4.3. Choice <strong>of</strong> the free <strong>diagnosis</strong> variables............................................................ 53<br />

3.5. Causality chains .................................................................................................. 55<br />

ix


3.5.1. General formulation <strong>of</strong> <strong>causal</strong>ity chains......................................................... 56<br />

3.5.2. Causality chains based on the dependence among free <strong>diagnosis</strong> variables. . 58<br />

3.5.3. Causality chains based on the dependence <strong>of</strong> free <strong>diagnosis</strong> variables with<br />

other variables. ............................................................................................................... 59<br />

3.6. Calculation <strong>of</strong> fuel impact on a time span. ....................................................... 60<br />

3.7. Conclusion ........................................................................................................... 61<br />

4. QUANTITATIVE CAUSALITY ANALYSIS AND OTHER DIAGNOSIS<br />

METHODS.................................................................................................................... 63<br />

4.1. Quantitative <strong>causal</strong>ity <strong>analysis</strong> and the fuel impact formula ......................... 63<br />

4.1.1. From free <strong>diagnosis</strong> variables to unit exergy cost and final product.............. 64<br />

4.1.2. Quantification <strong>of</strong> intrinsic and induced malfunctions .................................... 66<br />

4.1.3. Determination <strong>of</strong> the suitability <strong>of</strong> a productive structure.............................. 73<br />

4.1.4. Conclusion...................................................................................................... 75<br />

4.2. Application <strong>of</strong> neural networks and linear regression to the <strong>diagnosis</strong> <strong>of</strong><br />

energy systems. ............................................................................................................. 75<br />

4.2.1. Linear regression fundamentals...................................................................... 76<br />

4.2.2. Neural network fundamentals......................................................................... 77<br />

4.2.2.1. Neurons, layers and networks. .............................................................................. 78<br />

4.2.2.2. Feedforward networks and backpropagation......................................................... 80<br />

4.2.3. Application for <strong>diagnosis</strong>................................................................................ 82<br />

4.2.3.1. Diagnosis <strong>of</strong> the complete system......................................................................... 82<br />

4.2.3.2. Diagnosis <strong>of</strong> sub-systems. Hybrid configurations................................................. 83<br />

4.2.3.3. Other issues. .......................................................................................................... 84<br />

4.3. Conclusión ........................................................................................................... 84<br />

5 CASE STUDY: 3X 350 MW COAL-FIRED POWER PLANT............................ 85<br />

5.1 Teruel power plant description. ......................................................................... 85<br />

5.1.1 Steam cycle description................................................................................... 86<br />

5.1.2 Cooling system ................................................................................................ 89<br />

5.1.3 Boiler ............................................................................................................... 90<br />

5.2 Diagnosis model for a pulverized-coal power plant. ........................................ 95<br />

5.2.1 Steam cycle...................................................................................................... 96<br />

5.2.1.1 Steam turbine.......................................................................................................... 96<br />

5.2.1.2 Steam turbine seals and other secondary flows...................................................... 99<br />

5.2.1.3 Alternator ............................................................................................................. 100<br />

5.2.1.4 Feedwater heaters................................................................................................. 100<br />

5.2.1.5 Pumps................................................................................................................... 102<br />

5.2.1.6 Pressure losses...................................................................................................... 103<br />

5.2.2 Cooling system .............................................................................................. 104<br />

5.2.2.1 Condenser............................................................................................................. 105<br />

5.2.2.2 Cooling tower....................................................................................................... 107<br />

5.2.2.3 Pumps and ducts................................................................................................... 108<br />

5.2.3 Boiler ............................................................................................................. 108<br />

5.2.3.1 Boiler efficiency and coal consumption............................................................... 109<br />

5.2.3.2 Combustion .......................................................................................................... 114<br />

5.2.3.3 Heat transfer in boilers ......................................................................................... 115<br />

5.2.3.4 Air preheaters ....................................................................................................... 120<br />

5.2.3.5 Ancillary devices.................................................................................................. 122<br />

5.3 Monitoring and <strong>diagnosis</strong> model for Teruel power plant.............................. 123<br />

5.3.1 Measurement review. .................................................................................... 123<br />

5.3.2 Free <strong>diagnosis</strong> variables ................................................................................ 127<br />

5.3.3 Global efficiency indicators. ......................................................................... 131<br />

5.4 Conclusion .......................................................................................................... 132<br />

x


6 APPLICATION OF QUANTITATIVE CAUSALITY ANALYSIS................... 133<br />

6.1 Introduction ....................................................................................................... 133<br />

6.2 Evolution <strong>of</strong> efficiency indicators..................................................................... 136<br />

6.3 Diagnosis results <strong>analysis</strong>.................................................................................. 139<br />

6.3.1 Ambient conditions. ...................................................................................... 139<br />

6.3.2 Fuel quality.................................................................................................... 146<br />

6.3.3 Steam cycle set points. .................................................................................. 158<br />

6.3.4 Boiler set points............................................................................................. 164<br />

6.3.5 Steam cycle component parameters. ............................................................. 175<br />

6.3.6 Cooling system component parameters......................................................... 192<br />

6.3.7 Boiler component parameters........................................................................ 196<br />

6.3.8 Summary <strong>of</strong> results........................................................................................ 203<br />

6.4 Residual term <strong>analysis</strong>. ..................................................................................... 209<br />

6.4.1 Residual term distribution. ............................................................................ 209<br />

6.4.2 Non-dimensional residual term distribution.................................................. 212<br />

6.4.3 Conclusion..................................................................................................... 214<br />

6.5 Impacts aggregation .......................................................................................... 214<br />

6.5.1 Cycle heat rate ............................................................................................... 215<br />

6.5.2 Boiler efficiency ............................................................................................ 218<br />

6.5.3 Unit heat rate ................................................................................................. 223<br />

6.5.4 Conclusion..................................................................................................... 225<br />

6.6 Causality chains ................................................................................................. 226<br />

6.6.1 Determination <strong>of</strong> a correlation for cooling tower effectiveness.................... 227<br />

6.6.2 Decomposition <strong>of</strong> the impact <strong>of</strong> cooling tower effectiveness. ...................... 228<br />

6.6.3 Aggregated impacts ....................................................................................... 230<br />

6.7 Conclusion .......................................................................................................... 232<br />

6.7.1 Conclusions on the method ........................................................................... 232<br />

6.7.2 Practical results about the working example............................................ 233<br />

6.7.2.1 Steam cycle .......................................................................................................... 233<br />

6.7.2.2 Boiler.................................................................................................................... 234<br />

7 APPLICATION OF QUANTITATIVE CAUSALITY ANALYSIS TO THE<br />

QUANTIFICATION OF INTRINSIC AND INDUCED MALFUNCTIONS. ..... 235<br />

7.1 Description <strong>of</strong> the productive structure. ......................................................... 235<br />

7.1.1 Simplified physical structure......................................................................... 235<br />

7.1.2 Productive structure....................................................................................... 239<br />

7.2 Example <strong>of</strong> thermoeconomic <strong>diagnosis</strong> ........................................................... 243<br />

7.2.1 Preliminary results......................................................................................... 243<br />

7.2.2 Decomposition <strong>of</strong> unit exergy consumptions and plant products.................. 248<br />

7.2.3 Decomposition <strong>of</strong> malfunctions..................................................................... 253<br />

7.2.4 Malfunction cost decomposition. MFI and MFD tables. .............................. 255<br />

7.3 General quantification <strong>of</strong> intrinsic and induced effects................................. 263<br />

7.3.1 Decomposition <strong>of</strong> unit exergy consumptions and plant products.................. 263<br />

7.3.2 Decomposition <strong>of</strong> malfunctions..................................................................... 266<br />

7.3.3 Decomposition <strong>of</strong> malfunction costs. ............................................................ 268<br />

7.4 Conclusion .......................................................................................................... 272<br />

8 APPLICATION OF LINEAR REGRESSION AND NEURAL NETWORKS. 273<br />

8.1 Linear regression ............................................................................................... 274<br />

8.1.1 Variable change ............................................................................................. 274<br />

8.1.2 Results ........................................................................................................... 276<br />

8.2 Neural networks................................................................................................. 284<br />

xi


8.2.1 Neural network definition and training ......................................................... 285<br />

8.2.2 Results ........................................................................................................... 289<br />

8.3 Conclusion .......................................................................................................... 296<br />

9 CONCLUSION ........................................................................................................ 299<br />

9.1 Synthesis ............................................................................................................. 299<br />

9.2 Contributions ..................................................................................................... 301<br />

9.3 Perspectives ........................................................................................................ 304<br />

REFERENCES ........................................................................................................... 307<br />

xii


List <strong>of</strong> tables<br />

Table 3.1. Properties <strong>of</strong> streams in the cycle.................................................................. 36<br />

Table 4.1: Table <strong>of</strong> malfunctions and free <strong>diagnosis</strong> variables (MFD).......................... 69<br />

Table 4.2: Table <strong>of</strong> intrinsic and induced malfunctions (MFI). ..................................... 73<br />

Table 5.1. Main characteristics <strong>of</strong> the Teruel Power plant steam cycle. (Zaleta, 1997). 88<br />

Table 5.2. Condenser characteristics (Lazaro, 2001) ..................................................... 89<br />

Table 5.3. Cooling water design characteristics (Lazaro, 2001). ................................... 90<br />

Table 5.4. Teruel power plant boilers, nominal production (Díez, 2002). ..................... 90<br />

Table 5.5. Exchange area <strong>of</strong> the boiler heat exchangers. (Díez, 2002) .......................... 94<br />

Table 5.6.A. List <strong>of</strong> plant measurements used. ............................................................ 124<br />

Table 5.6.B. List <strong>of</strong> plant measurements used.............................................................. 125<br />

Table 5.6.C. List <strong>of</strong> plant measurements used.............................................................. 126<br />

Table 5.7.A. Free <strong>diagnosis</strong> variables. ......................................................................... 128<br />

Table 5.7.B. Free <strong>diagnosis</strong> variables........................................................................... 129<br />

Table 5.7.C. Free <strong>diagnosis</strong> variables.......................................................................... 130<br />

Table 5.8. Free <strong>diagnosis</strong> variables with constant value. ............................................. 131<br />

Table 6.1.A. Summary <strong>of</strong> <strong>diagnosis</strong> results.................................................................. 204<br />

Table 6.1.B. Summary <strong>of</strong> <strong>diagnosis</strong> results.................................................................. 205<br />

Table 6.1.C. Summary <strong>of</strong> <strong>diagnosis</strong> results.................................................................. 206<br />

Table 6.1.D. Summary <strong>of</strong> <strong>diagnosis</strong> results.................................................................. 207<br />

Table 6.1.E. Summary <strong>of</strong> <strong>diagnosis</strong> results. ................................................................. 208<br />

Table 7.1. A. Description <strong>of</strong> flows <strong>of</strong> the simplified physical structure. ..................... 236<br />

Table 7.1. B. Description <strong>of</strong> flows <strong>of</strong> the simplified physical structure....................... 237<br />

Table 7.2. Description <strong>of</strong> components <strong>of</strong> the physical structure. ................................. 239<br />

Table 7.3. Components <strong>of</strong> the productive structure ..................................................... 240<br />

Table 7.4. FPR Table.................................................................................................... 241<br />

Table 7.5.A. Free <strong>diagnosis</strong> variables in the actual and reference states. .................... 244<br />

Table 7.5.B. Free <strong>diagnosis</strong> variables in the actual and reference state. ...................... 245<br />

Table 7.6. Exergy and exergy costs <strong>of</strong> the productive flows (kW). ............................. 246<br />

Table 7.7. Exergy and exergy cost <strong>of</strong> waste flows (kW).............................................. 247<br />

Table 7.8. Unit exergy costs. ........................................................................................ 247<br />

Table 7.9. Unit exergy consumptions........................................................................... 248<br />

Table 7.10. Unit exergy consumptions associated with the residues. .......................... 248<br />

Table 7.11. Table <strong>of</strong> intrinsic and induced malfunctions (MFI). ................................. 258<br />

Table 7.12.A. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD). 259<br />

Table 7.12.B. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD). 260<br />

Table 7.12.C. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD). 261<br />

Table 7.12.D. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD) 262<br />

Table 8.1. Pairs <strong>of</strong> variables highly correlated. ............................................................ 274<br />

Table 8.2.A. Impact factors obtained by linear regression........................................... 277<br />

Table 8.2.B. Impact factors obtained by linear regression. .......................................... 278<br />

xiii


Table 8.2.C. Impact factors obtained by linear regression. .......................................... 279<br />

Table 8.2.D. Impact factors obtained by linear regression........................................... 280<br />

Table 8.2.E. Impact factors obtained by linear regression. .......................................... 281<br />

xiv


List <strong>of</strong> figures<br />

Figure 2.1. Example <strong>of</strong> a productive structure. ................................................................ 6<br />

Figure 2.2. RPS representation in the ω, σ, MFR space................................................. 20<br />

Figure 2.3: Dissipation temperature. .............................................................................. 21<br />

Figure 3.1. Scheme <strong>of</strong> the example cycle....................................................................... 35<br />

Figure 4.1. Single neuron. Source: Demuth and Beale (2002)....................................... 78<br />

Figure 4.2. Hard-lim transfer function. Source: Demuth and Beale (2002)................... 78<br />

Figure 4.3. Linear transfer function. Source: Demuth and Beale (2002) ....................... 79<br />

Figure 4.4. Log-sigmoid transfer function. Source: Demuth and Beale (2002)............. 79<br />

Figure 4.5. Tan-sigmoid transfer function. Source: Demuth and Beale (2002) ............. 79<br />

Figure 4.6. Layer <strong>of</strong> a neural network. Source: Demuth and Beale (2002).................... 80<br />

Figure 5.1. Steam cycle. ................................................................................................. 87<br />

Figure 5.2. Fans and preheaters...................................................................................... 91<br />

Figure 5.3. Schematic side view <strong>of</strong> the boiler. ............................................................... 93<br />

Figure 5.4. Temperature pr<strong>of</strong>ile in a feedwater heater. ................................................ 101<br />

Figure 6.1. Unit heat rate evolution.............................................................................. 137<br />

Figure 6.2. Steam cycle heat rate evolution. ................................................................ 138<br />

Figure 6.3. Boiler efficiency evolution......................................................................... 139<br />

Figure 6.4. Ambient temperature evolution.................................................................. 140<br />

Figure 6.5. Cycle heat rate impact due to ambient temperature................................... 141<br />

Figure 6.6. Impact <strong>of</strong> ambient temperature on boiler efficiency.................................. 142<br />

Figure 6.7. Unit heat rate impact due to ambient temperature. .................................... 142<br />

Figure 6.8. Ambient temperature variation and its impact on cycle heat rate.............. 143<br />

Figure 6.9. Cycle heat rate impact due to relative humidity......................................... 144<br />

Figure 6.10. Relative humidity variation and its impact on cycle heat rate. ................ 144<br />

Figure 6.11. Relative humidity variation and its impact on boiler efficiency.............. 145<br />

Figure 6.12. Wind speed evolution............................................................................... 146<br />

Figure 6.13. Coal high heating value evolution. .......................................................... 147<br />

Figure 6.14. Impact on boiler efficiency due to coal HHV. ......................................... 147<br />

Figure 6.15. Coal HHV variation and its impact on boiler efficiency.......................... 148<br />

Figure 6.16. Impact on boiler efficiency due to carbon in coal.................................... 149<br />

Figure 6.17. Carbon in coal variation and its impact on boiler efficiency. .................. 150<br />

Figure 6.18. Impact on boiler efficiency due to hydrogen in coal................................ 150<br />

Figure 6.19. Hydrogen in coal and its impact on boiler efficiency. ............................. 151<br />

Figure 6.20. Impact on boiler efficiency due to coal moisture..................................... 152<br />

Figure 6.21. Coal moisture variation and its impact on boiler efficiency. ................... 153<br />

Figure 6.22. Impact on boiler efficiency due to ash in coal. ........................................ 153<br />

Figure 6.23. Ash in coal variation and its impact on boiler efficiency......................... 154<br />

Figure 6.24. Impact on boiler efficiency due to sulphur in coal................................... 155<br />

Figure 6.25. Sulphur in coal variation and its impact on boiler efficiency. ................. 156<br />

Figure 6.26. Natural gas fraction evolution.................................................................. 156<br />

xv


Figure 6.27. Natural gas fraction variation and its impact on boiler efficiency........... 157<br />

Figure 6.28. Live steam temperature evolution............................................................ 158<br />

Figure 6.29. Live steam temperature variation and its impact on cycle heat rate. ....... 159<br />

Figure 6.30. Reheated steam temperature evolution. ................................................... 160<br />

Figure 6.31. Reheated steam temperature variation and its impact on cycle heat rate. 160<br />

Figure 6.32. Live steam pressure evolution.................................................................. 161<br />

Figure 6.33. Live steam pressure and its impact on cycle heat rate. ............................ 161<br />

Figure 6.34. Gross electric power evolution................................................................. 162<br />

Figure 6.35. Gross electric power variation and its impact on cycle heat rate............. 163<br />

Figure 6.36. Gross electric power variation and its impact on boiler efficiency.......... 163<br />

Figure 6.37. Oxygen set-point evolution...................................................................... 164<br />

Figure 6.38. Oxygen set point variation and its impact on boiler efficiency. .............. 165<br />

Figure 6.39. Evolution <strong>of</strong> average cold-side temperature in secondary air preheaters. 166<br />

Figure 6.40. Average cold-side temperature in secondary air preheaters variation and its<br />

impact on unit heat rate. ............................................................................................... 167<br />

Figure 6.41. Average cold-side temperature in primary air preheaters evolution........ 167<br />

Figure 6.42. Average cold-side temperature in primary air preheaters variation and its<br />

impact on unit heat rate. ............................................................................................... 168<br />

Figure 6.43. Sootblowing steam evolution................................................................... 169<br />

Figure 6.44. Sootblowing steam variation and its impact on unit heat rate. ................ 170<br />

Figure 6.45. Air for flue gas desulfuration unit evolution............................................ 171<br />

Figure 6.46. Air for flue gas desulfuration unit variation and its impact on boiler<br />

efficiency. ..................................................................................................................... 171<br />

Figure 6.47. Relation <strong>of</strong> tempering and primary air evolution..................................... 172<br />

Figure 6.48. Variation <strong>of</strong> the relation <strong>of</strong> tempering and primary air and its impact on<br />

boiler efficiency............................................................................................................ 173<br />

Figure 6.49. Primary air-coal ratio and its impact on boiler efficiency........................ 173<br />

Figure 6.50 Impact on boiler efficiency due to primary air-coal ratio. ........................ 174<br />

Figure 6.51. Evolution <strong>of</strong> temperature difference <strong>of</strong> flue gases entering air preheaters.<br />

...................................................................................................................................... 175<br />

Figure 6.52. Evolution <strong>of</strong> high pressure steam turbine isoentropic efficiency............. 176<br />

Figure 6.53. High pressure steam turbine isoentropic efficiency variation and its impact<br />

on cycle heat rate. ......................................................................................................... 177<br />

Figure 6.54. Medium pressure 1 steam turbine isoentropic efficiency evolution. ....... 177<br />

Figure 6.55. Medium pressure 1 steam turbine isoentropic efficiency variation and its<br />

impact on cycle heat rate. ............................................................................................. 178<br />

Figure 6.56. Medium pressure 2 steam turbine isoentropic efficiency evolution. ....... 179<br />

Figure 6.57. Medium pressure 2 steam turbine isoentropic efficiency variation and its<br />

impact on unit heat rate. ............................................................................................... 179<br />

Figure 6.58. Impact on cycle heat rate due to low pressure steam turbine isoentropic<br />

efficiency. ..................................................................................................................... 180<br />

Figure 6.59. Low pressure steam turbine isoentropic efficiency variation and its impact<br />

on unit heat rate. ........................................................................................................... 181<br />

Figure 6.60. Impact on unit heat rate due to flow coefficient <strong>of</strong> medium pressure steam<br />

turbine 1........................................................................................................................ 182<br />

Figure 6.61. Flow coefficient <strong>of</strong> medium pressure steam turbine 1 variation and its<br />

impact on cycle heat rate. ............................................................................................. 182<br />

Figure 6.62. Hot side temperature difference (TTD) <strong>of</strong> 6th preheater evolution. ........ 183<br />

Figure 6.63. TTD <strong>of</strong> 6th preheater variation and its impact on cycle heat rate............ 184<br />

Figure 6.64. Hot side temperature difference <strong>of</strong> 5th water preheater evolution........... 185<br />

xvi


Figure 6.65. Hot side temperature difference (TTD) <strong>of</strong> 5th water preheater variation and<br />

its impact on cycle heat rate. ........................................................................................ 186<br />

Figure 6.66. Hot side temperature difference (TTD) <strong>of</strong> 3rd water preheater evolution.<br />

...................................................................................................................................... 186<br />

Figure 6.67 Pressure drop in reheater evolution........................................................... 188<br />

Figure 6.68. Pressure drop in preheater variation and its impact on cycle heat rate. ... 189<br />

Figure 6.69. Pressure increment in turbo-pump evolution. .......................................... 189<br />

Figure 6.70. Pressure increment in turbo-pump variation and its impact on cycle heat<br />

rate. ............................................................................................................................... 190<br />

Figure 6.71. Water losses evolution. ............................................................................ 191<br />

Figure 6.72. Water losses variation and its impact on unit heat rate............................ 191<br />

Figure 6.73. Condenser effectiveness evolution........................................................... 192<br />

Figure 6.74. Condenser effectiveness variation and its impact on cycle heat rate....... 193<br />

Figure 6.75. Cooling water flow rate evolution............................................................ 193<br />

Figure 6.76. Cooling water flow rate variation and its impact on cycle heat rate........ 194<br />

Figure 6.77. Cooling tower effectiveness evolution..................................................... 195<br />

Figure 6.78. Cooling tower effectiveness variation and its impact on cycle heat rate. 195<br />

Figure 6.79. Secondary air preheater effectiveness evolution...................................... 196<br />

Figure 6.80. Secondary air preheater effectiveness variation and its impact on boiler<br />

efficiency. ..................................................................................................................... 197<br />

Figure 6.81. Primary air preheater effectiveness evolution.......................................... 198<br />

Figure 6.82. Primary air preheater effectiveness variation and its impact on boiler<br />

efficiency. ..................................................................................................................... 198<br />

Figure 6.83. Aggregated boiler effectiveness evolution............................................... 199<br />

Figure 6.84. Aggregated boiler effectiveness variation and its impact on boiler<br />

efficiency. ..................................................................................................................... 200<br />

Figure 6.85. Air infiltration in preheaters evolution..................................................... 200<br />

Figure 6.86. Air infiltration in preheaters and its impact on boiler efficiency............. 201<br />

Figure 6.87. Carbon in ashes evolution. ....................................................................... 202<br />

Figure 6.88. Carbon in ashes variation and its impact on boiler efficiency. ................ 202<br />

Figure 6.89. Distribution <strong>of</strong> residuals <strong>of</strong> cycle heat rate. ............................................. 210<br />

Figure 6.90. Distribution <strong>of</strong> residuals <strong>of</strong> boiler efficiency. .......................................... 211<br />

Figure 6.91. Distribution <strong>of</strong> residuals <strong>of</strong> unit heat rate................................................. 211<br />

Figure 6.92. Distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> cycle heat rate................... 212<br />

Figure 6.93. Distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> boiler efficiency ............... 213<br />

Figure 6.94. Distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> unit heat rate..................... 213<br />

Figure 6.95. Impact on cycle heat rate due to ambient conditions. .............................. 216<br />

Figure 6.96. Impact on cycle heat rate due to cycle set-points..................................... 216<br />

Figure 6.97. Impact on cycle heat rate due to boiler. ................................................... 217<br />

Figure 5.98. Impact on cycle heat rate due to cooling system...................................... 217<br />

Figure 6.99. Impact on cycle heat rate due to cycle components................................. 218<br />

Figure 6.100. Impact on boiler efficiency due to ambient conditions.......................... 219<br />

Figure 6.101. Impact on boiler efficiency caused by fuel variation............................. 219<br />

Figure 6.102. Impact on boiler efficiency caused by cycle and cooling system. ......... 220<br />

Figure 6.103. Impact on boiler efficiency caused by components and set-points <strong>of</strong> the<br />

boiler............................................................................................................................. 221<br />

Figure 6.104. Impact on boiler efficiency caused by boiler set-points......................... 222<br />

Figure 6.105. Impact on boiler efficiency caused by boiler components..................... 222<br />

Figure 6.106. Impact on unit heat rate caused by ambient conditions. ........................ 223<br />

Figure 6.106. Impact on unit heat rate caused by fuel.................................................. 224<br />

xvii


Figure 6.107. Impact on unit heat rate caused by set-points. ....................................... 224<br />

Figure 6.108. Impact on unit heat rate caused by components..................................... 225<br />

Figure 6.109. Dependence <strong>of</strong> cooling tower effectiveness with ambient temperature. 228<br />

Figure 6.110. Impact <strong>of</strong> cooling tower effectiveness on cycle heat rate, after filtering<br />

effects induced by ambient temperature....................................................................... 229<br />

Figure 6.111. Impact on cycle heat rate due to cooling tower effectiveness variation<br />

induced by ambient temperature................................................................................... 229<br />

Figure 6.112. Total impact on cycle heat rate due to ambient temperature, including<br />

induced effect on cooling tower effectiveness. ............................................................ 230<br />

Figure 6.113. Impact caused by cooling system on cycle heat rate after filtering effects<br />

induced by ambient temperature on cooling tower effectiveness................................. 231<br />

Figure 6.114. Impact caused by plant components on unit heat rate, after filtering effects<br />

induced by ambient temperature on cooling tower effectiveness................................. 231<br />

Figure 7.1. Simplified physical diagram <strong>of</strong> the cycle................................................... 238<br />

Figure 7.2. Simplified physical diagram <strong>of</strong> the boiler.................................................. 238<br />

Figure 7.3. Productive structure. .................................................................................. 242<br />

Figure 7.4. Decomposition <strong>of</strong> unit exergy consumptions............................................. 250<br />

Figure 7.5. Normalized decomposition <strong>of</strong> unit exergy consumptions.......................... 250<br />

Figure 7.6. Decomposition <strong>of</strong> the unit exergy consumptions <strong>of</strong> boiler. ....................... 251<br />

Figure 7.7. Decomposition <strong>of</strong> unit exergy consumptions associated with wastes. ...... 252<br />

Figure 7.8. Decomposition <strong>of</strong> plant products. .............................................................. 252<br />

Figure 7.9. Decomposition <strong>of</strong> malfunctions. ................................................................ 253<br />

Figure 7.10. Decomposition <strong>of</strong> malfunctions corresponding to the boiler................... 254<br />

Figure 7.11. Decomposition <strong>of</strong> malfunctions associated with the residues.................. 254<br />

Figure 7.12. Decomposition <strong>of</strong> malfunction cost. ........................................................ 256<br />

Figure 7.13. Decomposition <strong>of</strong> malfunction cost corresponding to boiler................... 256<br />

Figure 7.14. Decomposition <strong>of</strong> malfunction cost associated with wastes.................... 257<br />

Figure 7.15 Decomposition <strong>of</strong> unit exergy consumptions............................................ 264<br />

Figure 7.16. Decomposition <strong>of</strong> unit exergy consumptions........................................... 264<br />

Figure 7.17. Decomposition <strong>of</strong> unit exergy consumption associated with the boiler. . 265<br />

Figure 7.18. Decomposition <strong>of</strong> unit exergy consumptions associated with the residues.<br />

...................................................................................................................................... 265<br />

Figure 7.19. Decomposition <strong>of</strong> plant product............................................................... 266<br />

Figure 7.20. Decompositions <strong>of</strong> malfunctions. ............................................................ 267<br />

Figure 7.21. Decomposition <strong>of</strong> malfunctions associated with the boiler. .................... 267<br />

Figure 7.22. Decomposition <strong>of</strong> malfunctions associated with the residues.................. 268<br />

Figure 7.23. Decomposition <strong>of</strong> malfunction costs........................................................ 269<br />

Figure 7.24. Decomposition <strong>of</strong> malfunction cost associated with the boiler ............... 269<br />

Figure 7.25. Decomposition <strong>of</strong> malfunction costs associated with the residues. ......... 270<br />

Figure 7.26. Fuel impact <strong>of</strong> free <strong>diagnosis</strong> variables (1).............................................. 271<br />

Figure 7.27. Fuel impact <strong>of</strong> free <strong>diagnosis</strong> variables (2).............................................. 271<br />

Figure 8.1. Error in the value <strong>of</strong> impact factors and impact standard deviation .......... 283<br />

in boiler efficiency........................................................................................................ 283<br />

Figure 8.2. Error in the value <strong>of</strong> impact factors and impact standard deviation .......... 283<br />

in cycle heat rate........................................................................................................... 283<br />

Figure 8.3. Error in the value <strong>of</strong> impact factors and impact standard deviation .......... 284<br />

in unit heat rate. ............................................................................................................ 284<br />

Figure 8.4. Training process for boiler efficiency neural network............................... 286<br />

Figure 8.5. Training process for cycle heat rate neural network. ................................. 286<br />

Figure 8.6. Training process for unit heat rate neural network. ................................... 287<br />

xviii


Figure 8.7. Comparison <strong>of</strong> real and calculated values <strong>of</strong> boiler efficiency.................. 287<br />

Figure 8.8. Comparison <strong>of</strong> real and calculated cycle heat rate..................................... 288<br />

Figure 8.9. Comparison <strong>of</strong> real and calculated unit heat rate....................................... 288<br />

Figure 8.10. Variable increment and impact for the six variables most influent on boiler<br />

efficiency: coal HHV, aggregated boiler effectiveness, ambient temperature, moisture in<br />

coal, hydrogen in coal and carbon in ashes. ................................................................. 290<br />

Figure 8.11. Relation between variable increment and its impact on boiler efficiency for<br />

several variables: tempering and primary air relation, air infiltration in preheaters,<br />

average cold-side temperature in primary air preheaters, primary air-coal ratio, sulphur<br />

mass fraction in coal and low pressure isoentropic efficiency. .................................... 291<br />

Figure 8.12. Variable increment and impact for the six variables most influent on cycle<br />

heat rate: ambient temperature, low pressure turbine isoentropic efficiency, cooling<br />

tower effectiveness, condenser effectiveness, cooling water flow rate and high pressure<br />

turbine isoentropic efficiency. ...................................................................................... 292<br />

Figure 8.13. Relation between variable increment and its impact on cycle heat rate for<br />

several variables: relative humidity, intermediate pressure 1 isoentropic efficiency,<br />

intermediate pressure 2 isoentropic efficiency, gross electric power, coal high heating<br />

value and carbon mass fraction in coal......................................................................... 293<br />

Figure 8.14. Variable increment and impact for the six variables most influent on unit<br />

heat rate: low pressure turbine isoentropic efficiency, ambient temperature, coal high<br />

heating value, cooling tower effectiveness moisture mass fraction in coal and condenser<br />

effectiveness. ................................................................................................................ 294<br />

Figure 8.15. Relation between variable increment and its impact on unit heat rate for<br />

several variables: carbon mass fraction in coal, relative humidity, sootblowing steam<br />

flow rate, intermediate pressure 3 turbine isoentropic efficiency, hot side temperature<br />

difference in 6th water heater and average cold-side temperature in primary air<br />

preheaters...................................................................................................................... 295<br />

xix


Nomenclature<br />

a Generic point. Output <strong>of</strong> a neuron. Sound speed. Area per unit <strong>of</strong> volume<br />

A Area. Ambient<br />

b Generic point. Bias.<br />

B Boiler incomes different from fuel.<br />

Bo Boltzman number<br />

c Specific heat. Generic point. Speed.<br />

C Heat capacity rate<br />

CFD Computer fluid dynamics<br />

CP Component parameter<br />

d Differential variation<br />

D Diameter.<br />

DF Dysfunction<br />

e Intensive magnitude, Global efficiency indicator<br />

ee Error in the calculation <strong>of</strong> e from x<br />

E Exergy flow in a productive structure, Extensive magnitude. Energy entering the<br />

boiler<br />

E * Exergetic cost <strong>of</strong> a flow<br />

EP Electrostatic precipitator<br />

EZ1 Primary economizar<br />

EZ2 Secondary economizer<br />

f Global efficiency indicator. Function. Fraction.<br />

F Fuel. Factor. Fouling thermal resistence. View factor<br />

FDF Forced draft fan<br />

FT<br />

Fuel entering the plant<br />

g Intensive magnitude. Gravity constant.<br />

G Extensive magnitude. Irradiation<br />

h Enthalpy. Convection coefficient<br />

H Height. Enthalpy<br />

HR Heat rate<br />

i Coefficient <strong>of</strong> increment <strong>of</strong> intrinsic malfunctions.<br />

I Irreversibility. Impact<br />

IDF Induced draft fan<br />

k Constant. Thermal conductivity<br />

k * Unit exergy cost. Marginal unit exergy cost<br />

*<br />

k Average unit exergy cost.<br />

*<br />

k R Unit exergy cost associated with wastes<br />

L Losses<br />

LMTD Log-mean temperature difference<br />

m Flow rate<br />

xxi


M Coefficient<br />

MFR Mass flow ratio<br />

MF Malfunction<br />

MF * Malfunction cost<br />

MR Malfunction associated with wastes<br />

MR * Malfunction cost associated with wastes<br />

n Components in the thermoeconomic model. Number <strong>of</strong> observations in linear<br />

regression. Rotation speed. Exponent. Intermediate value in a neuron.<br />

NTU Number <strong>of</strong> transfer units<br />

N Dimension <strong>of</strong> predictor in linear regression<br />

P Product,<br />

p Pressure. Number <strong>of</strong> terms in linear regression<br />

PAF Primary air fan<br />

PAH Primary air heater<br />

PF Power factor<br />

q Flow rate<br />

Q Heat. Volumetric flow rate.<br />

r Number <strong>of</strong> governing parameters. Dimension <strong>of</strong> the input vector in a neuron.<br />

Relation<br />

RH Reheater<br />

RPS Reference performance state<br />

s Entropy<br />

S Number <strong>of</strong> neurons in a layer. Surface.<br />

SAH Secondary air preheater<br />

SH1 Primary superheater<br />

SH2 Secondary superheater<br />

SH3 Final superheater<br />

SP Set point<br />

T Temperature<br />

Td<br />

xxii<br />

Dissipation temperature<br />

TDCA Drain cooling approach temperature<br />

TTD Terminal temperature difference<br />

U Global heat transfer coefficient. Utile heat.<br />

UE Used enthalpy<br />

umb Umburned carbon related to coal<br />

umbres Umburned carbon related to ashes<br />

W Mass fraction <strong>of</strong> moisture<br />

W Power<br />

x Thermodynamic variable. Quality. Molar fraction<br />

ˆx Corrected variable<br />

y Response in linear regression<br />

z Teeth number in a seal.<br />

Z Ash in coal<br />

Matrices and arrays<br />

a Output <strong>of</strong> a layer <strong>of</strong> neurons<br />

AD Matrix <strong>of</strong> correction factors with free <strong>diagnosis</strong> variables<br />

AX Matrix <strong>of</strong> correction factors with all thermodynamic variables<br />

b Bias <strong>of</strong> a layer or neurons<br />

bd Independent terms vector for correction with free <strong>diagnosis</strong> variables


DD Matrix to connect free <strong>diagnosis</strong> variables <strong>of</strong> first and second level<br />

DE Matrix <strong>of</strong> impacts on e due to two levels <strong>of</strong> <strong>causal</strong>ity<br />

DP Matrix to connect measured variables and free <strong>diagnosis</strong> variables<br />

e Error vector<br />

EA Matrix <strong>of</strong> impacts induced by other free <strong>diagnosis</strong> variables<br />

EAD Matrix <strong>of</strong> sensitivity for impacts induced by other free <strong>diagnosis</strong> variables<br />

ed Vector <strong>of</strong> sensitivity coefficients <strong>of</strong> e related to xc<br />

ED Diagonal matrix <strong>of</strong> sensitivity coefficientys <strong>of</strong> e related to xc<br />

EDD Matrix <strong>of</strong> sensitivity coefficients <strong>of</strong> e related to two levels <strong>of</strong> <strong>causal</strong>ity<br />

EP Matrix to connect measured variables and impacts <strong>of</strong> free <strong>diagnosis</strong> variables.<br />

ep Vector to connect measured variables and efficiency indicator e.<br />

er Vector to connect residues and indicator e<br />

ex Vector <strong>of</strong> sensitivity coefficientys <strong>of</strong> e related to x<br />

g Restrictions. Gradient<br />

H Hessian matrix<br />

J Jacobian matrix<br />

JD General matrix <strong>of</strong> the <strong>diagnosis</strong> problem<br />

JP General matrix <strong>of</strong> the performance test problem<br />

JR Jacobian matrix <strong>of</strong> restrictions<br />

JRD Jacobian matrix <strong>of</strong> restrictions <strong>of</strong> the <strong>diagnosis</strong> problem<br />

JRP Jacobian matrix <strong>of</strong> restrictions <strong>of</strong> the performance test problem<br />

k *<br />

Unit exergy costs<br />

kd Vector to connect κ and xd<br />

KD Matrix to connect κ and xd<br />

Matrix <strong>of</strong> unit exergy consumptions<br />

kx Vector to connect κ and x<br />

M Generic matrix<br />

N Generic matrix<br />

p Input vector in a neuron.<br />

P Product<br />

Ps<br />

Product <strong>of</strong> the system<br />

P> Product operator<br />

R Restrictions. Correlation matrix<br />

RD Restrictions <strong>of</strong> the <strong>diagnosis</strong> problem<br />

rs Residues<br />

RP Restrictions for the performance test problem<br />

tx Vector to connect θ and x<br />

td Vector to connect θ and xd<br />

TD Matrix to connect θ and xd<br />

UD<br />

Unit matrix<br />

VD Matrix <strong>of</strong> <strong>diagnosis</strong> variables<br />

VP Matrix <strong>of</strong> measured variables<br />

W Weight matrix<br />

wx Vector to connect ω and x<br />

wd Vector to connect ω and xd<br />

x Thermodynamic variables. Predictor in linear regression<br />

X Diagonal matrix <strong>of</strong> thermodynamic variables. Matrix <strong>of</strong> observations <strong>of</strong> the<br />

predictor in linear regression.<br />

y Vector <strong>of</strong> observations <strong>of</strong> response in linear regression.<br />

xxiii


Greek<br />

α Coefficient. Learning rate. Absorptivity<br />

β Coefficient. Mass transfer coefficient. Angle<br />

β Vector <strong>of</strong> coefficients<br />

ˆβ Coefficient estimator in linear regression<br />

ˆ<br />

β Vector <strong>of</strong> coefficient estimator in linear regression<br />

Δ Increment<br />

ε Heat exchanger effectiveness. Error. Emisivity. Residual.<br />

ε Vector <strong>of</strong> errors in linear regression.<br />

η Isoentropic efficiency<br />

θ Unit exergy consumption associated with residues. Internal parameter<br />

κ Unit exergy consumption<br />

ˆκ Unit exegy consumption corrected by variations in operation point<br />

κe Unit exergy consumption <strong>of</strong> external resources<br />

μ Parameter in Levenberg-Marquardt algorithm<br />

ρ Density. Ratio <strong>of</strong> entropy increment.<br />

σ Entropy variation. Standard deviation. Stefan-Boltzman constant<br />

σ 2 Uncertainty vector<br />

τ Set <strong>of</strong> thermodynamic variables<br />

υ Specific volume<br />

φ Factor<br />

Φ Flow coefficient. Relative humidity.<br />

χ Internal or endogenous variables<br />

ω Product <strong>of</strong> a component leaving the system. Enthalpy variation<br />

Subscripts<br />

a Air<br />

ac Adiabatic combustion<br />

amb Ambient<br />

b Boiler<br />

B Radiation and convection<br />

c Cycle. Condenser. Cold. Corrected.<br />

calc Calculated<br />

CO Carbon monoxide<br />

conv Convection<br />

crs Cold reheated steam<br />

dep Dependent<br />

d Diagnosis. Design. Sensible heat in ashes<br />

dp Drum purge<br />

e Initial point<br />

ext External<br />

f Fuel<br />

F Furnace<br />

fdf Forced draft fan<br />

fw Feeding water<br />

g Final point. Gases<br />

gen Generator<br />

go Gases output<br />

gs Sensible heat in flue gases<br />

xxiv


h Hot<br />

H Hydrogen<br />

hrs Hot reheated steam<br />

ind Induced<br />

indep Independent<br />

int Intrinsic. Internal<br />

in Input<br />

l Linealization<br />

lm Log-mean<br />

LP Low pressure<br />

mA Air moisture<br />

mat Material<br />

max Maximum<br />

mec Mechanical<br />

mf Coal moisture<br />

min Minimum<br />

ms Main steam<br />

n Net<br />

nd Non dimensional<br />

ng Natural gas<br />

o Output<br />

ol Other loses<br />

p Pump. Measured in plant. Pressure. Point.<br />

P Product<br />

paf Primary air fan<br />

r Control system. Restrictions<br />

rad Radiation<br />

rd Restrictions <strong>of</strong> the <strong>diagnosis</strong> problem<br />

rp Restrictions for performance test problem<br />

res Residual<br />

ref Reference<br />

rs Residual<br />

s Isoentropic. Steam<br />

sat Saturation<br />

sb Sootblowing<br />

sl Saturated liquid<br />

sru Air to sulphur removing unit.<br />

ss Saturated steam<br />

st Static<br />

t Thermodynamic model. Turbine. Per unit <strong>of</strong> time.<br />

tot Total<br />

tp Turbo-pump<br />

tr Transmission<br />

tw Tempering water<br />

u Unit<br />

uc Unburned carbon in ashes<br />

ul Uncontrolled leakages<br />

v Volumetric. Valve<br />

var Variables<br />

w Water. Wall<br />

xxv


wb Wet bulb<br />

0 Environment, reference example<br />

1 First <strong>causal</strong>ity level<br />

1ch Primary air coil heater<br />

2 Second <strong>causal</strong>ity level<br />

2ch Secondary air coil heater<br />

Superscripts<br />

ac Ambient conditions<br />

av Average<br />

fq Fuel quality<br />

int Intrinsic<br />

lr Linear regression<br />

oc Other components<br />

r Residual<br />

rep Representative<br />

sp Set points<br />

t Transpose<br />

0 Reference<br />

1 Actual situation<br />

xxvi


1 Introduction<br />

The word <strong>diagnosis</strong> refers to the activities aimed at the detection <strong>of</strong> anomalies in any<br />

system. In general, the interest <strong>of</strong> locating these anomalies appears because they can<br />

lead, if not solved, to catastrophic effects.<br />

In energy systems, besides this kind <strong>of</strong> failures that potentially originates<br />

component breaking and, perhaps, plant shut <strong>of</strong>f, there are other anomalies which are <strong>of</strong><br />

interest. These situations might never originate a catastrophic failure but may entail<br />

huge economic expenses due to the additional fuel consumption they cause.<br />

In this context, it is interesting not only to detect anomalies but also to quantify the<br />

effect <strong>of</strong> each one <strong>of</strong> them in the variation <strong>of</strong> fuel consumption. This quantification is<br />

crucial in order to decide whether or not to repair or replace a component. This task is<br />

the objective <strong>of</strong> thermoeconomic <strong>diagnosis</strong>.<br />

Thermoeconomics can be considered as the science <strong>of</strong> energy saving and emerges<br />

from the combination <strong>of</strong> the Second Principle <strong>of</strong> Thermodynamics and Economics. It<br />

provides several tools and concepts to deal with the <strong>diagnosis</strong> problem. The main <strong>of</strong><br />

them is the fuel impact formula, which links the variations <strong>of</strong> the unit exergy<br />

consumptions <strong>of</strong> the plant components (the indicators <strong>of</strong> their individual behaviour) to<br />

the fuel increment they cause.<br />

Thermoeconomic <strong>analysis</strong> provides a solid conceptual basis for the <strong>diagnosis</strong>, but in<br />

the practical application to real examples, difficulties appear because <strong>of</strong> the problem <strong>of</strong><br />

induced effects. The presence <strong>of</strong> these effects makes it more difficult to identify the<br />

components where anomalies take place, and the development <strong>of</strong> <strong>methods</strong> to eliminate<br />

them is a wide field <strong>of</strong> research. (Valero et al., 2004a,b).<br />

Due to the difficulty to develop <strong>methods</strong> to deal successfully with induced effects,<br />

some authors have developed <strong>diagnosis</strong> methodologies which avoid the use <strong>of</strong><br />

Thermoeconomic <strong>analysis</strong> and are directly based on a thermodynamic representation <strong>of</strong><br />

1


Chapter 1<br />

the studied system. Among these proposals, the <strong>diagnosis</strong> algorithm developed by<br />

Correas (2001, 2004), constitutes a systematic approach, which provide good results in<br />

the application to real systems.<br />

Finally, there are other <strong>methods</strong>, based on techniques such as neural networks,<br />

which look directly for the relations among the variables, with the use <strong>of</strong> neither<br />

thermoeconomic nor thermodynamic models.<br />

1.1 Justification, objectives and content <strong>of</strong> the thesis<br />

In this framework, the <strong>diagnosis</strong> algorithm presents characteristics <strong>of</strong> rigor,<br />

completeness and suitability for practical application which makes it an interesting tool<br />

for the <strong>diagnosis</strong> <strong>of</strong> whatever energy system. However, their capabilities should be<br />

incremented and their drawbacks should be minimized. Besides, it is necessary to<br />

compare the different <strong>diagnosis</strong> <strong>methods</strong> available, in order to identify the strengths and<br />

weaknesses <strong>of</strong> each one, to state their common points and differences and to look for<br />

possible ways <strong>of</strong> integration. Last but not least, it should be kept in mind that, despite <strong>of</strong><br />

the interest <strong>of</strong> rigorous fundamentals, the aim <strong>of</strong> these techniques is to deal with real<br />

situations; so that, the application <strong>of</strong> the methodologies to an actual power plant is<br />

crucial.<br />

The objectives <strong>of</strong> this thesis are:<br />

2<br />

• The development <strong>of</strong> a complete <strong>diagnosis</strong> methodology (quantitative<br />

<strong>causal</strong>ity <strong>analysis</strong>); starting from the <strong>diagnosis</strong> algorithm, but introducing<br />

new features and concepts: effect <strong>of</strong> measurement errors, influence <strong>of</strong> the<br />

choice <strong>of</strong> the <strong>diagnosis</strong> variables, <strong>causal</strong>ity chains…<br />

• Practical application <strong>of</strong> the methodology in a real-life example. The <strong>analysis</strong><br />

has to be made for a time span <strong>of</strong> several years, in order to be able to<br />

appreciate the effect <strong>of</strong> the degradation <strong>of</strong> the components.<br />

• Demonstration that the potential disadvantage <strong>of</strong> the method (the<br />

linealization <strong>of</strong> equations) has negligible influence <strong>of</strong> results.<br />

• Connection <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> and thermoeconomic <strong>analysis</strong>:<br />

development <strong>of</strong> the theoretical basis and application <strong>of</strong> the working example.


Introduction<br />

• Analysis <strong>of</strong> the applicability <strong>of</strong> neural networks and linear regression to the<br />

<strong>diagnosis</strong> problem. Comparison <strong>of</strong> the results provided by these <strong>methods</strong> and<br />

quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

To accomplish these objectives, the thesis is developed in 7 chapters plus<br />

introduction and conclusion. The aim <strong>of</strong> chapter 2 is to review methodologies for the<br />

<strong>diagnosis</strong> <strong>of</strong> thermal systems; above all, the fundamentals <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

and the <strong>diagnosis</strong> algorithm. Quantitative Causality Analysis is developed in chapter 3:<br />

formulation, influence <strong>of</strong> measurement error, <strong>analysis</strong> <strong>of</strong> the effect <strong>of</strong> linearization,<br />

choice <strong>of</strong> suitable free <strong>diagnosis</strong> variables and theory <strong>of</strong> <strong>causal</strong>ity chains. Chapter 4 has<br />

two parts. In the first one, quantitative <strong>causal</strong>ity <strong>analysis</strong> and thermoeconomic <strong>diagnosis</strong><br />

are connected. In the second, the applicability <strong>of</strong> linear regression and neural networks<br />

to the <strong>diagnosis</strong> problem is analyzed.<br />

The thermal system chosen as a test-bench is a 3x350 MW conventional coal-fired<br />

power plant. It is presented in Chapter 5 as an example <strong>of</strong> the methodology to choose<br />

the suitable free <strong>diagnosis</strong> variables according to the phenomena to be diagnosed and<br />

information available. The study is focused on the steam cycles, while the <strong>analysis</strong> <strong>of</strong><br />

the boilers has been substantially simplified. Results <strong>of</strong> the anamnesis (or the repeated<br />

<strong>diagnosis</strong>), <strong>of</strong> the three units <strong>of</strong> the working example during more than 6 years are<br />

presented on Chapter 6. An exhaustive <strong>analysis</strong> <strong>of</strong> all free <strong>diagnosis</strong> variables is made,<br />

demonstrating the applicability and coherence <strong>of</strong> the approach. Besides, it is proved that<br />

the error provided by the method is negligible in practical applications. In chapter 7, the<br />

theory linking quantitative <strong>causal</strong>ity <strong>analysis</strong> and thermoeconomic <strong>diagnosis</strong>, developed<br />

in Chapter 4, is successfully applied to the working example. Finally, neural networks<br />

and linear regression are applied in Chapter 8, and results compared with those obtained<br />

with the quantitative <strong>causal</strong>ity <strong>analysis</strong>.


Chapter 1<br />

4


2. State <strong>of</strong> the art <strong>of</strong> thermoeconomic<br />

<strong>diagnosis</strong>.<br />

In this chapter, the main approaches to thermoeconomic <strong>diagnosis</strong> <strong>of</strong> thermal<br />

systems are revised. First <strong>of</strong> all, the main concepts and tools provided by<br />

Thermoeconomics are presented. Then, several proposals for applying these tools are<br />

described. Finally, some <strong>diagnosis</strong> systems are presented.<br />

2.1 Thermoeconomic approach to <strong>diagnosis</strong>.<br />

Exergy <strong>analysis</strong> is based on the application <strong>of</strong> this magnitude to the study <strong>of</strong> thermal<br />

systems (Kotas, 1985). By using this approach, the analyst can identify not only the<br />

losses <strong>of</strong> energy quantity but also the losses in energy quality (irreversibilities).<br />

However, according to the Principle <strong>of</strong> non-equivalence <strong>of</strong> the irreversibilities, the<br />

same variation <strong>of</strong> the irreversibility implies different additional consumption <strong>of</strong><br />

resources depending on the component <strong>of</strong> the plant in which it takes place.<br />

Thermoeconomic <strong>analysis</strong> goes a step forward than exergy <strong>analysis</strong> by taking into<br />

account this fact by introducing the concept <strong>of</strong> cost. In a few words, the cost <strong>of</strong> a flow<br />

can be defined as the amount <strong>of</strong> resources needed to produce that flow.<br />

2.1.1 Thermoeconomic model.<br />

Thermoeconomic <strong>analysis</strong> is based on the representation <strong>of</strong> a thermal system by<br />

using a thermoeconomic model. This is done by describing the system by means <strong>of</strong> a<br />

productive structure. The productive structure (also “functional diagram” or “structural<br />

scheme”) is a graph which shows a set <strong>of</strong> relations defining the interaction among<br />

components themselves and the environment. The nodes <strong>of</strong> the graph are:<br />

• Control volumes which can be physical devices or groups <strong>of</strong> them.<br />

5


Chapter 2<br />

• Junctions and branches <strong>of</strong> flows required to define the interaction between<br />

components. They can be either physical devices appearing in the plant or<br />

fictious elements introduced to build the productive structure (Frangopoulos,<br />

1983).<br />

The flows E <strong>of</strong> the productive structure are described by exergy (Valero et al.,<br />

1986a), sometimes split into thermal and mechanical components (Tsatsaronis et al.,<br />

1990) or using also negentropy flows (Frangopoulos, 1983; Frangopoulos, 1986; Von<br />

Spakovsky and Evans, 1987; Von Spakovsky and Evans, 1990). The name <strong>of</strong> this<br />

structure comes from the concept <strong>of</strong> purpose. Each subsystem has one or more entering<br />

flows that represent the resources or fuel and exiting flows (product). Sometimes, byproducts<br />

or residues can be also considered (Lozano and Valero, 1993a; Torres et al.,<br />

2003). Flows are named by using the notation Eij, and it means that the flow comes from<br />

the component i and goes to the component j. In Figure 2.1, an example <strong>of</strong> productive<br />

structure is shown.<br />

6<br />

E01<br />

2<br />

E25<br />

1<br />

E15<br />

E50<br />

5<br />

E32<br />

Figure 2.1. Example <strong>of</strong> a productive structure.<br />

Due to irreversibilities, in every component, product is smaller than fuel. So that, the<br />

amount <strong>of</strong> resources per unit <strong>of</strong> exergy increases throughout the plant. This leads us to<br />

the concept <strong>of</strong> cost.<br />

E53<br />

E54<br />

3<br />

4<br />

E30<br />

E40


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

2.1.2 Exergetic cost.<br />

The cost <strong>of</strong> a flow Eij (E * ij) is defined as the amount <strong>of</strong> external resources in terms <strong>of</strong><br />

exergy that are required to produce the flow Eij. The unit exergy cost is the relation<br />

between the cost <strong>of</strong> a flow and its exergy:<br />

k<br />

E<br />

= 2.1<br />

*<br />

* ij<br />

ij<br />

Eij<br />

Given a thermoeconomic model <strong>of</strong> a system, the cost <strong>of</strong> every flow can be calculated<br />

by the cost assessment rules provided by the exergy cost theory (Valero et al., 1986a;<br />

Lozano and Valero, 1993b; Valero, 2006a):<br />

P1: The cost <strong>of</strong> the external resources is equal to its exergy:<br />

E = E<br />

2.2<br />

*<br />

0i 0i<br />

P2: In each component, the cost is conserved, so that, the cost <strong>of</strong> the fuel is equal to<br />

the cost <strong>of</strong> the product.<br />

where:<br />

P = F<br />

2.3<br />

* *<br />

i i<br />

n<br />

* *<br />

i ij<br />

j=<br />

0<br />

P = ∑ E<br />

2.4<br />

n<br />

* *<br />

i ji<br />

j=<br />

0<br />

F = ∑ E<br />

2.5<br />

P3: The exergy cost <strong>of</strong> the flows produced in a component is proportional to their<br />

exergy, so that they have the same unit exergy cost:<br />

k = k<br />

2.6<br />

* *<br />

ij P, i<br />

The unit exergy consumption is defined as the number <strong>of</strong> units <strong>of</strong> exergy that each<br />

component requires from the other components to obtain a unit <strong>of</strong> its product:<br />

E<br />

ij<br />

κ ij = 2.7<br />

Pj<br />

The sum <strong>of</strong> all unit exergy consumptions in a component is the inverse <strong>of</strong> the exergy<br />

efficiency <strong>of</strong> that component:<br />

k<br />

F<br />

∑ 2.8<br />

n<br />

j = κij<br />

=<br />

i= 0<br />

j<br />

Pj<br />

7


Chapter 2<br />

According to the structural theory (Valero et al., 1992, 1993), the productive<br />

structure can be built considering that each product includes only one flow. So that, the<br />

characteristic equation <strong>of</strong> the Thermoeconomic model can be written as follows:<br />

8<br />

i i0 ij j<br />

j=<br />

1<br />

n<br />

P = B + ∑ κ ⋅P<br />

2.9<br />

Or, in matrix notation:<br />

P= PS+ KP P 2.10<br />

where Ps is a (n x 1) vector representing the contribution <strong>of</strong> each component to the<br />

product <strong>of</strong> the whole system, and KP is a (n x n) matrix whose elements are the unit<br />

exergy consumption κij. The previous expression allows to obtain the product <strong>of</strong> each<br />

component from the final product <strong>of</strong> the system.<br />

P = P P 2.11<br />

where<br />

S<br />

( ) 1 −<br />

D<br />

P ≡ U − KP 2.12<br />

The total resources <strong>of</strong> the plant may be obtained by:<br />

F = κ P P 2.13<br />

t<br />

T e s<br />

t<br />

where κ ( κ ,..., κ )<br />

≡ is a (n x 1) vector whose elements contain the unit<br />

e 01 0n<br />

consumption <strong>of</strong> external resources. Finally, the unit cost <strong>of</strong> the flows can be obtained by<br />

using the following equation:<br />

* t<br />

P = e<br />

k P κ 2.14<br />

2.1.3 The fuel impact formula<br />

This equation was suggested by Valero et al. (1990, 1999) and developed by Reini<br />

(1994), Lozano et al. (1994) and Torres et al. (1999). It is very important for<br />

Thermoeconomic <strong>diagnosis</strong> because it relates the variation <strong>of</strong> fuel consumption <strong>of</strong> a<br />

system with the variation <strong>of</strong> unit exergy consumptions <strong>of</strong> the components <strong>of</strong> that system<br />

and to the variation <strong>of</strong> production:<br />

t t * t *<br />

Δ F = Δ κ + k ( x) Δ KP P x + k ΔP<br />

2.15<br />

( ) ( 0 )<br />

T e P P s<br />

In scalar format:<br />

n ⎛ n<br />

⎞<br />

Δ FT = kP j x Δ jiPi x + kP i x Δ i<br />

i= 1⎝ j=<br />

0<br />

⎠<br />

* *<br />

∑∑ ⎜ , ( ) κ ( 0 ) , ( ) ϖ ⎟<br />

2.16


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

where ωi is the part <strong>of</strong> the product <strong>of</strong> the plant coming from the component i. When<br />

the unit exergy consumption <strong>of</strong> a component increases Δ κ ji , the irreversibility in this<br />

component also increases in a quantity which is called malfunction (Torres et al., 1999;<br />

Lerch et al., 1999; Torres et al., 2002; Valero et al., 2002).<br />

ji ji i<br />

( )<br />

MF =Δ κ P x<br />

2.17<br />

( )<br />

i i i 0<br />

ji<br />

j=<br />

0<br />

0<br />

n<br />

MF =Δ k P x =∑ MF<br />

2.18<br />

This malfunction implies an additional amount <strong>of</strong> fuel called the malfunction cost.<br />

( )<br />

MF = k x MF<br />

2.19<br />

* *<br />

ji P, j ji<br />

n<br />

* *<br />

i ji<br />

j=<br />

0<br />

MF = ∑ MF<br />

2.20<br />

If MF0 is defined as:<br />

n<br />

* *<br />

0 ,<br />

i=<br />

1<br />

∑ P i( ) i<br />

2.21<br />

MF = k x Δϖ<br />

it can be proved that:<br />

n<br />

*<br />

T i<br />

i=<br />

0<br />

Δ F =∑ MF<br />

2.22<br />

There are two types <strong>of</strong> malfunctions. When the unit exergy consumption <strong>of</strong> a<br />

component increases due to a degradation <strong>of</strong> this component, it is called intrinsic<br />

malfuncion. When an intrinsic malfunction occurs, the operation point <strong>of</strong> the other<br />

components varies. So that, since the efficiency curves <strong>of</strong> the components are usually<br />

non-flat, variations in the specific consumption <strong>of</strong> other components appear, which<br />

leads to induced malfunctions. The irreversibility <strong>of</strong> a component can also vary due to a<br />

change in its product. This is called dysfunction.<br />

( ( ) 1)<br />

DF = k x − Δ P<br />

2.23<br />

i i i<br />

The problem <strong>of</strong> induced malfunction is the main drawback <strong>of</strong> using the<br />

Thermoeconomic approach to diagnose energy systems, because it makes very difficult<br />

to identify the real components which do not work properly. To overcome this problem,<br />

several researchers have proposed techniques, most <strong>of</strong> them based in the filtering <strong>of</strong><br />

these induced effects. These procedures, among others, are revised in the section 2.2.<br />

9


Chapter 2<br />

More details on thermoeconomic <strong>analysis</strong> can be seen in Valero and Torres (2006a).<br />

For further information on thermoeconomic <strong>diagnosis</strong> see Valero and Torres (2006b)<br />

2.1.4. Cost formation process <strong>of</strong> wastes (or residues).<br />

In previous sections, only productive flows have been considered. However, there<br />

are other flows (such as gases leaving the stack or heat dissipated in a condenser) which<br />

are necessary for the operation <strong>of</strong> the thermal system but do not have a productive<br />

purpose. These flows are called wastes or residues (Torres et al., 2006; Rangel, 2005).<br />

The cost <strong>of</strong> the wastes has to be charged to the component where they have been<br />

produced; for example, the cost <strong>of</strong> gases leaving the stack should be charged to the<br />

combustion process, and the cost <strong>of</strong> heat dissipated in a condenser is usually shared<br />

according to the entropy produced in the different components.<br />

Consequently it is necessary to enlarge the productive structure to include these<br />

issues. A waste flow produced in a component i and charged to the component j is<br />

represented as Rij. So that, it is possible to define unit exergy consumptions associated<br />

with the wastes:<br />

10<br />

R<br />

ij<br />

θ ij = 2.24<br />

Pj<br />

And to calculate cost associated with residues:<br />

k<br />

R<br />

= 2.25<br />

*<br />

* ij<br />

Rij<br />

Rij<br />

The variation <strong>of</strong> these unit exergy consumptions, entails malfunctions associated<br />

with the wastes:<br />

ji ji i<br />

( )<br />

MR =Δ θ P x<br />

2.26<br />

0<br />

( )<br />

i i i 0<br />

ji<br />

j=<br />

0<br />

n<br />

MR =Δ θ P x =∑ MR<br />

2.27<br />

These malfunctions entail their corresponding malfunction cost:<br />

( )<br />

MR = k x MR<br />

2.28<br />

* *<br />

ji PR, j ji<br />

n<br />

* *<br />

i ji<br />

j=<br />

0<br />

MR = ∑ MR<br />

2.29<br />

These malfunction costs associated with the residues have to be considered in the<br />

expression <strong>of</strong> fuel impact as a summation <strong>of</strong> malfunction costs:


n<br />

*<br />

n<br />

*<br />

T i i<br />

i= 0 i=<br />

1<br />

Δ F = MF + MR<br />

State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

∑ ∑ 2.30<br />

Finally, the fuel impact formula has to be complemented with the variation <strong>of</strong> unit<br />

exergy consumption associated with the residues:<br />

n ⎛ n<br />

* * * ⎞<br />

Δ FT = ∑∑ ⎜ ( kP, j( x) Δ κ jiPi( x0 ) + kPR, j( x) Δ θ jiPi( x0 ) ) + kP, i( x)<br />

Δϖi⎟<br />

2.31<br />

i= 1⎝ j=<br />

0<br />

⎠<br />

2.1.5. Other issues.<br />

In the previous sections, a brief review <strong>of</strong> Thermoeconomics has been presented in<br />

order to explain the basic concepts and the fuel impact formula. However, there are<br />

other interesting issues, perhaps not so related straightforward by thermoeconomic<br />

<strong>diagnosis</strong> but worth to be presented. This is the aim <strong>of</strong> this section.<br />

The definition <strong>of</strong> cost as a quotient between fuel entering the plant and a given flow<br />

introduced previously corresponds to an average cost ( *<br />

k ) and it is suitable for cost<br />

accounting <strong>methods</strong>. For optimization <strong>methods</strong>, what it is interesting is to know the<br />

impact <strong>of</strong> fuel entailed by an additional unit <strong>of</strong> a given flow; in other words, a marginal<br />

* cost, k (Serra et al., 1995):<br />

k<br />

E<br />

= 2.32<br />

E<br />

* 0<br />

i<br />

* ⎛∂E⎞ 0 k = ⎜ ⎟<br />

2.33<br />

⎝ ∂Ei<br />

⎠conditions<br />

In an attempt to unify average and marginal cost and, subsequently,<br />

thermoeconomic theories based on cost accounting and aimed at system optimization,<br />

the structural theory was developed. This theory is based on four premises:<br />

1.- Every energy system can be structurally described as a number <strong>of</strong> physical or<br />

logical components, n, connected to one another and to their environment though a set<br />

<strong>of</strong> m relations (mass, energy, money, information).<br />

2.- Every relation among component must be defined completely using an extensive<br />

magnitude which reckons, at the same time, its quantity and quality:<br />

Ei = qi⋅ ei<br />

or Gj = qj⋅ g j<br />

2.34<br />

3.- Each component, l, needs to have as many characteristic equations as relations<br />

which affect (enter) it. Characteristic equation is defined as that which relates the<br />

magnitude <strong>of</strong> a flow entering the component, Ei, to the flows leaving it, G1, G2,…, Gj,<br />

11


Chapter 2<br />

and a set <strong>of</strong> internal variables or endogenous variables: {χl} which describe the<br />

behaviour <strong>of</strong> the component, l:<br />

E = Eˆ G , G ,..., G , χ<br />

2.35<br />

12<br />

( 1 2 )<br />

i i j l<br />

4.- The characteristic equation is a homogeneous function <strong>of</strong> first order in each<br />

production interval, concerning the subset {Gl} <strong>of</strong> independent variables:<br />

( 1, 2,..., , ) ( 1, 2,...,<br />

, )<br />

λE = Eˆ λG λG λG χ = λEˆ G G G χ<br />

2.36<br />

i i j l i j l<br />

Under the previous hypotheses, average and marginal cost coincide. Thanks to this<br />

result, structural theory can serve as a common formalism to unify different<br />

thermoeconomic theories previously presented. For more details see: Valero et al.,<br />

1992; Valero et al., 1993; Serra et al., 1995; Erlach et al., 1998; Serra and Torres, 2006.<br />

An interesting point <strong>of</strong> structural theory is that it is general, so that it is possible to<br />

use other thermodynamic functions different from exergy to characterize flows and<br />

different sets <strong>of</strong> parameters to characterize the behaviour <strong>of</strong> the component. Royo and<br />

Valero (1995) propose to use internal parameters as universal indicators to characterize<br />

simple thermal processes:<br />

θ<br />

eg<br />

h − h<br />

=<br />

s − s<br />

e g<br />

e g<br />

Where h and s are enthalpy and entropy and e and g are the initial and the final<br />

points <strong>of</strong> the process. The use <strong>of</strong> these parameters allows to simplify characteristics<br />

equations <strong>of</strong> thermal systems (Valero et al., 1995).<br />

Although the use <strong>of</strong> exergy is widely accepted, it has several limitations (Valero,<br />

2006b), which have lead to the proposal <strong>of</strong> other magnitudes such as processable exergy<br />

(Valero and Guallar, 1991) or relative free energy (Valero and Royo, 1992).<br />

Thermoeconomic <strong>analysis</strong> provides numerical results very valuable for the <strong>analysis</strong>,<br />

synthesis, <strong>diagnosis</strong> and optimization <strong>of</strong> complex energy systems. However, if the<br />

concepts <strong>of</strong> exergy cost theory are combined with symbolic manipulators, its<br />

capabilities increase. This line <strong>of</strong> research is called symbolic exergoeconomics, and<br />

allows to obtain explicit formulae relating the efficiency <strong>of</strong> the whole system with<br />

efficiencies <strong>of</strong> components forming it (Valero and Torres, 1988; Valero and Torres,<br />

1990; Torres, 2006). The application <strong>of</strong> symbolic exergoeconomics for thermal system<br />

simulation can be seen in Torres et al. (1989) and an <strong>analysis</strong> <strong>of</strong> how this theory can<br />

contribute to the unification <strong>of</strong> different thermoeconomic theories is made in Valero et<br />

al. (1989). Valero et al. (1991) describe a program for symbolic computation <strong>of</strong><br />

2.37


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

exergoeconomic variables. A detailed description <strong>of</strong> the requirements for the<br />

development <strong>of</strong> s<strong>of</strong>tware for thermoeconomic <strong>analysis</strong> can be seen in Torres et al.<br />

(2007).<br />

Although Thermoeconomics has an academic origin, its final aim is to serve as<br />

useful tool for improving efficiency in both design and operation stages. So that,<br />

practical applications have been developed in parallel with theoretical research. The<br />

concept <strong>of</strong> exergy has been applied to analyze steam power plants: simulation<br />

(Alconchel, 1988; Alconchel et al., 1989) and monitoring (Valero et al., 1986b).<br />

Thermoeconomic <strong>analysis</strong> has been applied to a steam boiler (Lozano and Valero,<br />

1987) cogeneration plants (Valero et al., 1987; Muñoz and Valero, 1989), a sugar<br />

factory (Guallar and Valero, 1988) an air conditioning system with cogeneration (Tozer<br />

et al., 1996) fuel cell systems (Álvarez et al., 2006) a pressurized fluidised bed<br />

combustion power plant (Schwarz et al., 1997) and <strong>analysis</strong> and optimization <strong>of</strong> dualpurpose<br />

power and desalination plants (Uche et al., 2001; Uche et al., 2006).<br />

Plant monitoring has also been tackled by Thermoeconomics (Valero et al., 1996;<br />

Valero et al., 1999; Correas et al., 1999).<br />

Finally, it should be noted that the field <strong>of</strong> application <strong>of</strong> Thermoeconomics is not<br />

limited to artificial energy systems but can be a valuable tool to analyze and quantify<br />

natural resources (Valero, 1995). Furthermore, it has been proposed to connect<br />

Thermoconomics with the Aristotelian concept <strong>of</strong> cause (Valero and Carreras, 1990).<br />

2.2 Thermoeconomic <strong>diagnosis</strong> methodologies.<br />

To deal with the problem <strong>of</strong> induced malfunctions, several <strong>methods</strong> have been<br />

proposed. Some <strong>of</strong> them are directly based on the ideas proposed above, combined with<br />

the use <strong>of</strong> several types <strong>of</strong> filtration techniques. Others, however, use different<br />

approaches.<br />

To compare these methodologies, the TADEUS (Thermoeconomic Approach to the<br />

Diagnosis <strong>of</strong> Energy Utility Systems) initiative was launched. In the first paper <strong>of</strong> this<br />

series (Valero et al., 2004a), the problem <strong>of</strong> thermoeconomic <strong>diagnosis</strong> is presented and<br />

an example problem is proposed in order to encourage researchers interested in the topic<br />

to try their methodologies and then compare results. This example is a combined cycle<br />

power plant made up <strong>of</strong> two 125 MW gas turbines, two heat recovery steam generators<br />

13


Chapter 2<br />

and a 110 MW steam turbine, where several malfunctions have been simulated. In a<br />

second paper (Valero et al., 2004b), concepts <strong>of</strong> productive structure, cost, fuel impact<br />

formula and malfunctions and dysfunctions are revised. The third paper (Verda et al.,<br />

2003) focuses on the characteristics that a <strong>diagnosis</strong> system installed in a power plant<br />

should have and on the concept <strong>of</strong> reference condition. In other papers, several authors<br />

explain their methodologies and test them with the example proposed: (T<strong>of</strong>folo and<br />

Lazaretto, 2004; Zaleta et al., 2004a; Reini and Taccani, 2004; Verda, 2004, and<br />

Correas, 2004). Valero et al. (2004c) analyzes the importance <strong>of</strong> the fuel impact formula<br />

and describes the working example. Lazzaretto et al. (2006) compare results provided<br />

by several approaches.<br />

2.2.1 The filtration <strong>of</strong> the effects induced by the control system.<br />

When a malfunction occurs in a component <strong>of</strong> an energy system, the operation<br />

condition tends to move to a point that does not accomplish the set-points fixed by the<br />

operators (temperatures, power produced…). So that the control system intervenes in<br />

order to restore the values <strong>of</strong> these set-points. For example, the set points <strong>of</strong> a gas<br />

turbine are the power produced and the temperature <strong>of</strong> the gas entering the turbine. To<br />

maintain these two values, the control system modifies the angle <strong>of</strong> the guide vanes and<br />

the amount <strong>of</strong> fuel introduced in the combustion chamber. If the isoentropic efficiency<br />

<strong>of</strong> the turbine decreases (e. g. due to blade erosion), the power decreases. In this point,<br />

the control system intervenes in order to increase the fuel entering the combustion<br />

chamber. Due to the increment in fuel, temperature <strong>of</strong> gas entering the turbine increases,<br />

so the inlet guide vanes open in order to increase the air mass. As a result, the<br />

intervention <strong>of</strong> the control system allows to keep the set-points but modifies the<br />

operation points <strong>of</strong> all the parts <strong>of</strong> the plant, not only the ones where an anomaly takes<br />

place. In conclusion, the control system intervention produces induced anomalies that<br />

makes more difficult the identification <strong>of</strong> the actual malfunctioning component.<br />

The filtration <strong>of</strong> effects induced by the control system is the origin <strong>of</strong> a<br />

thermoeconomic <strong>diagnosis</strong> methodology developed by Verda (Verda et al., 2005a;<br />

Verda et al., 2005b; Verda 2004, Verda et. al. 2004, Verda et. al. 2002a, Verda et. al.<br />

2002b). This author defines the free condition <strong>of</strong> a malfunctioning thermal system as the<br />

state <strong>of</strong> that system characterized by the same position <strong>of</strong> the governing parameters as<br />

that for the reference condition, but containing the anomalies occurring at the actual<br />

operating condition. In the example <strong>of</strong> the gas turbine, the free condition would have the<br />

14


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

same isoentropic efficiency as the malfunctioning condition but the same amount <strong>of</strong> fuel<br />

and guide vanes angle as the reference condition.<br />

The free condition is a fictitious state <strong>of</strong> the system, so that it can not be determined<br />

by any measurement. It can only be calculated mathematically, for example, with a<br />

simulator. If the anomalies are low enough, the operation condition is close to the<br />

reference condition. So that, the values <strong>of</strong> the exergy flows in the free condition can be<br />

obtained by using a linear approximation:<br />

( )<br />

∂E<br />

E E ∑ x x<br />

2.38<br />

i =<br />

free i +<br />

op<br />

r<br />

i<br />

j= 1 ∂x<br />

j<br />

j −<br />

ref jop<br />

where xj is one <strong>of</strong> the r governing parameters <strong>of</strong> the system. The partial derivatives can<br />

be obtained by using r different particular working conditions. The idea <strong>of</strong> control<br />

system intervention leads us to two new concepts: Malfunction induced by the control<br />

system and cost <strong>of</strong> the control system (Verda et. al. 2004). Considering the effect <strong>of</strong> the<br />

control system, the fuel impact can be divided into three kinds <strong>of</strong> malfunctions:<br />

1. The intrinsic malfunction, which corresponds to the component where the<br />

anomaly takes place. Assuming that Δκgh is the intrinsic effect in the h th<br />

component:<br />

( κ κ )<br />

MF = − ⋅ P<br />

2.39<br />

int ghfree ghref href<br />

2. The malfunction induced by variations in the component productions is<br />

expressed by:<br />

( κ κ ) free ref ref int<br />

n n<br />

⎛ ⎞<br />

MFind = ∑∑ ⎜ ij − ij ⎟⋅Pj−MF<br />

2.40<br />

j= 1⎝ i=<br />

0<br />

⎠<br />

3. Malfunction induced by the control system intervention can be obtained as<br />

follows:<br />

( κ κ )<br />

n n<br />

⎛ ⎞<br />

MFr = ∑∑ ⎜ ij −<br />

op ij P<br />

free ⎟⋅<br />

j<br />

2.41<br />

ref<br />

j= 1⎝ i=<br />

0<br />

⎠<br />

The cost <strong>of</strong> control system intervention is defined as:<br />

k<br />

F − F<br />

* Top Tfree<br />

r =<br />

Pext − P<br />

op ext free<br />

2.42<br />

where FT is the total fuel entering the system and Pext is the product <strong>of</strong> the plant.<br />

This parameter is an indicator <strong>of</strong> the influence <strong>of</strong> the control system intervention in the<br />

15


Chapter 2<br />

efficiency <strong>of</strong> the plant. If this cost is greater than the cost <strong>of</strong> the product, it means that<br />

the control system intervention induces a reduction in the efficiency <strong>of</strong> the system.<br />

The comparison <strong>of</strong> free condition and reference condition has been used to diagnose<br />

a gas-turbine based cogeneration system (Verda et. al. 2004). This approach has been<br />

suitable to detect single anomalies, above all compared to the direct comparison <strong>of</strong><br />

operation condition and reference condition. However, when several anomalies appear,<br />

the filtration <strong>of</strong> the influence <strong>of</strong> the control system is not enough.<br />

Also in the free condition, the operation point <strong>of</strong> the components <strong>of</strong> the system is not<br />

the same as in the reference condition. So that, due to the fact that efficiency curves are,<br />

in general, non flat, induced malfunctions appear. To overcome this problem, the<br />

behaviour <strong>of</strong> the components depending on the product should be introduced. One<br />

possibility is to consider the variation <strong>of</strong> unit exergy consumption (Verda et. al. 2002a,<br />

2002b):<br />

16<br />

n ⎛∂κ⎞ ji<br />

ˆ κ ji = κ ji + E<br />

ref ∑ ⎜ ⋅Δ<br />

⎜ jl ⎟<br />

l= 0 ∂E<br />

⎟<br />

⎝ jl ⎠<br />

2.43<br />

where the derivatives can be obtained by using a set <strong>of</strong> operation conditions, and<br />

ΔEjl is calculated by comparing free and reference condition. The value ˆ k ji takes into<br />

account the variation <strong>of</strong> unit exergy consumption due to the variation <strong>of</strong> operation point,<br />

which should be taken into account in order to identify the actual malfunctioning<br />

component.<br />

( )<br />

free ref<br />

Δ κ = κ −κ −Δ ˆ κ<br />

2.44<br />

jiint ji ji ji<br />

Another possibility to filtrate this effect is to consider the dependence <strong>of</strong> the product<br />

on each resource (Verda 2004):<br />

n<br />

ˆ<br />

⎛ ∂E<br />

⎞<br />

i<br />

Ei = Ei + E<br />

ref ∑ ⎜ ⋅Δ li⎟´<br />

2.45<br />

l= 0 ⎝∂Eli ⎠<br />

and then to calculate ˆ k ji<br />

Eji E<br />

ˆ<br />

free jiref<br />

Δ k ji = − 2.46<br />

Eˆ<br />

P<br />

i<br />

iref<br />

The filtration <strong>of</strong> induced effects due to control system intervention and to non-flat<br />

efficiency curves allow to highlight intrinsic anomalies that could not be detected by the<br />

direct comparison <strong>of</strong> operation and reference conditions. This approach is applied by to<br />

diagnose a combined cycle (Verda, 2004; Verda et al., 2005a, 2004b and 2005b). In the


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

first two papers, a zooming strategy is proposed in order to first locate the macrocomponent<br />

where the anomaly occurs and then to identify the specific malfunctioning<br />

component inside the macro-component. In the two last papers, the effects <strong>of</strong> splitting<br />

the exergy into its thermal and mechanical components in order to diagnose a<br />

malfunction in a heat recovery steam generator are studied.<br />

Although the ideas presented above can be used to identify the malfunctioning<br />

components, sometimes they are not enough, mainly when there are several anomalies.<br />

The procedure can be improved by using anamnesis, which is the <strong>analysis</strong> <strong>of</strong> the history<br />

<strong>of</strong> the system (Verda, 2004). This technique is based on the <strong>diagnosis</strong> <strong>of</strong> the plant<br />

several times: malfunctions corresponding to the components where anomalies occur<br />

are expected to increase while induced malfunctions are expected to have a random<br />

evolution. To overcome the limitations <strong>of</strong> a linear model, the use <strong>of</strong> neural-networks<br />

based thermoeconomic model has also been proposed. (Verda 2004).<br />

Once the malfunctioning components have been identified, the fuel impact <strong>of</strong> each<br />

anomaly should be determined. Since this figure corresponds to the expected fuel<br />

savings if the anomaly is removed, this step can be named as prognosis (Verda, 2004).<br />

This information is very important to decide whether it is worth to repair the component<br />

or not. In theory, this could be determined by appliying the fuel impact formula<br />

previously presented. However, due to induced effects, this is not so easy. If this<br />

formula is rearranged, a corrected fuel impact can be defined.<br />

n<br />

*<br />

n<br />

*<br />

Tcorr T<br />

i= 1<br />

piei i=<br />

1<br />

i<br />

∑ ∑ 2.47<br />

Δ F =ΔF − k ⋅Δ P = MF<br />

Some <strong>of</strong> the malfunctions are intrinsic while the other are induced. If there are r<br />

intrinsic malfunctions, the impact <strong>of</strong> induced malfunctions should be split among the<br />

intrinsic ones.<br />

r<br />

*<br />

∑ int ( 1 )<br />

2.48<br />

Δ F = MF ⋅ + i<br />

Tcorr l l<br />

l=<br />

1<br />

The coefficients il can be adjusted by using the information available for the<br />

prognosis. In (Verda 2004) the complete approach including filtering <strong>of</strong> control system<br />

effects, filtering <strong>of</strong> non-flat efficiency curves effects and prognosis is applied to the<br />

Tadeus problem. Prognosis results are not as good as expected due to non-linearities.<br />

17


Chapter 2<br />

2.2.2 Other <strong>methods</strong> based on the filtration <strong>of</strong> induced effects<br />

Since the filtration <strong>of</strong> the intervention <strong>of</strong> the control system is usually not enough to<br />

detect the malfunctioning components, some authors propose <strong>methods</strong> based in the<br />

filtration <strong>of</strong> induced effects in only one step.<br />

Reini and Taccani (2004) propose a method based in the calculation <strong>of</strong> the cost <strong>of</strong><br />

malfunction induced by the fuel variation. First, they calculate the variation <strong>of</strong> specific<br />

consumption due to the variation <strong>of</strong> the product <strong>of</strong> the component:<br />

18<br />

∂kij<br />

∂ Pk<br />

ij =<br />

2.49<br />

∂E<br />

i<br />

This derivative can be obtained by using historical data. Then it is used to calculate<br />

the cost <strong>of</strong> malfunction induced by the variation <strong>of</strong> product:<br />

*<br />

*<br />

( MF ) ∑ k ∂<br />

i<br />

P<br />

= j<br />

j<br />

P<br />

k<br />

ij<br />

E<br />

i<br />

2.50<br />

Finally, intrinsic malfunction can be detected by analyzing the values <strong>of</strong> the cost <strong>of</strong><br />

malfunction (<br />

*<br />

MF i ) and the cost <strong>of</strong> malfunction induced by product variation ( ( MF i ) P<br />

*<br />

and using the following rules:<br />

a) If<br />

b) If<br />

malfunction.<br />

c) If<br />

*<br />

MF i is negative, it is probably an effect <strong>of</strong> induced malfunction.<br />

*<br />

MFi is very close to/smaller than ( MF i ) P<br />

*<br />

, it is the effect <strong>of</strong> induced<br />

*<br />

MF i related to a junction is not equal to zero, it is the effect <strong>of</strong> induced<br />

malfunction because inside junctions control volumes there are no exergy losses.<br />

This methodology is applied to the TADEUS problem. It is able to highlight the<br />

malfunctions separately, but when there is more than one at the same time it is not so<br />

useful.<br />

T<strong>of</strong>folo and Lazzaretto (2004) analyse the use <strong>of</strong> several indicators to identify the<br />

component where the anomaly takes place. They consider that the use <strong>of</strong> exergetic or<br />

thermoeconomic indicators is not suitable to identify malfunctions, because anomalies<br />

can be detected due a variation <strong>of</strong> specific curve <strong>of</strong> the components and not due a<br />

variation <strong>of</strong> specific consumption. They propose to isolate the component and to<br />

calculate an indicator based on the irreversibility variation but filtering the effect <strong>of</strong><br />

thermodynamic variables.<br />

h ∂I<br />

n<br />

h<br />

I res,<br />

h I h<br />

τ h,<br />

i<br />

i 1 τ h<br />

Δ ⋅<br />

= Δ − ∑<br />

= ∂<br />

2.51<br />

)


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

where Ires,h is the residual irreversibility <strong>of</strong> the h th component that is due to intrinsic<br />

malfunctions, ΔIh is the irreversibility variation and τh is the set <strong>of</strong> nh thermodynamic<br />

variables that influence the behaviour <strong>of</strong> component h. The indicator proposed is the<br />

non-dimensional value <strong>of</strong> Ires,h that is obtained by dividing it into the absolute value <strong>of</strong><br />

ΔIh. This approach has been applied to the TADEUS problem; it is able to highlight<br />

malfunctioning components but it not provided the fuel impact originated by each one<br />

<strong>of</strong> them. These authors analyze several thermoeconomic <strong>diagnosis</strong> methodologies and<br />

conclude that the true independent causes <strong>of</strong> malfunctions are the variation <strong>of</strong><br />

parameters <strong>of</strong> the thermodynamic model, and this information has to be included in the<br />

thermoeconomic <strong>diagnosis</strong> (Lazzaretto and T<strong>of</strong>folo, 2006).<br />

Valero et al. (1999) propose to use a simulator <strong>of</strong> the thermal system, in order to<br />

determine the effect <strong>of</strong> the variation <strong>of</strong> an operating parameter xr on a unit exergy<br />

consumption:<br />

0 0<br />

( x ) ( )<br />

Δ κ = κ x +Δ −κ<br />

x 2.52<br />

r<br />

ij ij r ij<br />

The application <strong>of</strong> the previous equation allows to decompose malfunctions in two<br />

terms, intrinsic and induced, depending on whether the operating parameters are<br />

associated to the studied component or not. It should be noted that these operating<br />

parameters are not only associated with the analyzed component, but parameters<br />

characterizing the behavior <strong>of</strong> all the components <strong>of</strong> the plant. The approach has been<br />

successfully applied to a steam cycle <strong>of</strong> a power plant (Valero et al., 1999). Zhang et al.<br />

(2007) apply a similar approach.<br />

2.2.3 Representation <strong>of</strong> the malfunctions in the h-s plane<br />

It has been shown that the direct use <strong>of</strong> classical Thermoeconomic <strong>analysis</strong> and the<br />

fuel impact formula is not always useful due to induced effects. For this reason, several<br />

authors have proposed <strong>diagnosis</strong> methodologies based on the filtration <strong>of</strong> these effects.<br />

Other authors consider that Thermoeconomic <strong>analysis</strong> is a good option for cost<br />

calculation but the <strong>diagnosis</strong> problem deals with the <strong>analysis</strong> <strong>of</strong> deviations around the<br />

reference operation point, so that the evolution <strong>of</strong> thermodynamic properties when a<br />

malfunction occurs should be the starting point <strong>of</strong> a <strong>diagnosis</strong> methodology. In this<br />

sense, several methodologies based on the trajectory in the h-s plane <strong>of</strong> the points when<br />

a malfunction occurs have been proposed.<br />

19


Chapter 2<br />

Zaleta et al. (2004b) proposes a thermo-characterization <strong>of</strong> power system<br />

components based on the representation in the ω, σ, MFR space. ω is defined as the<br />

enthalpy <strong>of</strong> the entering flow minus the enthalpy <strong>of</strong> the exiting flow, while σ is the<br />

entropy difference. The mass flow ratio (MFR) is the mass flow divided into the<br />

maximum mass flow. In this 3D space, the Reference Performance State (RPS) can be<br />

represented (Figure 2.2). RPS can be defined as “The range <strong>of</strong> operation <strong>of</strong> both<br />

intensive and extensive thermodynamic conditions (at full or partial load) guaranteed by<br />

the manufacturer, when neither intrinsic nor induced malfunctions arise in the<br />

component”.<br />

20<br />

MFR<br />

ω<br />

RPS<br />

Figure 2.2. RPS representation in the ω, σ, MFR space.<br />

Once the reference behaviour <strong>of</strong> a component has been represented by using RPS,<br />

malfunctions can be detected by comparing the actual values <strong>of</strong> ω and σ with the<br />

reference values for the same MFR. Examples <strong>of</strong> RPS and the impact <strong>of</strong> several types <strong>of</strong><br />

malfunctions in ω and σ can be seen in Zaleta et. al. (2004b) for several components <strong>of</strong><br />

a cycle and in Zaleta (1997) for several types <strong>of</strong> steam turbines.<br />

Once the deviations in ω and σ have been determined (the malfunctions have been<br />

detected), Zaleta et. al. (2004b) provide expressions to relate Δω and Δσ to the variation<br />

they produce in the heat rate <strong>of</strong> a steam cycle.<br />

An important parameter related to the <strong>analysis</strong> <strong>of</strong> malfunctions is the dissipation<br />

temperature (Royo et al., 1997). It is defined as the quotient between the increment in<br />

enthalpy and the increment in entropy <strong>of</strong> the exiting flow <strong>of</strong> a component when a<br />

differential malfunction occurs:<br />

σ


T<br />

d j<br />

⎛∂h⎞ out j<br />

= ⎜ ⎟<br />

⎜∂s⎟ ⎝ out j ⎠ Malfunction<br />

State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

This derivative is the tangent <strong>of</strong> the trajectory described by the state <strong>of</strong> the exiting<br />

point <strong>of</strong> a malfunctioning component when a malfunction occurs. For example, in<br />

Figure 2.3, an intrinsic malfunction in a turbine has been represented in the h-s plane,<br />

where the state <strong>of</strong> the exiting flow moves from 2 to 2’. The value <strong>of</strong> Td is determined by<br />

the behaviour <strong>of</strong> the system interacting with the malfunctioning component. So that, it<br />

can have two main origins: i) environmental condition, such as ambient pressure in the<br />

discharge <strong>of</strong> a gas turbine and ii) control system, such as temperature <strong>of</strong> the gas leaving<br />

a combustion chamber.<br />

Figure 2.3: Dissipation temperature.<br />

By using the concept <strong>of</strong> dissipation temperature, information about the behaviour <strong>of</strong><br />

a thermal system when a malfunction occurs can be introduced in the <strong>analysis</strong>, which<br />

allow to develop more accurate <strong>diagnosis</strong> methodologies. Zaleta et. al. (2003) have<br />

related the variation <strong>of</strong> the specific exergy consumption <strong>of</strong> a turbine in a steam cycle to<br />

the variation <strong>of</strong> the specific exergy consumption that it induced in the whole cycle. The<br />

following equation shows this relation for a low pressure turbine:<br />

⎛ ⎞<br />

⎛P⎞ ⎜<br />

1<br />

⎟<br />

dK ≅ ⎜ ⎟⎜ ⎟dk<br />

⎜ ⎟<br />

bT ,<br />

LP<br />

⎝ PT<br />

⎠ T0<br />

⎜<br />

1−kLP<br />

−<br />

TdLP<br />

,<br />

⎟<br />

LP<br />

h<br />

⎝ ⎠<br />

1<br />

2<br />

2' atn Td<br />

2.53<br />

2.54<br />

where Kb,T is the unit exergy consumption <strong>of</strong> the cycle, PLP and PT are the power<br />

produced by the low pressure turbine and the whole cycle, kLP is the unit exergy<br />

s<br />

21


Chapter 2<br />

consumption <strong>of</strong> the low pressure turbine, T0 is the reference temperature and Td,LP is the<br />

dissipation temperature <strong>of</strong> the low turbine pressure.<br />

These expressions have been applied to the <strong>diagnosis</strong> <strong>of</strong> a conventional steam cycle<br />

power plant <strong>of</strong> 158 MW (Zaleta et al. 2003). When they are compared with a<br />

simulation, they provided very good results. In the same paper, the fuel impact formula<br />

and a <strong>diagnosis</strong> system proposed by ASME (1996) have also been tested. Their results<br />

are not so good.<br />

The methodology proposed by ASME is based on the work developed by Cotton<br />

and Wescott (1960) and related the variation in isoentropic efficiency <strong>of</strong> a turbine to the<br />

impact in heat rate that it produces. For example, the heat rate impact produced by an<br />

anomaly in the low pressure steam turbine can be calculated as follows:<br />

22<br />

⎛ UEhp ⋅ m<br />

hp ⎞<br />

Δ HR%<br />

=Δηlp % ⎜1− ⎟<br />

2.55<br />

⎝ W<br />

Gen ⎠<br />

where ΔHR% and Δη% are the percentage variation <strong>of</strong> heat rate and isoentropic<br />

efficiency, UE is the used energy (enthalpy drop), m is the mass flow and W is the<br />

power produced by the cycle. Expressions for high and medium pressure steam turbines<br />

are more complex. This comparison should be made at constant throttle flow.<br />

2.2.4 Methods directly based on the thermodynamic representation<br />

<strong>of</strong> the system<br />

As it have been seen, <strong>methods</strong> based on exergetic and thermoeconomic indicators<br />

are general but <strong>of</strong>ten need the use <strong>of</strong> not only some kind <strong>of</strong> filtration but also the<br />

intervention <strong>of</strong> an analyst that may introduce subjectivity. Other <strong>methods</strong> like the ones<br />

proposed in the previous section provide quite good results but can only be used to<br />

diagnose a restricted type <strong>of</strong> components (e.g. turbines). In this context, the<br />

development <strong>of</strong> <strong>diagnosis</strong> systems that can be applied to real-life power plants need<br />

other <strong>methods</strong> that should be general (capable <strong>of</strong> diagnose all types <strong>of</strong> malfunctions)<br />

and reliable (the solution should not be probabilistic and should not need the<br />

intervention <strong>of</strong> an analyst). To achieve this, some authors consider that thermoeconomic<br />

tools are not the only solution and prefer to develop <strong>diagnosis</strong> techniques directly based<br />

in the thermodynamic description <strong>of</strong> the system.<br />

Zaleta et al. (2004a) propose a reconciliation in heat rate and power methodology<br />

based in the use <strong>of</strong> a simulator. Variables needed to describe thermodynamically the<br />

model can be classified in either dependent or independent. Dependent variables are


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

external (ambient conditions and fuel quality) or internal (indicators <strong>of</strong> equipment<br />

efficiency and set points <strong>of</strong> the control system). A simulator is able to obtain all the<br />

variables (the complete description <strong>of</strong> the system) only from the value <strong>of</strong> the<br />

independent ones. The procedure consists in calculating the test operation condition <strong>of</strong><br />

the plant and moving towards the reference operation condition by modifying the value<br />

<strong>of</strong> one independent variable in each step until all independent variables have the value<br />

corresponding to the reference condition. The increment in heat rate and power obtained<br />

in each step correspond to the impact due to the variable modified in this step. Authors<br />

remark that the method is not based in any linearization. On the other hand, a simulator<br />

is needed and results might vary depending on the order <strong>of</strong> variable modification. The<br />

method has been successfully applied to the TADEUS problem and it is the core <strong>of</strong> the<br />

<strong>diagnosis</strong> systems installed in more than eight combined cycle power plants in Mexico.<br />

Remiro and Lozano (2007) consider that fuel impact formula and the use <strong>of</strong> a<br />

simulator are complementary <strong>methods</strong> and propose to use both <strong>of</strong> them.<br />

Correas (2001, 2004) proposes a <strong>diagnosis</strong> algorithm also aimed at sharing the<br />

variation in efficiency among the set <strong>of</strong> thermodynamic independent variables (“free<br />

<strong>diagnosis</strong> variables”). The main advantage <strong>of</strong> this method is that it does not need a<br />

tuned simulator. It has been applied to solve the TADEUS problem (Correas, 2004) and<br />

is used in a <strong>diagnosis</strong> system at Elcogas IGCC power plant in Puertollano, Spain<br />

(García-Peña et al., 2000; García-Peña et al., 2001, and Correas 2001). This<br />

methodology is applied and improved in this thesis, so that it is explained in detail in the<br />

next section.<br />

2.2.5. Concept <strong>of</strong> the <strong>diagnosis</strong> algorithm and mathemathical<br />

formulation.<br />

The aim <strong>of</strong> this section is to describe the <strong>diagnosis</strong> algorithm proposed and applied<br />

by Correas (2001, 2004), which is the origin <strong>of</strong> the quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

developed in this thesis.<br />

The algorithm is based on a thermodynamic description <strong>of</strong> the energy system. The<br />

plant is represented as a graph, whose nodes are components where processes take<br />

place. The edges linking these nodes are material or energy streams. Every stream is<br />

characterized by using an extensive property (flow rate or power) and several intensive<br />

properties such as pressure, temperature or quality. The components are characterized<br />

by using parameters like efficiencies, pressure losses and so on.<br />

23


Chapter 2<br />

The value <strong>of</strong> these variables has been previously determined by another calculation<br />

method different from the <strong>diagnosis</strong> for the two states to be compared by the <strong>diagnosis</strong><br />

methodology. Usually, one <strong>of</strong> these states is the reference one. This previous calculation<br />

method can be a simulator or a “performance test code” that solves the mass and energy<br />

balances from field instrumentation.<br />

All the variables (properties <strong>of</strong> streams and parameters <strong>of</strong> components) can be<br />

classified as dependent and independent. Usually, the independent variables are<br />

environmental conditions, fuel quality, set points and component parameters. There is<br />

also an indicator <strong>of</strong> the global efficiency <strong>of</strong> the whole system. The aim <strong>of</strong> the algorithm<br />

is to determine the influence <strong>of</strong> the variations <strong>of</strong> the independent variables in the<br />

variation <strong>of</strong> the global efficiency indicator.<br />

Let’s consider a plant that is represented by a set <strong>of</strong> variables x. It is also a global<br />

efficiency indicator that, obviously, depends <strong>of</strong> this set. The variation <strong>of</strong> this indicator<br />

between two plant states 0 and 1 is expressed by:<br />

24<br />

1 0 ( ) ( )<br />

Δ f = f x − f x 2.56<br />

The aim <strong>of</strong> the <strong>diagnosis</strong> algorithm is to share this efficiency variation in a<br />

summation <strong>of</strong> impacts, each one due to the variation <strong>of</strong> an independent variable:<br />

n n<br />

indep indep<br />

0 1 * 0 1<br />

∑ ( , , ) ∑ cos , ( , )<br />

Δf≅ I Δ x x x = k x x ⋅Δx<br />

2.57<br />

f i t i i<br />

i= 1 i=<br />

1<br />

So that, the purpose <strong>of</strong> the algorithm is to obtain the set <strong>of</strong> impact factors<br />

k that<br />

relate the variation <strong>of</strong> the indicator to the variation to the independent variables. This<br />

task is not direct because the indicator depends not only on the values <strong>of</strong> the<br />

independent variables but also on the dependent ones:<br />

( ) = ( dep, indep )<br />

*<br />

cos t,<br />

i<br />

f x f x x 2.58<br />

The variation <strong>of</strong> the indicator can be related to the variation <strong>of</strong> x by using a Taylor<br />

series:<br />

n<br />

n<br />

n<br />

var<br />

indep<br />

dep<br />

∂f<br />

∂f<br />

∂f<br />

Δf ≅ ∑ ⋅ Δxi<br />

= ∑ ⋅ Δxindep,<br />

i + ∑<br />

i= 1 ∂xi<br />

i= 1 ∂xindep,<br />

i<br />

i= 1 ∂xdep,<br />

i<br />

⋅ Δx<br />

*<br />

This relation can be expressed in matrix notation by defining k f i = ∂f<br />

/ ∂xi<br />

:<br />

* T * T ⎡Δxindep ⎤<br />

Δf≅kf ⋅Δ x = kf<br />

⋅⎢<br />

Δ<br />

⎥<br />

⎣ xdep<br />

⎦<br />

dep,<br />

i<br />

,<br />

2.59<br />

2.60


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

At this point, the variation <strong>of</strong> f has been related to the variation <strong>of</strong> all the variables <strong>of</strong><br />

the system. To relate f to the independent variables only, the ndep restrictions (g(x) = 0)<br />

<strong>of</strong> the problem that relate dependent and independent variables should be introduced. If<br />

the function g is expanded in a Taylor series and terms <strong>of</strong> higher order are neglected, it<br />

can be written:<br />

1 0 ( ) ( ) ( ) 0<br />

g x ≅ g x + J x ⋅Δx<br />

2.61<br />

x<br />

Both vectors x 1 and x 0 satisfy the set <strong>of</strong> constraints because they are solutions <strong>of</strong><br />

g(x) = 0. So that, the previous equation can be simplified to:<br />

J( x) 0 ⋅Δx ≅0<br />

2.62<br />

x<br />

The size <strong>of</strong> the previous equation system is (ndep x nvar), so that it must be completed<br />

to be solved:<br />

( ) 0 xindep<br />

⎡J x<br />

⎢<br />

⎢⎣ U D<br />

⎤ ⎡Δ x<br />

⎥⋅⎢ 0 ⎥ ⎣ Δxdep ⎦<br />

⎤ ⎡ 0 ⎤<br />

⎥ ≅ ⎢<br />

Δ<br />

⎥<br />

⎦ ⎣ xindep<br />

⎦<br />

2.63<br />

If the matrix is inverted, the relationship between dependent and independent<br />

variables is obtained:<br />

⎡Δxindep ⎤ ⎡ 0<br />

⎢ ⎥ ≅ ⎢<br />

⎣ Δxdep⎦ ⎣M U ⎤ ⎡ 0 ⎤<br />

⎥⋅⎢<br />

Δ<br />

⎥<br />

N⎦<br />

⎣ xindep<br />

⎦<br />

2.64<br />

Δxdep ≅ N⋅Δx indep<br />

2.65<br />

Once this relation is known, the relation <strong>of</strong> variation <strong>of</strong> the indicator to variation <strong>of</strong><br />

independent variables is determined.<br />

The method presented above is approximate because only first order terms in the<br />

series expansion have been considered. It would be exact if both the variation in the<br />

indicator as well as the constraints were linear. The error induced by linearization is<br />

studied in this thesis.<br />

2.2.6. Other examples.<br />

Rusinowski et al. (2007) have developed a <strong>diagnosis</strong> method <strong>of</strong> a boiler based on a<br />

neural network which links temperature <strong>of</strong> gases leaving the boiler with seven<br />

independent variables.<br />

A method based on neural networks has also been used by Gluch et al. (1998). First<br />

<strong>of</strong> all, symptoms <strong>of</strong> failure (difference between measured and reference parameters) are<br />

detected. Afterwards, a two step approach is applied: first <strong>of</strong> all, a module for searching<br />

25


Chapter 2<br />

defected apparatuses; second, a set <strong>of</strong> modules to diagnose the causes <strong>of</strong> degradation <strong>of</strong><br />

individual apparatuses.<br />

Sciubba (2004) proposes a hybrid system based on both qualitative and quantitative<br />

approaches. It combines reconciled data and physical modelling to extract a limited<br />

number <strong>of</strong> numerical coefficients that introduce a sufficient degree <strong>of</strong> quantification.<br />

The potentialities <strong>of</strong> a qualitative method for the <strong>analysis</strong> <strong>of</strong> system operating<br />

condition by means <strong>of</strong> an expert system with fuzzy rules have been explored by T<strong>of</strong>folo<br />

and Lazzaretto (2007). The knowledge base <strong>of</strong> the expert system is obtained by a given<br />

list <strong>of</strong> malfunctioning conditions. A multi-objective evolutionary algorithm is used in<br />

order to accomplish two conflicting objectives: the number <strong>of</strong> correct predictions (to be<br />

maximized) and the complexity <strong>of</strong> the set <strong>of</strong> rules (to be minimized). Two approaches<br />

are considered: global approach for the overall system and local approach by splitting<br />

the system into subsystems.<br />

El-Sayed (2007) proposes to analyze the fingerprint <strong>of</strong> malfunctions. The idea is to<br />

use a simulator to obtain fingerprint tables which show the variation <strong>of</strong> thermal system<br />

parameters when a malfunction in a component appears; the comparison <strong>of</strong> the<br />

operating situation with the different fingerprints allows to determine the actual<br />

malfunctioning component.<br />

26<br />

2.3. State <strong>of</strong> the art <strong>of</strong> systems for <strong>diagnosis</strong> and other related<br />

issues.<br />

After the revision <strong>of</strong> the methodologies for thermoeconomic <strong>diagnosis</strong>, the aim <strong>of</strong><br />

this section is to present the systems developed to perform this task. Obviously, a<br />

<strong>diagnosis</strong> system has to be based on a <strong>diagnosis</strong> method, so that, this section is strongly<br />

linked to the previous one; in other words, the separation is mainly due to the<br />

organization <strong>of</strong> the text and should not be considered as a strict rule.<br />

Diagnosis systems are seldom <strong>of</strong>fered as independent items, but they are usually<br />

part <strong>of</strong> packages covering other issues related to plant management which use<br />

information available: performance monitoring, mechanical condition monitoring, online<br />

optimization (set-points, sootblowing, compressor washing in gas turbines…). In<br />

this framework, the aim <strong>of</strong> this section is to review the <strong>diagnosis</strong> systems and <strong>methods</strong><br />

commercially available and to show how the thermoeconomic <strong>diagnosis</strong> task is a part <strong>of</strong>


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

the total plant management. In order to organize the revision, information is classified<br />

into several sub-sections: <strong>diagnosis</strong> systems in general, <strong>diagnosis</strong> centres and<br />

collaboration, <strong>diagnosis</strong> <strong>of</strong> boilers, and other issues. Again, this classification should not<br />

be seen as a strict rule.<br />

2.3.1. Diagnosis systems in general.<br />

Plant equipment manufacturers usually develop the plant control system and are the<br />

people who best known the plant characteristics. In this situation, they have developed<br />

monitoring and <strong>diagnosis</strong> systems which are sold in the same package <strong>of</strong> the plant or,<br />

sometimes, can be added afterwards. It is usual that these systems have a modular<br />

structure based on a core system which can be optionally complemented with several<br />

packages with different aims.<br />

GE Energy (2005a) has developed the System 1® Optimization and Diagnostic<br />

Platform in order to organize and distribute plant information. This system can include<br />

the ‘Machine Performance. Condition Monitoring and Diagnosis’ package (GE Energy,<br />

2005b) which can monitor thermodynamic monitoring <strong>of</strong> gas and steam turbines and<br />

allow the user to combine this information with mechanical condition monitoring, in<br />

order to take decisions on maintenance. Modules for modelling and detection <strong>of</strong><br />

inconsistent measurements are also included.<br />

Alstom have developed the ALSPA P320 flexible automation system. It can include<br />

an integrated plant management system OPTIPLANT (Alstom, 2007).<br />

ABB provides the Optimax suite, which includes several packages for plant<br />

monitoring and optimization: maintenance and lifecycle management, optimization <strong>of</strong><br />

power and desalination plants or optimization <strong>of</strong> compressor washing (ABB, 2005a-d;<br />

Morton, 2002). According this company, plant operation optimization needs: process<br />

models to determine fuel consumption, stress models to compute lifetime cost,<br />

operation constraints and an optimisation routine (Weisenstein et al., 2002). Finally,<br />

Antoine (2005) describes how parameters indicative <strong>of</strong> gas turbine degradation can be<br />

estimated from measurements, applies the idea <strong>of</strong> probability <strong>of</strong> failure and presents a<br />

compressor washing optimiser.<br />

Siemens <strong>of</strong>fers the SPPA-P3000 system process optimization which includes<br />

packages for optimizing different items such as: boiler start-up, optimum set-points,<br />

sootblowing, combustion… (Siemens, 2006a-h).<br />

27


Chapter 2<br />

Besides these proposals <strong>of</strong> plant manufacturers, other companies have developed<br />

programs aimed at monitoring and diagnose power plants. SOLCEP is a s<strong>of</strong>tware<br />

implemented in power plants <strong>of</strong> Endesa (Pablo Moreno S.A., 2006). This s<strong>of</strong>tware is<br />

based on a modular structure easy to configure by the plant staff. It not only calculates<br />

the plant heat rate, but also compares it with the expected one, sharing this deviation<br />

first into sub-systems (boiler, steam cycle and ancillary devices) and then into main<br />

components. It has other features such as data validation, historian retrieval and<br />

distribution <strong>of</strong> information among users according to different levels.<br />

An on-line supervision system <strong>of</strong> energetic efficiency has been installed in<br />

Compostilla power station, a coal-fired power plant owned by Endesa (Gómez-Yagüe et<br />

al., 1999). This system is able to calculate global and individual efficiencies and<br />

evaluate their deviations with respect to current plant objectives and to dynamic<br />

references established by the operator.<br />

DADIC is a system to diagnose combined cycle power plants (Arranz et al., 2007).<br />

The <strong>diagnosis</strong> is performed by using a multi-agent system: the detection is based on<br />

agents that use neural-networks based models, and the <strong>diagnosis</strong> is performed by agents<br />

based on an expert-system structure.<br />

Kim and Joo (2005) describe a <strong>diagnosis</strong> system based on EfficiencyMap from GE<br />

Energy applied in two Korean combined cycles. Corrected gas turbine power is<br />

calculated by a simulator, and the comparison <strong>of</strong> this parameter with the actual power is<br />

an indicator <strong>of</strong> turbine degradation. Compressor wash advisor and a module to<br />

optimally allocate load among the different turbines <strong>of</strong> the two plants are also included.<br />

Barbucci et al. (2002) present a system for monitoring and <strong>diagnosis</strong> <strong>of</strong> two gas<br />

turbines owned by Enel. The system features includes: gas path <strong>analysis</strong> to determine<br />

the thermodynamic parameters <strong>of</strong> turbine components, a neural network based system<br />

to analyze the pattern <strong>of</strong> flue gas temperatures, vibration monitoring and emissions and<br />

stability <strong>analysis</strong> <strong>of</strong> the combustion system. A system to diagnose a gas turbines from<br />

Ansando Energia based on the use <strong>of</strong> a tuned simulator is presented by Crosa and<br />

Torelli (2005).<br />

Lopez et al. (1999) describe an expert system based on decision trees and fuzzy<br />

logic to determine causes <strong>of</strong> power losses in C<strong>of</strong>rentes Nuclear Power Plant (Spain).<br />

The problem <strong>of</strong> lost electric output in nuclear power plants has also been tackled by Heo<br />

et al. (2005) and Heo and Chang (2005), which propose a system based on statistical<br />

regression whose coefficients have been determined by using a simulator.<br />

28


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

2.3.2. Diagnosis centres and manufacturer-operator collaboration.<br />

Manufacturers <strong>of</strong> power plant equipment sometimes propose collaboration<br />

agreements with their customers in order to sell not only the product and sometimes its<br />

maintenance but also a monitoring and <strong>diagnosis</strong> service which can be performed<br />

remotely in one or more centres by using information technologies widespread available<br />

today. Besides, it should be noted that, although it is not usually acknowledged<br />

explicitly, manufacturers are also interested in knowing the true behaviour <strong>of</strong> their<br />

products in service. Examples <strong>of</strong> these centres are the Ansaldo Diagnosis Centre<br />

(Rebizzo et al., 2002; Rebizzo et al., 2003), the diagnostic centres <strong>of</strong> Siemens in<br />

Orlando and Erlangen (Scheidel et al., 2004; Bauch and Killich, 2005), and the Alstom<br />

centre (Decoussemaeker et al., 2005). Collaboration <strong>of</strong> manufacturers and users is also<br />

proposed by the engine manufacturer Wärtsilä (Klimnstra, 2004). The idea <strong>of</strong> common<br />

<strong>diagnosis</strong> centres for several power plants has also been applied by utilities such as<br />

Iberdrola (Mendivil and Álvarez, 2002) and RWE (Grimwade and Gornm, 2005).<br />

2.3.3. Monitoring, <strong>diagnosis</strong> and optimization <strong>of</strong> boilers.<br />

Coal-fired boilers are complex systems whose operation can be improved in several<br />

ways: combustion in burners, sootblowing optimization, fuel or start-up.<br />

Several tools have been developed to optimize operation <strong>of</strong> burners in coal boilers:<br />

suitable amount <strong>of</strong> primary and secondary air, choice <strong>of</strong> burners used… They are<br />

proposed as affordable tools to reduce NOx while boiler efficiency is improved. The use<br />

<strong>of</strong> additional instrumentation is usually needed in order to determine coal and air mass<br />

flows to burners. Sometimes, systems to measure carbon in ash are also used. Suitable<br />

models (<strong>of</strong>ten empirical, built by tools such as neural networks) are developed capable<br />

to predict the consequence <strong>of</strong> modification <strong>of</strong> operation. Usually, an optimization<br />

module is used to decide the optimum operation, and sometimes the use <strong>of</strong> closed-loop<br />

operation is possible. Examples <strong>of</strong> these systems are: Jarc and Lang, 1998; Otero et al.,<br />

1999; Ray and Howard, 1999; Copado and Rodríguez, 2002; Thulen and Peper, 2003;<br />

Reismann et al., 2005; Siemens 2006a.<br />

Sootblowing <strong>of</strong> heat transfer surfaces is a key point to achieve high efficiency and<br />

operation without problems <strong>of</strong> coal boilers. However, this operation needs the use <strong>of</strong><br />

steam, which implies an additional consumption <strong>of</strong> coal. In this context, there is a big<br />

interest in substituting the traditional sootblowing programming at regular time intervals<br />

by intelligent systems able to determine the reduction in the heat transfer in the different<br />

29


Chapter 2<br />

surfaces <strong>of</strong> the boiler, to predict its future evolution and to use this information to<br />

decide the optimum moment for sootblowing. This is a complex task because both<br />

complex monitoring techniques and models <strong>of</strong> fouling and slagging are needed: Cortés<br />

et al., 1989; Cortés et al., 1993; Bella et al., 1994; Valero and Cortés, 1996; Bartels et<br />

al., 2003; Teruel et al., 2005; Siemens, 2006b; Romeo and Gareta, 2006.<br />

A detailed monitoring <strong>of</strong> boilers is a difficult task due to the complexity <strong>of</strong><br />

combustion and heat transfer processes and the lack <strong>of</strong> instrumentation. Cortés et al.<br />

(1998) and Díez et al. (1998, 2001) have studied this task.<br />

A computer system including sootblowing optimization, combustion set-points<br />

optimization, and on-line monitoring <strong>of</strong> heat transfer and air-coal system has been<br />

developed to the boiler <strong>of</strong> the power plant studied in this thesis (Carrasco et al., 2001).<br />

Boilers have been designed to burn a determined type <strong>of</strong> fuel, so that, its<br />

modification has to be carefully analyzed. An application <strong>of</strong> a program to perform this<br />

task is described by Gálvez et al. (2004). Systems for adapting burners operation to coal<br />

properties have also been developed (Siemens, 2006c).<br />

Boiler start-up is an operation which has to be done carefully and progressively in<br />

order to avoid excess <strong>of</strong> thermal stress which can produce damage. However, the<br />

necessity to adapt the operation to competitive markets and to reduce fuel consumption<br />

recommends to perform this operation in a time span as short as possible. Several<br />

systems have been developed to conjugate these contradictory requirements (ABB,<br />

2005a; Siemens 2006d).<br />

One important cause <strong>of</strong> reduction in boilers’ efficiency is air leakages in preheaters.<br />

Drobnic et al. (2006) propose to use a numerical model to determine sealing failure by<br />

using information from air and gases temperatures.<br />

2.3.4. Other issues.<br />

Ioanilli et al. (1998) describe a method to detect the state <strong>of</strong> feedwater heaters in<br />

order to optimise their maintenance. The goal is reduce the additional fuel consumption<br />

caused by tube plugging or heater’s by-pass while keeping low the associated<br />

maintenance operations.<br />

GE Energy has developed a system for closed-loop optimal control <strong>of</strong> set-points in<br />

combined cycle and co-generation plants (Davari and de Boer, 2004).<br />

An indicator <strong>of</strong> mechanical failure in turbomachinery is the presence <strong>of</strong> abnormal<br />

vibrations. Although vibrations above a certain level should be always detected for<br />

30


State <strong>of</strong> the art <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

safety reasons, the detection <strong>of</strong> anomalies in an early stage is very convenient in order to<br />

avoid further damage <strong>of</strong> the turbine and additional fuel consumption. (Herbert and<br />

Davies, 2005; Parenti et al., 1998).<br />

Malfunctions in the combustion chamber <strong>of</strong> gas turbines can lead not only to<br />

additional fuel consumption but also to damages in the turbine. Verhage et al. (1999)<br />

present a system to detect instabilities in combustion, while Siemens (2007) has<br />

developed a system to optimize the operation <strong>of</strong> this component.<br />

The <strong>analysis</strong> <strong>of</strong> previous examples <strong>of</strong> component degradation and their symptoms<br />

can serve as valuable experience to diagnose new failures. Tirone et al. (1998) present<br />

some examples <strong>of</strong> steam turbines owned by Enel, and Kubiak et al. (2002) analyse<br />

degradation <strong>of</strong> gas turbine components.<br />

2.4. Conclusion<br />

The main concepts and tools provides by Thermoeconomics for <strong>diagnosis</strong> have been<br />

presented, as well as other techniques to filter induced effects. Proposals which avoid<br />

the use <strong>of</strong> Thermoeconomics have also been reviewed, especially the <strong>diagnosis</strong><br />

algorithm which is further developed in next chapter. Finally, systems for monitoring<br />

and diagnose thermal systems have been presented. This allows to see the capabilities<br />

and requirements <strong>of</strong> industrial applications as well as other related issues which can<br />

result in synergies (e.g. mechanical condition monitoring, on-line optimization…).<br />

31


Chapter 2<br />

32


3. Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

In this chapter the <strong>diagnosis</strong> methodology applied in this thesis is presented. This<br />

method is based on the <strong>diagnosis</strong> algorithm proposed by Correas (2001), which is the<br />

core <strong>of</strong> a <strong>diagnosis</strong> system applied to an IGCC power plant (Correas 2001, García-Peña<br />

et. al., 2000 and 2001)and has also been successfully applied to the TADEUS problem<br />

(Correas, 2004). It has been presented in section 2.2.5.<br />

The idea <strong>of</strong> the method is to share the variation <strong>of</strong> an indicator <strong>of</strong> the efficiency <strong>of</strong><br />

the whole system (e.g. efficiency, heat rate…) in the summation <strong>of</strong> several impact terms<br />

each one due to the variation <strong>of</strong> an independent variable. These independent <strong>diagnosis</strong><br />

variables are called free <strong>diagnosis</strong> variables and can be ambient conditions, fuel quality<br />

parameters, set-points or indicators <strong>of</strong> the state <strong>of</strong> the components <strong>of</strong> the energy system.<br />

The original <strong>diagnosis</strong> algorithm is improved by developing several points. First, the<br />

nomenclature is adapted in order to easily introduce the improvements. Then, the<br />

methodology is extended to include the influence <strong>of</strong> measurement errors. Afterwards,<br />

the problem <strong>of</strong> non-linearities is analyzed. Finally, the concept <strong>of</strong> <strong>causal</strong>ity chains is<br />

introduced. A very simple system is used as an example to explain the concepts.<br />

3.1. Re-formulation <strong>of</strong> the <strong>diagnosis</strong> problem<br />

In this paragraph, a new nomenclature for the <strong>diagnosis</strong> problem variables is<br />

introduced in order to clarify the notation. This task is important because new problems,<br />

such as the influence <strong>of</strong> error measurement, are going to be considered.<br />

A thermal system described by a set <strong>of</strong> nt thermodynamic variables (x) is considered.<br />

nd <strong>of</strong> these variables are the free <strong>diagnosis</strong> variables. There are a set <strong>of</strong> nr restrictions<br />

33


Chapter 3<br />

that link the variables. These restrictions are matter and energy balances and definitions<br />

<strong>of</strong> parameters <strong>of</strong> equipments. So that, they are general restrictions that are common to<br />

the problems <strong>of</strong> <strong>diagnosis</strong>, performance test, simulation and optimization.<br />

R( x ) = 0<br />

3.1<br />

34<br />

There may be also other specific restrictions for the <strong>diagnosis</strong> problem:<br />

RD( x ) = 0<br />

3.2<br />

The number <strong>of</strong> these restrictions nrd should accomplish the following relation:<br />

nrd = nt−nr− nd<br />

3.3<br />

For small deviations, both restrictions can be expanded in a first order Taylor series:<br />

1 0 ( ) ( ) ( ) 0<br />

R x ≅ R x + JR x ⋅Δx<br />

3.4<br />

1 0<br />

( ) ( ) ( ) 0<br />

x<br />

RD x ≅ RD x + JRD x ⋅Δx<br />

3.5<br />

x<br />

Where JR and JRD are the Jacobian matrices corresponding to restrictions R and<br />

1 0<br />

RD and Δ x = x − x . Since both x 0 and x 1 are solutions <strong>of</strong> the systems, they accomplish<br />

the restrictions. So that:<br />

⎡ JR ⎤<br />

⎢ ⎥⋅Δx<br />

≅0<br />

3.6<br />

⎣JRD⎦ To obtain a resoluble system, the variation <strong>of</strong> the free <strong>diagnosis</strong> variables should be<br />

introduced.<br />

⎡ JR ⎤<br />

⎢ ⎥ ⎡ 0 ⎤<br />

⎢<br />

JRD<br />

⎥<br />

⋅Δ x = JD ⋅Δx ≅ ⎢<br />

Δ<br />

⎥<br />

⎣ xd<br />

⎢ ⎥<br />

⎦<br />

⎣ VD ⎦<br />

Where JD is a nt x nt matrix including JR, JRD and VD. VD is a nd x nt matrix<br />

defined as:<br />

VDij = 1 if the j th element <strong>of</strong> xt is the ith element <strong>of</strong> xd.<br />

VDij = 0 in other case.<br />

The use <strong>of</strong> the matrix VD instead <strong>of</strong> an identity and a null matrices, makes<br />

unnecessary that the vector x is ordered. This is very important when the same vector x<br />

is used for other problems different from the <strong>diagnosis</strong>. If x were ordered, VD would be<br />

composed <strong>of</strong> an identity and a null matrices. The VD matrix allows to obtain the xd<br />

vector from the x vector:<br />

d = ⋅ x VD x d<br />

3.7<br />

Δ x = VD⋅Δx 3.8<br />

By inverting the JD matrix, the variation <strong>of</strong> x is related to the variation <strong>of</strong> xd.


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

−1 ⎡ 0 ⎤ −1<br />

⎡ 0 ⎤<br />

Δx ≅ JD ⋅ ⎢ ⎥ = JD ⋅⎢ ⎥⋅Δxd<br />

3.9<br />

⎣Δxd⎦ ⎣UD⎦ An indicator <strong>of</strong> the global efficiency <strong>of</strong> the system e(x) is considered. Its variation<br />

can be expressed by using a Taylor series again:<br />

∂e<br />

∑ ex x 3.10<br />

n<br />

t<br />

Δe≅ ⋅Δ xi=<br />

⋅Δ<br />

i= 1 ∂xi<br />

Finally, variation <strong>of</strong> e can be related to the variation <strong>of</strong> the free <strong>diagnosis</strong> variables:<br />

t −1 ⎡0⎤ t<br />

Δe≅ex ⋅JD ⋅⎢ ⋅Δ d = ⋅Δ d<br />

U<br />

⎥ x ed x<br />

⎣ ⎦<br />

Where ed is the vector containing the sensitivity <strong>of</strong> the efficiency indicator e respect<br />

to the free <strong>diagnosis</strong> variables. For convenience for further developments, it is<br />

interesting to express ed t as a product <strong>of</strong> an unit vector times a diagonal matrix ED:<br />

t<br />

Δe≅ ⋅ ⋅Δ d<br />

u ED x 3.11<br />

3.1.1. Example<br />

To clarify the concepts and nomenclature presented above, an example consisting <strong>of</strong><br />

a very simple steam cycle is proposed. It is going to be used to explain other ideas in<br />

this chapter.<br />

2<br />

1<br />

6<br />

Wb<br />

Qc<br />

Figure 3.1. Scheme <strong>of</strong> the example cycle.<br />

5<br />

3<br />

4<br />

Wt<br />

35


Chapter 3<br />

The steam cycle is represented in Figure 3.1. It is composed <strong>of</strong> a pump, a boiler, a<br />

turbine and a condenser. There are two flows: the flow describing the water-steam cycle<br />

that comprises streams 1 to 4 and the cooling water flow that includes streams 5 and 6.<br />

The properties <strong>of</strong> the streams are summarized in Table 3.1.<br />

POINT m (kg/s) p (bar) T (ºC) x (-) h (kJ/kg) s<br />

(kJ/kg·K)<br />

36<br />

1 97.882 0.07 39.0 0 163.4 0.559<br />

2 97.882 40 39.3 168.1 0.561<br />

3 97.882 35 540.0 3541.8 7.272<br />

4 97.882 0.07 39.0 0.9767 2515.4 8.094<br />

5 10000 1 20 83.9 0.296<br />

6 10000 1 25.5 106.9 0.374<br />

Table 3.1. Properties <strong>of</strong> streams in the cycle.<br />

Other important values related to the example cycle are: heat provided by the boiler,<br />

heat rejected in the condenser, power produced in the turbine, power consumed by the<br />

pump, net power produced, cycle efficiency, condenser effectiveness and isoentropic<br />

efficiencies <strong>of</strong> pump and turbine:<br />

Q boiler = 330,220 kW<br />

Q condenser = 230,220 kW<br />

W turbine = 100,463 kW<br />

W pump = 463 kW<br />

W net = 100,000 kW<br />

ηcycle = 0.303<br />

εcondenser = 0.29<br />

ηpump = 0.85 tpu<br />

ηturbine = 0.8 tpu<br />

To represent the cycle, 18 variables are considered (nt = 18): pressure and<br />

temperature <strong>of</strong> points 2 and 3, condensing pressure and temperature, quality <strong>of</strong> points 1<br />

and 4, temperatures <strong>of</strong> points 5 and 6 (cooling water is considered as incompressible<br />

liquid with constant specific heat), two mass flows (steam and cooling water), two


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

isoentropic efficiencies (pump and turbine), one pressure drop (in the boiler), condenser<br />

efficiency, the power and cycle efficiency. It should be noted that the cycle could be<br />

represented by using more variables, e. g. six mass flows instead <strong>of</strong> only two. In this<br />

case, four mass balances should be introduced in the set <strong>of</strong> restrictions. So that, the x<br />

vector containing the n variables necessary to describe the system is:<br />

x<br />

t<br />

( psat p2 p3 TsatT2 T3 T5 T6 x1 x4 m1 m5 W n εcηcηpηtpb) = Δ<br />

3.12<br />

These 18 variables are linked by 9 restrictions:<br />

• Net power calculation<br />

W−m ⋅ h p , x − h p , T + h p , T − h p , x = 0 3.13<br />

( ( ) ( ) ( ) ( ) )<br />

n 1 1 sat 1 2 2 2 3 3 3 4 sat 4<br />

• Condenser energy balance<br />

( ) ( ( ) ( ) )<br />

m ⋅c ⋅ T −T −m ⋅ h p , x − h p , x = 0<br />

3.14<br />

5 pw , 6 5 1 4 sat 4 1 sat 1<br />

• Pressure drop in boiler<br />

Δp − p + p = 3.15<br />

b<br />

2 3 0<br />

• Condenser effectiveness<br />

( )<br />

ε ⋅ − − + = 3.16<br />

c T5 Tsat Tsat T5<br />

0<br />

• Cycle efficiency<br />

( ( ) ( ) )<br />

η ⋅m ⋅ h p , T − h p , T + W=<br />

0<br />

3.17<br />

c 1 3 3 3 2 2 2 n<br />

• Pump isoentropic efficiency<br />

( h2( p2 T2) h1( p x1) ) h2 ( p2 T2 p x1) h1( p x1)<br />

η ⋅ , − , − , , , + , = 0 3.18<br />

p sat s sat sat<br />

• Steam turbine isoentropic efficiency:<br />

( h3( p3 T3) h4 ( p x4 p3 T3) ) h3( p3 T3) h4( p x4)<br />

η ⋅ , − , , , − , + , = 0 3.19<br />

t s sat sat<br />

• Saturated liquid in point 1:<br />

x = 0<br />

3.20<br />

1<br />

• Saturation conditions:<br />

sat sat sat<br />

( ) 0<br />

T − T P = 3.21<br />

37


Chapter 3<br />

These restrictions are common to the problems <strong>of</strong> <strong>diagnosis</strong>, performance test,<br />

simulation and optimization; so that, they all corresponds to the equation R(x) = 0. By<br />

deriving the operator R(x) in the reference point, the JR matrix can be obtained:<br />

38<br />

⎡ − ⋅ −<br />

⎢<br />

⎢ 67940 0.000242 −0.00025 −0.000115 8.60⋅10 2.20⋅10 ⎢ 0 −1<br />

1 0 0 0<br />

⎢<br />

−5<br />

⎢−2.20⋅10<br />

0 0 0.6698 0 0<br />

−5<br />

JR = ⎢ −0.52336 −2.637 −28.757 −1.70⋅10 −123.26<br />

67.052<br />

⎢<br />

−6<br />

⎢ −1081.3 −0.024588 1.00⋅100 3.547 0<br />

⎢<br />

−1074.<br />

7 0 3.68 0 0 −1.1446<br />

⎢<br />

⎢ 0 0 0 0 0 0<br />

⎢ −9<br />

⎣ −304.01 −2.00⋅10 0 1 0 0<br />

1<br />

0<br />

0<br />

0<br />

−1<br />

0<br />

0<br />

0<br />

0<br />

−<br />

67942 8.6096 93.889<br />

−5<br />

5.60 10 402.43 218.92<br />

−5 −5<br />

0.<br />

000102<br />

− 20915<br />

0<br />

0.<br />

3302<br />

3.<br />

10 ⋅10<br />

1.<br />

00 ⋅10<br />

0<br />

0<br />

0<br />

−5<br />

−6<br />

0.<br />

003037<br />

− 0.<br />

012514<br />

0<br />

16.<br />

168<br />

− 0.<br />

00093<br />

2.<br />

70 ⋅10<br />

0<br />

0<br />

0<br />

−5<br />

6.<br />

60E<br />

− 5<br />

−<br />

20915<br />

0<br />

−1<br />

2.<br />

00 ⋅10<br />

1.<br />

00 ⋅10<br />

0<br />

0<br />

0<br />

0.<br />

006642<br />

0.<br />

00426<br />

0<br />

0<br />

3.<br />

26 ⋅10<br />

5.<br />

90 ⋅10<br />

0<br />

0<br />

0<br />

5<br />

−5<br />

−5<br />

−6<br />

−<br />

2.<br />

33⋅10<br />

2.<br />

33⋅10<br />

−<br />

0<br />

0<br />

0.<br />

000623<br />

−<br />

2055.<br />

8<br />

0<br />

1<br />

0<br />

0.<br />

002393<br />

0.<br />

018231<br />

0<br />

0<br />

− 0.<br />

000733<br />

4.<br />

7251<br />

0<br />

0<br />

0<br />

5<br />

5<br />

2.<br />

33⋅10<br />

−<br />

−<br />

2.<br />

33⋅10<br />

0<br />

0<br />

0.<br />

000641<br />

1.<br />

80 ⋅10<br />

2415<br />

0<br />

0<br />

0.<br />

002543<br />

− 0.<br />

013099<br />

0<br />

0<br />

− 0.<br />

000779<br />

2.<br />

20 ⋅10<br />

1302,<br />

1<br />

0<br />

0<br />

−5<br />

5<br />

5<br />

−5<br />

−1036.<br />

9<br />

−<br />

2348.<br />

6<br />

0<br />

0<br />

1036.<br />

9<br />

0<br />

0<br />

0<br />

0<br />

0.<br />

000407 ⎤<br />

0.<br />

00107<br />

⎥<br />

⎥<br />

1 ⎥<br />

⎥<br />

0 ⎥<br />

− 0.<br />

000125⎥<br />

4.<br />

00 ⋅10<br />

0<br />

0<br />

0<br />

−6<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎦<br />

0<br />

45.<br />

299<br />

9 free <strong>diagnosis</strong> variables have been chosen: the two isoentropic efficiencies,<br />

temperature and pressure at inlet <strong>of</strong> steam turbine, pressure drop in the boiler, condenser<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

3.22


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

efficiency, temperature and mass flow <strong>of</strong> the cooling water and net power. So that, the<br />

xd vector can be defined as:<br />

( p3 T3 T5 m5 W ε η η p )<br />

x = Δ<br />

3.23<br />

t<br />

d n c p t b<br />

And the VD matrix that connects the vectors xd and x:<br />

⎡0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎤<br />

⎢<br />

0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0<br />

⎥<br />

⎢ ⎥<br />

⎢0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0⎥<br />

⎢ ⎥<br />

⎢0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0⎥<br />

VD = ⎢0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0⎥<br />

3.24<br />

⎢ ⎥<br />

⎢0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0⎥<br />

⎢0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0⎥<br />

⎢ ⎥<br />

⎢0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0⎥<br />

⎢<br />

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1<br />

⎥<br />

⎣ ⎦<br />

There are 18 total variables, 9 restrictions corresponding to R(X) = 0 and 9<br />

<strong>diagnosis</strong> variables. So that, no additional conditions specific for the <strong>diagnosis</strong> problem<br />

are needed (the dimension <strong>of</strong> RD(x) = 0, nrd, is equal to zero) and the matrix JD is only<br />

formed by JR and VD.<br />

Finally, the indicator for the efficiency <strong>of</strong> the global system (e) belongs to the set <strong>of</strong><br />

variables x; so that, the vector ex has a very easy expression composed <strong>of</strong> only one 1<br />

and zeros:<br />

( 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0)<br />

t<br />

ex =<br />

3.25<br />

By using the matrices calculated above and the expressions obtained in the previous<br />

section, the vector that relates the variations <strong>of</strong> the free <strong>diagnosis</strong> variables (Δxd) to the<br />

variations <strong>of</strong> the efficiency indicator (Δe) can be obtained:<br />

ed<br />

t −6 −7<br />

= ( 0.0009514 0.00014003 −0.0013413 4.3373⋅10 −2.1687⋅10 0.<br />

032378<br />

0.<br />

0012553<br />

0.<br />

42385<br />

)<br />

−5<br />

− 2.<br />

6692 ⋅10<br />

3.26<br />

According to this result, when turbine inlet pressure increases one bar, efficiency<br />

increases 0.0009514 units and if turbine inlet temperature increases one degree,<br />

efficiency increases 0.00014 units. When cooling water temperature increases one<br />

degree, efficiency decreases 0.0013413 units. If cooling water flow rate increases one<br />

kilogram per second, efficiency increases 4.3373·10 -6 units. If power increases one<br />

kilowatt, efficiency decreases 2.1687·10 -7 . If the condenser efficiency increases one<br />

39


Chapter 3<br />

unit, the efficiency <strong>of</strong> the plant increases 0.032378. The increment <strong>of</strong> isoentropic<br />

efficiencies <strong>of</strong> pump and turbine makes efficiency to increase, although the effect <strong>of</strong> the<br />

turbine is 0.42385 units <strong>of</strong> cycle efficiency per unit <strong>of</strong> turbine efficiency while the effect<br />

<strong>of</strong> the pump is only 0.0012553. Finally, an increment <strong>of</strong> one bar in the boiler pressure<br />

drop produces an efficiency decrement <strong>of</strong> only 2.6692·10 -5 units. As it can be seen, the<br />

information contained in this vector is very important because it allows to connect the<br />

variation <strong>of</strong> the free <strong>diagnosis</strong> variables to the variation <strong>of</strong> the efficiency indicator,<br />

which is the purpose <strong>of</strong> the <strong>diagnosis</strong> algorithm. It should be noted that this vector is not<br />

dimensionless.<br />

40<br />

3.2. Influence <strong>of</strong> measurement errors.<br />

The methodology presented above is able to perform a quite accurate <strong>diagnosis</strong><br />

starting from two sets <strong>of</strong> variables x 1 and x 0 corresponding to the actual and reference<br />

states <strong>of</strong> the plant. However, these plant states (at least the actual one) are determined<br />

by a performance test. This test is not exact because it relies on information provided by<br />

instruments that present errors as well as hypotheses made to overcome the lack <strong>of</strong><br />

information. These errors in the performance test may affect not only the deviation <strong>of</strong><br />

the efficiency indicator but also the share <strong>of</strong> this value in the free <strong>diagnosis</strong> variables.<br />

Measurements affecting the value <strong>of</strong> the efficiency indicator are usually a few, but<br />

measurements affecting the <strong>diagnosis</strong> results are a lot. For example, in the case <strong>of</strong> the<br />

cycle, temperature <strong>of</strong> the cooling water entering the condenser does not affect the value<br />

<strong>of</strong> the cycle efficiency, but it modifies the contribution <strong>of</strong> the condenser efficiency and<br />

the water mass flow. Another example is the value <strong>of</strong> a temperature between two stages<br />

<strong>of</strong> a turbine, which does not affect the cycle efficiency but it modifies the isoentropic<br />

efficiency <strong>of</strong> the stages upstream and downstream the measurement.<br />

The impact <strong>of</strong> a measurement error in the value <strong>of</strong> the efficiency or heat rate has<br />

been studied and it is widely applied to determine the uncertainty <strong>of</strong> the calculated<br />

efficiency. In this section, a method to determine the impact <strong>of</strong> a measurement error in<br />

the <strong>diagnosis</strong> result is explained. The methodology is very similar as that <strong>of</strong> the<br />

<strong>diagnosis</strong>.<br />

The thermal system to be analyzed is described by means <strong>of</strong> a set <strong>of</strong> nt<br />

thermodynamic variables (x). These variables correspond to properties <strong>of</strong> streams (mass


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

flow, pressure, temperature) and parameters <strong>of</strong> equipments (isoentropic efficiencies,<br />

pressure drops). It should be noted that this vector x is the same as that used for the<br />

<strong>diagnosis</strong> problem, so that, there are the same nr restrictions (R(x) = 0). A set <strong>of</strong> np<br />

variables from x (xp) is measured by the field instrumentation. So that, a VP matrix (np<br />

x nt) can be defined as follows<br />

VPij = 1 if the variable i <strong>of</strong> xp is the variable j <strong>of</strong> x<br />

VPij = 0 in other case<br />

Depending <strong>of</strong> the problem a set <strong>of</strong> nrp additional restriction may be needed:<br />

RP ( x ) = 0<br />

3.27<br />

The value <strong>of</strong> nrp should accomplish the following rule.<br />

nt = nr + nrp + np<br />

3.28<br />

Our goal is to relate the variation <strong>of</strong> the set <strong>of</strong> variables x to the variation to the<br />

measured ones xp. This can be done in the same way as in the <strong>diagnosis</strong> problem, but<br />

considering that now the free variables are the measured ones. So that:<br />

⎡ JR ⎤<br />

⎢ ⎥ ⎡ 0 ⎤<br />

⎢<br />

JRP<br />

⎥<br />

⋅Δ x = JP ⋅Δx ≅ ⎢<br />

Δ<br />

⎥<br />

⎣ x p<br />

⎢ ⎥<br />

⎦<br />

⎣ VP ⎦<br />

3.29<br />

where JR and JRP are the Jacobian matrices <strong>of</strong> the restrictions R(x) = 0 and RP(x)<br />

= 0. The relation we were looking appears by inverting the JP matrix:<br />

0 −1 ⎡ ⎤ −1<br />

⎡0⎤ Δx ≅ JP ⋅ ⎢ ⎥ = JP ⋅⎢ ⋅Δ p<br />

Δ p U<br />

⎥ x<br />

3.30<br />

⎣ x ⎦ ⎣ ⎦<br />

The introduction <strong>of</strong> the VD matrix allows to relate the variation <strong>of</strong> the <strong>diagnosis</strong><br />

variables to the variation <strong>of</strong> the measured ones:<br />

−1 ⎡0⎤ Δxd ≅VD⋅JP ⋅⎢ ⋅Δ p = ⋅Δ p<br />

U<br />

⎥ x DP x 3.31<br />

⎣ ⎦<br />

DP is a nd x np matrix whose elements DPij are the sensitivity <strong>of</strong> the <strong>diagnosis</strong><br />

variable i to the measured variable j. It is very important because it shows the impact <strong>of</strong><br />

a measurement error in the values <strong>of</strong> the free <strong>diagnosis</strong> variables. Finally, the sensitivity<br />

<strong>of</strong> the <strong>diagnosis</strong> to the measurements can be obtained by combining equations 3.11 and<br />

3.31:<br />

Δe≅ u ⋅ED⋅Δ x = u ⋅ED⋅DP⋅Δ x = u ⋅EP⋅Δx 3.32<br />

t t t<br />

d p p<br />

41


Chapter 3<br />

42<br />

where EP is a nd x np matrix whose elements EPij are the sensitivity <strong>of</strong> the impact<br />

due to the <strong>diagnosis</strong> variable i to the measured variable j. This matrix allows to<br />

understand how an error in a measurement can affect the <strong>diagnosis</strong> result.<br />

Matrices DP and EP link the problems <strong>of</strong> performance test and <strong>diagnosis</strong>. They are<br />

very useful to determine the <strong>diagnosis</strong> uncertainty due to measurement uncertainty and<br />

to choose the right instrumentation for a correct <strong>diagnosis</strong> system. In fact, <strong>diagnosis</strong><br />

uncertainty <strong>of</strong> each <strong>diagnosis</strong> variables can be obtained by:<br />

2 2 2<br />

, = ∑ ⋅ ,<br />

= 1<br />

p n<br />

xd i DPij<br />

xp j<br />

j<br />

σ σ 3.33<br />

where<br />

σ is a nd dimensional vector containing the uncertainties <strong>of</strong> the <strong>diagnosis</strong><br />

2<br />

xd<br />

2<br />

variables and σ is the np dimensional vector <strong>of</strong> the instrumentation uncertainties. The<br />

xp<br />

uncertainties <strong>of</strong> the impacts due to the measurement can be obtained as follows:<br />

2 2<br />

=<br />

, ∑ ⋅ ,<br />

=<br />

p n<br />

2<br />

e EP<br />

xdi ij xp j<br />

j i<br />

σ σ 3.34<br />

2<br />

where σ is a nd dimensional vector containing the uncertainties <strong>of</strong> the impacts<br />

exd<br />

due to the instrumentation errors. It should be highlighted that this impact only<br />

considers the effect <strong>of</strong> the measurements. There are other errors due to the <strong>diagnosis</strong><br />

methodology and the hypothesis <strong>of</strong> linearization that should be added. The effect <strong>of</strong><br />

linearization is studied in a following section.<br />

Finally, a vector ep can be defined that directly links the variation <strong>of</strong> the efficiency<br />

indicator to the variation <strong>of</strong> the measurements:<br />

Δe ≅u ⋅EP⋅Δ x = ep ⋅Δx<br />

3.35<br />

t t<br />

p p<br />

Each position <strong>of</strong> this vector is the summation <strong>of</strong> a row <strong>of</strong> the EP matrix, and<br />

represents the sensibility <strong>of</strong> the efficiency indicator to the variation <strong>of</strong> the<br />

measurements. These sensitivities are usually used to determine the uncertainty <strong>of</strong> the<br />

indicator in a performance test:<br />

2<br />

np n<br />

⎛ ⎞<br />

p<br />

2 ∂e<br />

2 2 2<br />

e = ∑⎜ ⎟ ⋅ = ⋅<br />

⎜ ⎟ x , ∑ep<br />

p i i xp,<br />

i<br />

i= 1 ∂xpi<br />

,<br />

i=<br />

1<br />

σ σ σ 3.36<br />

⎝ ⎠<br />

3.2.1 Example<br />

In this section the example presented in 3.1.1 is continued in order to clarify the<br />

concepts introduced above.


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

9 from the 18 variables <strong>of</strong> the thermodynamic model <strong>of</strong> the cycle are measured:<br />

pressure and temperature in points 2 and 3, net power, condenser pressure, inlet and<br />

outlet cooling water temperature and inlet boiler water mass flow. As it can be seen,<br />

some <strong>of</strong> the measured variables are the same as the free <strong>diagnosis</strong> ones while other does<br />

not. So that, the xp vector and the VP matrix are:<br />

( p p2 p3 T2 T3 T5 T6 m1 W)<br />

x =<br />

3.37<br />

p sat n<br />

⎛1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎞<br />

⎜ ⎟<br />

⎜<br />

0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0<br />

⎟<br />

⎜0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0⎟<br />

⎜ ⎟<br />

⎜0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0⎟<br />

VP = ⎜0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0⎟<br />

3.38<br />

⎜ ⎟<br />

⎜0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0⎟<br />

⎜0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0⎟<br />

⎜ ⎟<br />

⎜0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0⎟<br />

⎜<br />

0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0<br />

⎟<br />

⎝ ⎠<br />

Since n = 18, nr = 9 and np= 9, additional restrictions are not needed (nrp = 0). So<br />

that, the DP and EP matrices can be calculated:<br />

⎡ 0 0 1 0 0<br />

⎢<br />

⎢<br />

0<br />

⎢ 0<br />

⎢<br />

⎢−0.076898<br />

DP<br />

= ⎢ 0<br />

⎢<br />

⎢ −12.594<br />

⎢ 228.84<br />

⎢<br />

⎢ 0.28433<br />

⎢<br />

⎣ 0<br />

0<br />

0<br />

−0.1910 0<br />

0<br />

0.0052028<br />

−5<br />

6.86⋅10 1<br />

0<br />

0<br />

−2.0726 0<br />

0<br />

−7 6.45⋅10 −0.002079 −1<br />

0<br />

0<br />

−8.8836<br />

0<br />

0<br />

−0.75067 0.0032048<br />

0<br />

1<br />

0<br />

4.8328<br />

0<br />

0<br />

−9<br />

3.01⋅10 −0.000864<br />

0<br />

43


Chapter 3<br />

44<br />

0 0 0 0 ⎤<br />

0 0 0 0<br />

⎥<br />

⎥<br />

1 0 0 0 ⎥<br />

⎥<br />

461.71 −461.72 74.738 −0.022076⎥ 0 0 0 1 ⎥⎥<br />

−0.020424 0.061852 0 0 ⎥<br />

−8 −7 −8<br />

−9.49⋅10 −5.65⋅10 6.11⋅10 0 ⎥<br />

⎥<br />

−10 −9 −6<br />

3.18⋅10 2.02⋅10 −0.008258 7.96⋅10⎥ 0 0 0 0<br />

⎥⎦<br />

The first row <strong>of</strong> the DP matrix has a 1 in the third position and zeros in the other<br />

positions, because the first <strong>diagnosis</strong> variable is the turbine inlet pressure, which<br />

corresponds to the third measured variable. A similar situation occurs in the second,<br />

third and fifth rows (turbine inlet temperature, cooling water temperature and power<br />

produced). The fourth <strong>diagnosis</strong> variable is water cooling mass flow that depends <strong>of</strong> all<br />

the variables because quality at the exit turbine is not measured. The same occurs with<br />

the isoentropic efficiency <strong>of</strong> the turbine (eighth variable), although its dependence on<br />

cooling water temperatures is negligible and due to numerical errors. The sixth variable<br />

is the condenser efficiency that depends <strong>of</strong> the cooling water temperatures and the<br />

condensing pressure. The seventh variable is the pump isoentropic efficiency that only<br />

depends on pressure and temperature in point 2 and condensing pressure. Finally, the<br />

ninth <strong>diagnosis</strong> variable is the pressure drop in the boiler, so that in this row only<br />

appears one 1 and one –1.<br />

EP matrix is obtained by multiplying each row <strong>of</strong> the ED matrix times the impact<br />

factor <strong>of</strong> the <strong>diagnosis</strong> variable corresponding to that row:<br />

⎡ 0 0 0.0009514 0 0<br />

⎢<br />

⎢<br />

0 0 0 0 0.00014<br />

⎢ 0 0 0 0 0<br />

⎢ −7 −7 −6 −5 −5<br />

⎢−3.34⋅10<br />

−8.25⋅10 −8.99⋅10 −3.85⋅10 2.096⋅10 EP = ⎢ 0 0 0 0 0<br />

⎢<br />

⎢ −0.40778<br />

0 0 0 0<br />

⎢<br />

−6 −10<br />

0.28726 6.531⋅10 8.097⋅10−0.000942 0<br />

⎢<br />

−5<br />

⎢ 0.12051 2.906⋅10 −0.000881<br />

0.0013584<br />

⎢<br />

−5 −5<br />

⎣ 0 −2.67⋅10 2.669⋅10 0<br />

3.39


0 0 0 0 ⎤<br />

0 0 0 0<br />

⎥<br />

⎥<br />

−0.001341<br />

0 0 0 ⎥<br />

−8<br />

⎥<br />

0.0020026 −0.002003 0.0003242 −9.575⋅10 ⎥<br />

−7<br />

0 0 0 −2.169⋅10 ⎥<br />

⎥<br />

−0.000661 0.0020026 0 0 ⎥<br />

−10 −10<br />

−1.19⋅10 −7.09⋅10 0 0 ⎥<br />

⎥<br />

−10 −10 −6<br />

1.349⋅10 8.568⋅10 −0.0035 3.375⋅10 ⎥<br />

0 0 0 0<br />

⎥⎦<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

The interpretation <strong>of</strong> this matrix is very similar as that <strong>of</strong> the previous one. An<br />

important detail is that measured variables can affect in all the <strong>diagnosis</strong> variables,<br />

although they are <strong>diagnosis</strong> variables themselves. For example, temperature in point 3 is<br />

a <strong>diagnosis</strong> variable (so that the first row has only one 1), but it also affects other<br />

<strong>diagnosis</strong> variables such as the isoentropic efficiency <strong>of</strong> steam turbine. Finally, ep<br />

vector can be obtained:<br />

ep<br />

t −6 −6 −5<br />

(<br />

= 2.10⋅10 8.08⋅10 8.81⋅10 0.00037752 −0.0002054<br />

)<br />

3.40<br />

−10 −6<br />

⋅ − ⋅ 3.41<br />

0 2.73 10 0.003176 3.06 10<br />

The <strong>analysis</strong> <strong>of</strong> these two matrices allows to understand how a variation in a<br />

measurement modifies the results <strong>of</strong> the <strong>diagnosis</strong>. For example, a variation in the<br />

measured value <strong>of</strong> the entering cooling water temperature (sixth <strong>diagnosis</strong> variable)<br />

does not affect the calculated value <strong>of</strong> the cycle efficiency but it modifies the <strong>diagnosis</strong><br />

result. If the measured value <strong>of</strong> this temperature increases 1 degree, the impact in cycle<br />

efficiency due to the cooling water temperature decreases 0.001341 units, the impact<br />

due to condenser mass flow increases 0.0020026 units and the impact due to condenser<br />

efficiency decreases 0.000661 units. As it can be seen, an error in this measurement<br />

makes the <strong>diagnosis</strong> results to vary substantially, so that an effort to ensure good quality<br />

instrumentation in this point should be made. It should be highlighted that matrices DP,<br />

EP and vector ep are not dimensionless, so that they not only depend on the analysed<br />

plant ant the variable choice but also on the units <strong>of</strong> the variables <strong>of</strong> e, xd and xp.<br />

45


Chapter 3<br />

46<br />

3.3. Analysis <strong>of</strong> non-linearities<br />

In the previous section, the problems <strong>of</strong> <strong>diagnosis</strong> and performance test have been<br />

connected and the influence <strong>of</strong> the measurement errors in the results <strong>of</strong> <strong>diagnosis</strong> has<br />

been studied. Another source <strong>of</strong> <strong>diagnosis</strong> inaccuracy is related to the equation<br />

linearization, which is studied here.<br />

3.3.1. Evaluation <strong>of</strong> the influence <strong>of</strong> linearization errors in the<br />

<strong>diagnosis</strong> result<br />

The <strong>diagnosis</strong> methodology presented above is based on the linearization <strong>of</strong> the sets<br />

<strong>of</strong> restrictions R(x) = 0 and RD(x) = 0. If non-linear behaviour is considered, equations<br />

3.4 and 3.5 become:<br />

( )<br />

( )<br />

( )<br />

( )<br />

( )<br />

( )<br />

⎡ R x ⎤ ⎡ R x ⎤ ⎡ JR x ⎤<br />

= + ⋅Δ x+ rs<br />

1 0<br />

⎢<br />

1 ⎢<br />

⎣<br />

RD x<br />

⎥<br />

⎥<br />

⎦<br />

⎢<br />

0 ⎢<br />

⎣<br />

RD x<br />

⎥<br />

⎥<br />

⎦<br />

0<br />

x<br />

⎢ ⎥<br />

⎢ ⎥ 0 ⎣<br />

JRD x x ⎦<br />

l<br />

3.42<br />

where rsl is a (nr + nrd) x 1 vector that has been introduced in order to make exact<br />

that approximate equation. The previous equation can be reordered and expanded in<br />

order to consider also the <strong>diagnosis</strong> variables:<br />

⎡ rs ⎤<br />

JD ⋅Δ x = ⎢<br />

Δx<br />

⎥<br />

3.43<br />

⎣ d ⎦<br />

where rs is a (nr + nrd) x 1 vector that takes into account the errors due to<br />

linearization and to the non-exact restriction accomplishment <strong>of</strong> vectors x0 and x1.<br />

( )<br />

( )<br />

( )<br />

( )<br />

0 1<br />

⎡ ⎤ ⎡ ⎤<br />

R x R x<br />

rs = rs ⎢ ⎥ ⎢ ⎥<br />

l + −<br />

0 1<br />

⎢ ⎥ ⎢ ⎥<br />

⎣<br />

RD x<br />

⎦ ⎣<br />

RD x<br />

⎦<br />

Since JD, Δx and the residues are known, vectors rs and rsl can be determined by<br />

using the previous equations. It should be noted that equations corresponding to the<br />

lower part <strong>of</strong> the system are identities (Δxd,i = Δxd,i), so that they are not affected by<br />

linearization errors. This error due to restrictions (rs) is propagated when it is multiplied<br />

by JD -1 . So that, the value <strong>of</strong> Δx that can be calculated is the summation <strong>of</strong> two terms,<br />

the first one obtained from the <strong>causal</strong>ization <strong>of</strong> the <strong>diagnosis</strong> variables and the other<br />

corresponds to the error <strong>of</strong> restrictions.<br />

⎡ 0 ⎤ ⎡rs⎤ ⎣Δ⎦ ⎣ ⎦<br />

−1 −1<br />

Δ x = JD ⋅ ⎢ ⎥+ JD ⋅ ⎢ =Δ calc +Δ rs<br />

d 0<br />

⎥ x x<br />

x<br />

3.44<br />

3.45


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

There is another linearization in the <strong>diagnosis</strong> algorithm, which corresponds to the<br />

dependence <strong>of</strong> the indicator e <strong>of</strong> the set <strong>of</strong> values x:<br />

t<br />

Δ e= ex ⋅Δ x + ee<br />

3.46<br />

Since ee is the only unknown quantity in the previous equation, it can be easily<br />

determined by substituting Δx:<br />

Δ e= ex ⋅Δ x + ex ⋅Δ x + ee<br />

3.47<br />

t t<br />

calc rs<br />

if we define er as:<br />

t t −1 ⎡U D ⎤<br />

er = ex ⋅JD ⋅⎢⎥ ⎣ 0 ⎦<br />

3.48<br />

the exact equation for Δe becomes:<br />

Δ e= ed ⋅Δ x + er ⋅ rs + ee= ed ⋅Δ x + ε<br />

3.49<br />

t t t<br />

d d<br />

The previous equation is very important because it demonstrates that the efficiency<br />

variation is due to three contributions: the first one is the summation <strong>of</strong> impacts<br />

calculated by the <strong>diagnosis</strong> algorithm, the second is an error due to the linearization <strong>of</strong><br />

the restrictions and inaccuracy <strong>of</strong> the vectors x 0 and x 1 and the last one is the error due<br />

to the linearization <strong>of</strong> the expression <strong>of</strong> the efficiency indicator. If there were no<br />

linearization error, terms two and three would be null (ε = 0) and the method would be<br />

exact. However, this is not true in general, so that errors appear mainly when the two<br />

situations to be compared are not close enough. Methods to reduce these errors are<br />

presented in the next section.<br />

3.3.2. Strategies to reduce the errors <strong>of</strong> the methodology due to nonlinearities<br />

Once the error sources <strong>of</strong> the <strong>diagnosis</strong> algorithm have been studied, the <strong>analysis</strong> <strong>of</strong><br />

strategies to reduce this error can be tackled.<br />

According to the expressions presented above, matrices JR and JRD and vector ex<br />

are calculated at reference conditions (x = x 0 ). However, results may improve if these<br />

derivatives are evaluated in a point x m between x 0 and x 1 . In fact, the Mean Value<br />

Theorem states that:<br />

Let f be a function continuous on a closed interval [a, b] and differentiable in an<br />

open interval (a, b), then there is at least one point c in (a, b) where:<br />

( ) − ( )<br />

f b f a<br />

f '(<br />

c)<br />

=<br />

b−a One possibility would be to evaluate the derivatives at point x av defined as:<br />

3.50<br />

47


Chapter 3<br />

0 1<br />

av +<br />

=<br />

2<br />

x x<br />

x 3.51<br />

Another option is to calculate the matrices in points x 0 and x 1 and then use average<br />

matrices:<br />

48<br />

( ( ) 0 ( ) 1<br />

x x )<br />

1<br />

JR = ⋅ JR x + JR x 3.52<br />

2<br />

( ( ) 0 ( ) 1<br />

x x )<br />

1<br />

JRD = ⋅ JRD x + JRD x 3.53<br />

2<br />

( ( ) 0 ( ) 1<br />

x x )<br />

1<br />

ex = ⋅ ex x + ex x 3.54<br />

2<br />

This strategy can be implemented with no big difficulties and allows obtaining quite<br />

good results, as it is shown in Chapter 6.<br />

To improve the results, a second order approach might be considered. The goal<br />

would be to obtain an expression linking the variation <strong>of</strong> the efficiency indicator to the<br />

variation <strong>of</strong> the free <strong>diagnosis</strong> variables that has quadratic terms:<br />

Δ e= n<br />

⋅Δ x +<br />

n<br />

⋅ Δ x +<br />

n i−1<br />

⎛<br />

⋅Δx ⋅Δx<br />

⎞<br />

i= 1 i= 1 i= 1⎝ j=<br />

1<br />

⎠<br />

d d d<br />

2<br />

∑αi d, i ∑βii ( d, i) ∑⎜∑ βij<br />

d, i d, j⎟<br />

3.54<br />

where αi and βij are the sensitivity coefficients to be determined by the <strong>diagnosis</strong><br />

procedure.<br />

This second order approach has not only the problem <strong>of</strong> calculation complexity but<br />

also the interpretation <strong>of</strong> the crossed terms. A <strong>diagnosis</strong> procedure without crossed<br />

terms provides a more or less exact solution consisting in the share <strong>of</strong> a deviation <strong>of</strong> a<br />

global efficiency indicator into several elements each one corresponding to an<br />

independent “free” <strong>diagnosis</strong> variable. In this way, each impact represents the<br />

improvement <strong>of</strong> the efficiency indicator e that would be obtained if the corresponding<br />

anomaly were repaired. However, when crossed terms appear, this is not so simple,<br />

because the removal <strong>of</strong> two anomalies does not have exactly the same effect as the<br />

summation <strong>of</strong> the effects obtained by the removal <strong>of</strong> each anomaly separately. As an<br />

example, a very simple system with only two free <strong>diagnosis</strong> variables (x1and x2) can be<br />

considered. To fix ideas, the system could be a steam cycle, whose efficiency is related<br />

to the efficiency <strong>of</strong> a turbine and to the ambient temperature. In the reference state<br />

x = 0.7 ,<br />

0<br />

1<br />

0<br />

x 2 = 10 and 0<br />

e = 0.372 and in the operation state 1<br />

1<br />

x = 0.68,<br />

1<br />

e = 0.364 The equation linking the variations <strong>of</strong> these three variables is:<br />

15 x = and<br />

1<br />

2


( ) ( )<br />

2 2<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Δ e= 0.15⋅Δx−0.0002⋅Δ y+ 2.5⋅ Δx −0.00016⋅ Δ y + 0.01⋅Δx⋅Δ<br />

y 3.55<br />

Application <strong>of</strong> the previous equation shows that if x1 is restored to the previous<br />

value (increases 0.02), efficiency increases 0.003. If x2 is restored after restoring x1,<br />

efficiency increases 0.005, so that it reaches the reference value. On the other hand, if x2<br />

is restored first, efficiency increases 0.006, and if x1 is then restored, efficiency<br />

increases 0.002 additional points. As it can be seen, the removal <strong>of</strong> anomalies does not<br />

have exactly the same effect depending on whether the starting point is the plant with all<br />

the anomalies or if some anomalies have previously been eliminated. However, this<br />

difference is negligible in most cases, and only has a conceptual interest.<br />

This situation leads to consider that the use <strong>of</strong> crossed terms should be avoided in<br />

most cases, because the additional difficulty in <strong>analysis</strong> results does not balance the<br />

accuracy increment. If these developments are to be used, this fact should be clarified.<br />

In the text above the interest <strong>of</strong> using crossed terms has been discussed starting from<br />

the hypothesis that there is a <strong>diagnosis</strong> methodology that provides the values <strong>of</strong> α and β.<br />

However, a methodology to obtain this information has to be developed. One way to do<br />

this might be to perform a development similar to the one used above for the linear<br />

methodology. This approach has several problems. First is that the second derivatives <strong>of</strong><br />

every restriction should be introduced, which means a very important effort in the<br />

development <strong>of</strong> the <strong>diagnosis</strong> tool. Second, the <strong>diagnosis</strong> procedure would be more<br />

complex and the calculation time would increase.<br />

Another option that is not much more complex and can provide very interesting<br />

advantages is explored here.<br />

First <strong>of</strong> all, it should be highlighted that the methodology based on the linealization<br />

is exact if Δxd tends to zero. So that, it is possible to write:<br />

( )<br />

0 0<br />

t<br />

0<br />

= ⋅ d<br />

de ed dx<br />

3.56<br />

where<br />

0<br />

de and<br />

in the reference condition.<br />

0<br />

dx d are a differential variation <strong>of</strong> these variables around their value<br />

the derivatives also in the reference condition:<br />

( ( ) ) ( )<br />

0<br />

ed means that this vector has been calculated by evaluating<br />

( )<br />

t −1 0<br />

⎡ 0 ⎤<br />

ed = ex x 0 ⋅ JD x 0 ⋅ x x ⎢ ⎥<br />

⎣U ⎦<br />

3.57<br />

The same relation can be obtained around the point x 1 :<br />

( )<br />

1 1<br />

t<br />

1<br />

= ⋅ d<br />

de ed dx<br />

3.58<br />

49


Chapter 3<br />

50<br />

( ( ) ) ( )<br />

( )<br />

t −1 1<br />

⎡ 0 ⎤<br />

ed = ex x 1 ⋅ JD x 1 ⋅ x x ⎢ ⎥<br />

⎣U ⎦<br />

3.59<br />

Vectors ed represents the derivatives <strong>of</strong> the function e related to the set <strong>of</strong><br />

independent variables xd. This information is known, as well as the values <strong>of</strong> e 1 , e 0 , x 1<br />

and x 0 and can be used to obtain a relation that links Δe to Δxd.<br />

A first idea is to look for quadratic expressions including crossed terms. However,<br />

this is not possible because more information would be needed. For example, a function<br />

depending on only two variables x1 and x2 can be considered. The goal is to obtain an<br />

expression like the following:<br />

2 2<br />

( ) ( ) ( ) ( ) ( ) ( )<br />

Δ e= e−e ≅α ⋅ x − x + α ⋅ x − x + α ⋅ x − x + α ⋅ x − x + α ⋅ x −x ⋅ x − x<br />

0 0 0 0 0 0 0<br />

1 1 1 2 2 2 1,1 1 1 2,2 2 2 1,2 1 1 2 2<br />

To obtain the values <strong>of</strong> the coefficients α, this equation can be derived:<br />

∂e<br />

∂x<br />

∂e<br />

∂x<br />

∂e<br />

∂x<br />

0 0<br />

x1 , x2<br />

1<br />

1 1<br />

x1, x2<br />

1<br />

2<br />

∂e<br />

∂x<br />

2<br />

2<br />

1<br />

0 0<br />

x1 , x2<br />

1 1<br />

x1, x2<br />

= α<br />

1<br />

1 0 1 0<br />

( x x ) ( x x )<br />

= α + 2⋅α<br />

⋅ − + α ⋅ −<br />

= α<br />

1 1,1 1 1 1,2 2 2<br />

2<br />

2<br />

∂ e<br />

= 2⋅α1,1<br />

∂x<br />

2<br />

∂ e<br />

= 2⋅α<br />

∂x<br />

2<br />

2<br />

2<br />

∂ e<br />

∂x∂x 1 2<br />

1 0 1 0<br />

( x x ) ( x x )<br />

= α + 2⋅α<br />

⋅ − + α ⋅ −<br />

2 2,2 2 2 1,2 1 1<br />

2,2<br />

= 2⋅α<br />

1,2<br />

3.60<br />

3.61<br />

3.62<br />

3.63<br />

3.64<br />

3.65<br />

3.66<br />

3.67<br />

From the previous set <strong>of</strong> equations, only the values <strong>of</strong> first-order derivatives and the<br />

values <strong>of</strong> x1 and x2 in points 0 and 1 are known. So that, seven equations are available to<br />

determine the five αs and the three second-order derivatives. As it can be seen, more<br />

information is needed. One solution <strong>of</strong> this problem is not to consider the crossed term.<br />

This also allows to overcome the conceptual problem related to crossed terms presented<br />

above. The expression <strong>of</strong> the variation <strong>of</strong> e becomes:


0<br />

nd<br />

i=<br />

1<br />

αi i<br />

0<br />

i βi<br />

i<br />

0<br />

i<br />

( ) ( ) 2<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

e−e ≅ ⎡ ⋅ x − x + ⋅ x −x<br />

⎤<br />

∑ ⎢⎣ ⎥<br />

3.68<br />

⎦<br />

If this equation is derived, 3·nd conditions can be obtained:<br />

∂e<br />

∂x<br />

i<br />

∂e<br />

∂x<br />

i<br />

= α<br />

0<br />

x= x i<br />

1<br />

x= x<br />

2<br />

∂ e<br />

= ⋅β<br />

∂x<br />

2 2 i<br />

i<br />

1 0 ( x x )<br />

= α + 2⋅β<br />

⋅ −<br />

i i i i<br />

By solving the previous system, the values for α, β and the second order derivatives<br />

can be obtained. Values <strong>of</strong> alpha are direct and the others are:<br />

∂<br />

∂e ∂e<br />

−<br />

∂ −<br />

0<br />

2<br />

x= x x= x<br />

e ∂xi ∂xi<br />

= 2⋅<br />

β<br />

2 i = 1 0<br />

xi xi xi<br />

By substituting, the expression <strong>of</strong> e depending <strong>of</strong> the value <strong>of</strong> x can be obtained:<br />

⎡ ∂e ∂e<br />

⎤<br />

1 0<br />

n ⎢ −<br />

d<br />

x= x x= x ⎥<br />

0 ∂e<br />

0 ∂xi ∂xi<br />

0<br />

2<br />

e− e =<br />

⎢<br />

0<br />

x x ( xi xi )<br />

1 0 ( xi x<br />

⎥<br />

∑ ⋅ − + ⋅ −<br />

=<br />

i )<br />

3.73<br />

⎢ i= 1 ∂xi 2⋅(<br />

xi −xi<br />

)<br />

⎥<br />

⎢ ⎥<br />

⎢⎣ ⎥⎦<br />

If the previous function is evaluated in x = x 1 , it yields:<br />

∂e ∂e<br />

1 +<br />

n 0<br />

d x= x x= x<br />

1 0 ∂xi ∂xi<br />

1 0<br />

Δ e= e − e = ∑ ⋅( xi −xi<br />

)<br />

3.74<br />

i=<br />

1 2<br />

As it can be seen, when crossed terms are non considered, evaluation <strong>of</strong> derivatives<br />

in points 1 and 0 allows obtaining a second order approximation by using a linear<br />

approach. The difference between Equation 3.74 and Equations 3.52, 3.53 and 3.54 is<br />

that in the first one the average is performed after calculating ed (at the end <strong>of</strong> the<br />

process) and in the other approach the average is made at the beginning and then ed is<br />

computed.<br />

3.69<br />

3.70<br />

3.71<br />

3.72<br />

51


Chapter 3<br />

52<br />

3.4. Free <strong>diagnosis</strong> variables<br />

The <strong>diagnosis</strong> methodology presented above is able to relate the variation <strong>of</strong> an<br />

indicator <strong>of</strong> the global efficiency <strong>of</strong> the system to the variation <strong>of</strong> a set <strong>of</strong> free <strong>diagnosis</strong><br />

variables. So that, the <strong>diagnosis</strong> result depends not only on the accuracy <strong>of</strong> the<br />

methodology and the quality <strong>of</strong> the measurements but also on the choice <strong>of</strong> the<br />

efficiency indicator e and mainly the free <strong>diagnosis</strong> variables xd. The problem <strong>of</strong> free<br />

<strong>diagnosis</strong> variable election is tackled in this section. First, the mathematical conditions<br />

that these variables must fulfill are expressed. Then, some characteristic that the<br />

variables should accomplish are presented.<br />

3.4.1 Mathematical conditions required to the set <strong>of</strong> free <strong>diagnosis</strong><br />

variables<br />

The conditions that the free <strong>diagnosis</strong> variables must accomplish are imposed by the<br />

<strong>diagnosis</strong> methodology. This methodology is based on the inversion <strong>of</strong> the JD matrix,<br />

so that, the free <strong>diagnosis</strong> variables set must be chosen in order to accomplish this<br />

condition. First, the dimension <strong>of</strong> vector xd (nd) has to be the difference <strong>of</strong> the total<br />

number <strong>of</strong> variables nt minus the dimension <strong>of</strong> the independent restrictions nr and nrd.<br />

nd = nt −nr − nrd<br />

3.75<br />

Second, the rows <strong>of</strong> matrix JD have to be independent. This means that both general<br />

restrictions (R(x) = 0) and <strong>diagnosis</strong> restrictions (RD(x) = 0) have to be independent,<br />

which is obvious. Second, these restrictions and the fixation <strong>of</strong> the values <strong>of</strong> the free<br />

<strong>diagnosis</strong> variables should be independent. As an example, the pressure drop in a heat<br />

exchanger cannot be chosen as <strong>diagnosis</strong> variable if its value is simulated in the<br />

restrictions <strong>of</strong> the problem. In this case, there are three possibilities.<br />

The first one is to consider the pressure drop as an independent free <strong>diagnosis</strong><br />

variable.<br />

The second one is to introduce an equation in the restrictions that fix the value <strong>of</strong><br />

this variable or introduces their relation to other variables <strong>of</strong> the problem:<br />

2<br />

⎛ m<br />

⎞<br />

Δ p=Δp0⋅⎜ ⎟<br />

m<br />

0<br />

⎝ ⎠<br />

where 0 p Δ and m 0 are not additional variables but constants.<br />

A third possibility is to use the previous equation but adding a factor that becomes<br />

the free <strong>diagnosis</strong> variable:<br />

3.76


2<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

⎛ m<br />

⎞<br />

Δ p= ϕ ⋅ p0⋅⎜ ⎟<br />

3.77<br />

⎝m0⎠ The choice <strong>of</strong> one <strong>of</strong> the options above depends on the preferences <strong>of</strong> the analyst,<br />

and influences the results <strong>of</strong> the <strong>diagnosis</strong>. This is the objective <strong>of</strong> the following<br />

paragraphs. It should be highlighted that if second option is chosen, the equation should<br />

be accomplished by the two sets <strong>of</strong> variables entering the <strong>diagnosis</strong> procedure (x 0 and<br />

x 1 ). If this is not true (e.g. both pressures are measured in the performance test), errors<br />

appears because restriction are not accomplished (see paragraph 3.3.1).<br />

3.4.2. Influence <strong>of</strong> a modification in the free <strong>diagnosis</strong> variable choice<br />

on the <strong>diagnosis</strong> result.<br />

In this paragraph, it is shown how the variation <strong>of</strong> only a few free <strong>diagnosis</strong><br />

variables can affect the impact <strong>of</strong> all the free <strong>diagnosis</strong> variables.<br />

The impact in the efficiency <strong>of</strong> a system for a certain set <strong>of</strong> free <strong>diagnosis</strong> variables<br />

is given by the following equation:<br />

e<br />

⎡ JR ⎤<br />

ex<br />

⎢<br />

JRD<br />

⎥<br />

−1<br />

⎡ 0 ⎤<br />

t<br />

Δ ≅ ⋅<br />

⎢ ⎥<br />

⋅⎢ Δx<br />

⎥<br />

d<br />

⎢⎣ VD ⎥⎦<br />

⎣ ⎦<br />

If the set <strong>of</strong> free <strong>diagnosis</strong> variables is modified, VD and xd vary (VD’ and xd’) and<br />

the other terms remain constant:<br />

−1<br />

⎡ JR ⎤<br />

t<br />

e<br />

⎢ ⎥ ⎡ 0 ⎤<br />

Δ ≅ex ⋅<br />

⎢<br />

JRD<br />

⎥<br />

⋅⎢ ,<br />

Δ<br />

⎥<br />

d<br />

⎢ ' ⎥<br />

⎣ x ⎦<br />

⎣VD ⎦<br />

The previous equation can be expanded:<br />

−1 −1<br />

⎡ JR ⎤ ⎡ JR ⎤<br />

0 0<br />

t ⎡ ⎤ t ⎡ ⎤<br />

Δ e = ex ⋅<br />

⎢ ⎥ ⎢ ⎥<br />

⎢<br />

JRD<br />

⎥<br />

⋅ ⎢ ⎥+ ex ⋅<br />

⎢<br />

JRD<br />

⎥<br />

⋅⎢<br />

,<br />

Δ<br />

⎥<br />

d Δ d −Δ d<br />

⎢ '⎥ ⎣ x ⎦<br />

⎢ '⎥<br />

⎣ x x ⎦<br />

⎣VD ⎦ ⎣VD ⎦<br />

3.78<br />

3.79<br />

3.80<br />

−1 −1 −1 −1<br />

⎧ ⎫<br />

⎡ JR ⎤ ⎛⎡ JR ⎤ ⎡ JR ⎤ ⎞ ⎡ JR ⎤<br />

0 0 0<br />

t ⎪ ⎡ ⎤ ⎜ ⎟ ⎡ ⎤ ⎡ ⎤⎪<br />

⎪<br />

Δ e = ex ⋅<br />

⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥<br />

⎨⎢ JRD<br />

⎥<br />

⋅ ⎢ ⎥+ ⎜⎢ JRD<br />

⎥<br />

−<br />

⎢<br />

JRD<br />

⎥ ⎟⋅<br />

⎢ ⎥+ ⎢<br />

JRD<br />

⎥<br />

⋅⎢<br />

, ⎥⎬<br />

⎪ ⎣Δxd⎦ ⎜ Δ d Δ d −Δ d<br />

⎢ ⎥ ⎢ '⎥ ⎢ ⎥ ⎟ ⎣ x ⎦<br />

⎢ '⎥<br />

⎣ x x ⎦⎪<br />

⎪⎩⎣ VD ⎦ ⎝⎣VD ⎦ ⎣ VD ⎦ ⎠ ⎣VD ⎦<br />

⎪⎭<br />

3.81<br />

53


Chapter 3<br />

The first term <strong>of</strong> the right-hand side is the impact <strong>of</strong> the initial free <strong>diagnosis</strong><br />

variables. The last term represents the direct impact due to the modification <strong>of</strong> the free<br />

<strong>diagnosis</strong> variables. The second term represent how the original variables xd have<br />

additional impact due to the modification <strong>of</strong> the variables. This additional term confirms<br />

that the modification <strong>of</strong> one or more variables can affect the whole <strong>diagnosis</strong> result.<br />

3.4.3. Choice <strong>of</strong> the free <strong>diagnosis</strong> variables.<br />

In the previous paragraphs the conditions that free <strong>diagnosis</strong> variables have to<br />

accomplish have been presented. However, there are several possibilities so that the<br />

analyst can choose the best option depending on his preferences. It has also been shown<br />

that depending on the choice <strong>of</strong> the variables the result <strong>of</strong> the <strong>diagnosis</strong> may vary<br />

substantially.<br />

Free <strong>diagnosis</strong> variables are classified into: i) ambient conditions and fuel quality<br />

(external variables), ii) set-points and iii) parameters <strong>of</strong> component efficiency. Usually,<br />

there are no problems in the choice <strong>of</strong> <strong>diagnosis</strong> variables for external causes and set<br />

points, because they are universally accepted and each variable corresponds to a<br />

measured quantity: temperature, pressure, mass flow, composition, heating value and so<br />

on.<br />

However, a right choice <strong>of</strong> a <strong>diagnosis</strong> variable to represent efficiency <strong>of</strong> a<br />

component is usually more complex. The main reason is that component efficiencies are<br />

usually determined by indicators including several variables <strong>of</strong> the flows entering and<br />

exiting the component. For example: isoentropic efficiency <strong>of</strong> a turbine, efficiency <strong>of</strong> a<br />

heat exchanger, pressure drop in a pipe and so on. The direct use <strong>of</strong> these indicators as<br />

free <strong>diagnosis</strong> variables has two advantages: i) they can be calculated directly from the<br />

properties <strong>of</strong> the streams in the plant and ii) the have a widely accepted definition.<br />

On the other hand, it should be remembered that the degradation <strong>of</strong> the components<br />

behaviour comes from a microscopic level (fouling, erosion). Although this degradation<br />

implies a variation <strong>of</strong> component efficiency indicators, this variation may be also caused<br />

by variations <strong>of</strong> the operation point <strong>of</strong> the component (induced malfunctions). For<br />

example, pressure drop in a heat exchanger can vary due to fouling but also due to a<br />

variation <strong>of</strong> the flow crossing this component. If the evolution <strong>of</strong> this parameter is<br />

plotted, an oscillating tendency could be seen. This problem can be overcome by<br />

introducing information <strong>of</strong> the component behaviour:<br />

54


2<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

⎛ m<br />

⎞<br />

Δ p= ϕ ⋅Δp0⋅⎜ ⎟<br />

3.77<br />

⎝m0⎠ Now the <strong>diagnosis</strong> variable is the factor φ that substitutes the pressure drop. Since a<br />

new equation and a new variable have been introduced, the mathematical conditions for<br />

the free <strong>diagnosis</strong> variables can be fulfilled. The use <strong>of</strong> the factor φ allows to filtrate the<br />

pressure drop variations due to the modification <strong>of</strong> the operation point <strong>of</strong> the heat<br />

exchanger. So that if the evolution <strong>of</strong> this factor is plotted, the graph would be<br />

composed <strong>of</strong> long slowly increasing zones showing the fouling increment separated by<br />

dramatic drops corresponding to the cleaning. Another option would be to use a factor<br />

which is added instead <strong>of</strong> multiplied:<br />

2<br />

⎛ m<br />

⎞<br />

Δ p=Δ pint +Δp0⋅⎜ ⎟<br />

3.82<br />

⎝m0⎠ As it can be seen, the use <strong>of</strong> models <strong>of</strong> the component and factors instead <strong>of</strong><br />

indicators does not imply additional complexity in the formulation <strong>of</strong> the <strong>diagnosis</strong><br />

problem: only the addition <strong>of</strong> other equation and other variable. However, this solution<br />

has two important problems. First, the behaviour models <strong>of</strong> the component have to be<br />

implemented (which can be complex) and perhaps periodically revised because due to<br />

degradation the behaviour <strong>of</strong> the plant may vary. Second, the interpretation <strong>of</strong> results<br />

may become more difficult, because there is not a universal acceptation <strong>of</strong> these kinds<br />

<strong>of</strong> factors. So that, the decision <strong>of</strong> the free variables definitions should be made by<br />

taking into account the preferences <strong>of</strong> the people that are going to use the <strong>diagnosis</strong> tool.<br />

In general, the best option is to use conventional parameters that accept a widely<br />

accepted definition, which can be sometimes substituted by other variables by using<br />

behaviour models <strong>of</strong> the components. Another possibility for some cases may be the use<br />

<strong>of</strong> <strong>causal</strong>ity chains, which is presented below.<br />

3.5. Causality chains<br />

In the previous section, it has been explained how the analyst need to choose<br />

suitable definitions for the free <strong>diagnosis</strong> variables. This task may be difficult in some<br />

parameters describing components, because an equilibrium between an easy definition<br />

widely accepted and a good reproduction <strong>of</strong> the physical behaviour is required.<br />

55


Chapter 3<br />

If there is no definition which accomplish these two conditions at the same time, it<br />

may be interesting to use two sets <strong>of</strong> free <strong>diagnosis</strong> variables, one corresponding to<br />

parameters close to the real behaviour (first level <strong>of</strong> <strong>causal</strong>ity) and other with parameters<br />

with easy and widely accepted definition (second level <strong>of</strong> <strong>causal</strong>ity). It should be noted<br />

that most variables would be the same in both sets (ambient conditions, fuel quality<br />

indicators, set-points and some indicators <strong>of</strong> components). Besides, an interesting way<br />

to obtain an indicator close to the physical behaviour is by considering a simple<br />

indicator but with some corrections to take into account the deviation from the design<br />

point.<br />

The objective <strong>of</strong> this section is to develop a methodology to use these two levels <strong>of</strong><br />

free <strong>diagnosis</strong> variables. The idea is to make a double decomposition. First, the<br />

variation <strong>of</strong> the global efficiency indicator is divided into a summation <strong>of</strong> terms, each<br />

one caused by a free <strong>diagnosis</strong> variable <strong>of</strong> the second level. Second, each one <strong>of</strong> these<br />

terms is again separated into a summation <strong>of</strong> terms corresponding to the free <strong>diagnosis</strong><br />

variables <strong>of</strong> the first level. For example, part <strong>of</strong> the variation <strong>of</strong> the efficiency <strong>of</strong> a<br />

steam cycle is caused by a variation in the condenser effectiveness, but this variation is<br />

partly caused by degradation <strong>of</strong> the component, and partly due to variation in other<br />

parameters (ambient temperature, cooling water mass flow…).<br />

It should be noted that the definition <strong>of</strong> the free <strong>diagnosis</strong> variables <strong>of</strong> first level is a<br />

problem <strong>of</strong> modelling. If the accuracy required is not very high, perhaps the use <strong>of</strong> only<br />

one level <strong>of</strong> <strong>causal</strong>ity is enough. If this requirement increases, the use <strong>of</strong> two levels and<br />

a model may be necessary. The accuracy <strong>of</strong> the result depends on the accuracy <strong>of</strong> the<br />

model. However, the formulation <strong>of</strong> <strong>causal</strong>ity chains theory, which is tackled in this<br />

section, is independent <strong>of</strong> the complexity <strong>of</strong> the models. First, a general formulation is<br />

made, which allows to introduce any sets <strong>of</strong> free <strong>diagnosis</strong> variables. Afterwards, a<br />

formulation based on models which relate only free <strong>diagnosis</strong> variables among them is<br />

presented. Finally, a formulation which considers models which relate the free <strong>diagnosis</strong><br />

variables with all variables <strong>of</strong> the thermodynamic model is presented.<br />

3.5.1. General formulation <strong>of</strong> <strong>causal</strong>ity chains.<br />

The thermal system considered is represented by a set <strong>of</strong> nt thermodynamic variables<br />

(x). These variables can be temperatures, pressures, mass flows, indicators <strong>of</strong><br />

component’s behaviour and so on. They are linked by nr restrictions such as mass and<br />

energy balances, and definitions <strong>of</strong> parameters:<br />

56


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Rx ( )= 0<br />

3.1<br />

Besides these general restrictions, there may be other nrd, specific for the <strong>diagnosis</strong><br />

problem:<br />

RD( x )= 0<br />

3.2<br />

If the previous equations are linealized, it is possible to write:<br />

⎡ JR ⎤<br />

⎢ ⎥⋅Δx<br />

≅0<br />

3.6<br />

⎣JRD⎦ where JR and JRD are the Jacobian matrices <strong>of</strong> sets <strong>of</strong> equations R and RD.<br />

Besides, there are two sets <strong>of</strong> nd free <strong>diagnosis</strong> variables, xd1 and xd2. Associated with<br />

each one <strong>of</strong> these sets, there is one VD matrix defined as:<br />

VDij = 1 if variable j <strong>of</strong> vector x is variable i <strong>of</strong> vector xd<br />

VDij = 0 in other case.<br />

Which allows to calculate the vector <strong>of</strong> free <strong>diagnosis</strong> variables xd from the vector <strong>of</strong><br />

variables x:<br />

x = VD ⋅x<br />

3.83<br />

d1<br />

1<br />

x = VD ⋅x<br />

3.84<br />

d2<br />

2<br />

In this situation, it is possible to write:<br />

⎡ JR ⎤<br />

⎢ ⎥ ⎡ 0 ⎤<br />

⎢<br />

JRD<br />

⎥<br />

⋅Δ x = JD1 ⋅Δx ≅ ⎢<br />

Δ<br />

⎥<br />

⎢ ⎥<br />

⎣ xd<br />

1 ⎦<br />

⎣VD1 ⎦<br />

⎡ JR ⎤<br />

⎢ ⎥ ⎡ 0 ⎤<br />

⎢<br />

JRD<br />

⎥<br />

⋅Δ x = JD2 ⋅Δx ≅ ⎢<br />

Δ<br />

⎥<br />

⎢ ⎥<br />

⎣ xd2<br />

⎦<br />

⎣VD2 ⎦<br />

By inverting JD matrices, it is possible to relate the variation <strong>of</strong> all the variables<br />

with the variation <strong>of</strong> the free <strong>diagnosis</strong> variables.<br />

−1<br />

1<br />

3.85<br />

3.86<br />

Δx≅ JD ⋅Δx d1<br />

3.87<br />

−1<br />

Δ ≅ 2 ⋅Δ d2<br />

x JD x 3.88<br />

At this point, it is possible to connect the variations <strong>of</strong> both sets <strong>of</strong> free <strong>diagnosis</strong><br />

variables:<br />

Δ x = VD ⋅Δx ≅VD ⋅JD ⋅Δ x = DD ⋅Δx<br />

3.89<br />

d2 2 2<br />

−1<br />

1 d1 d1<br />

57


Chapter 3<br />

where DD is the matrix which connects the variations <strong>of</strong> the two sets <strong>of</strong> free<br />

<strong>diagnosis</strong> variables. Each element DDij indicates the variation <strong>of</strong> the i th component <strong>of</strong><br />

xd2 caused by an increment <strong>of</strong> one unit <strong>of</strong> the value <strong>of</strong> the j th component <strong>of</strong> xd1.<br />

A global efficiency indicator e is also considered. Since this indicator is a function<br />

<strong>of</strong> the set <strong>of</strong> thermodynamic variables x, there is a vector ex which relates their<br />

variations:<br />

58<br />

t<br />

Δe≅ ⋅Δ<br />

ex x 3.10<br />

The variation <strong>of</strong> this global efficiency indicator can be related to the variations <strong>of</strong><br />

both sets <strong>of</strong> free <strong>diagnosis</strong> variables:<br />

Δe≅ex ⋅JD ⋅Δ x = ed ⋅Δx<br />

3.90<br />

t −1<br />

t<br />

1 d1 1 d1<br />

Δe≅ex ⋅JD ⋅Δ x = ed ⋅Δx<br />

3.91<br />

t −1<br />

t<br />

2 d2 2 d2<br />

For convenience, matrices ed can also be expressed as a product <strong>of</strong> a unit vector and<br />

a diagonal matrix:<br />

t<br />

Δe≅ ⋅ 1⋅Δ d1<br />

u ED x 3.92<br />

t<br />

Δe≅ ⋅ 2⋅Δ d 2<br />

u ED x 3.93<br />

The last point in the development is to connect the variation <strong>of</strong> the global efficiency<br />

indicator and the variation <strong>of</strong> xd1 through the variation <strong>of</strong> xd2, by using the matrix DD:<br />

Δe≅ u ⋅ED ⋅Δx ≅ u ⋅ED ⋅DD⋅Δ x = u ⋅EDD⋅Δx 3.94<br />

t t t<br />

2 d2 2 d1 d1<br />

Matrix EDD is calculated by multiplying matrix ED2 times matrix DD. It is very<br />

important because each element EDDij represents the impact on e due to an increment <strong>of</strong><br />

a unit <strong>of</strong> the j th variable <strong>of</strong> xd1 through the i th variable <strong>of</strong> xd2.<br />

Finally, if the increment <strong>of</strong> xd1 is expressed as the product <strong>of</strong> a diagonal matrix times<br />

a unit vector, it is possible to write:<br />

Δe≅u ⋅EDD⋅Δ x = u ⋅EDD⋅ΔX ⋅ u= u ⋅DE⋅u 3.95<br />

t t t<br />

d1 d1<br />

Where DE is the product <strong>of</strong> EDD and ΔXd1. Each element DEij represents the impact<br />

on e due to the j th variable <strong>of</strong> xd1 through the i th variable <strong>of</strong> xd2. This is the double<br />

decomposition we were looking for.<br />

3.5.2. Causality chains based on the dependence among free <strong>diagnosis</strong><br />

variables.<br />

Theory previously developed can be used for any pair <strong>of</strong> sets <strong>of</strong> free <strong>diagnosis</strong><br />

variables. However, in the practical application, such wide range <strong>of</strong> application is


Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

usually not needed. It should be kept in mind that the main purpose <strong>of</strong> the use <strong>of</strong><br />

<strong>causal</strong>ity chains is to use free <strong>diagnosis</strong> variables which have widely accepted<br />

definitions while induced effects are minimized. For this purpose, to introduce<br />

correlations linking free <strong>diagnosis</strong> variables among themselves is <strong>of</strong>ten enough. For<br />

example, to consider the influence <strong>of</strong> power produced or ambient conditions on the<br />

efficiency <strong>of</strong> a component. This is the task tackled in this section.<br />

Given a thermal system, described by a thermodynamic model, the decomposition<br />

<strong>of</strong> the variation <strong>of</strong> a global efficiency indicator e into a summation <strong>of</strong> terms<br />

corresponding to a set <strong>of</strong> free <strong>diagnosis</strong> variables xd is given by the following<br />

expression:<br />

t<br />

Δe≅ u ⋅ED⋅Δx 3.11<br />

d<br />

Besides, it is possible to decompose the free <strong>diagnosis</strong> into an intrinsic and an<br />

induced term, by using a model which connects each one <strong>of</strong> the free <strong>diagnosis</strong> variables<br />

with the others:<br />

x = x + AD⋅ x + bd 3.96<br />

d d, int<br />

d<br />

where AD is a matrix <strong>of</strong> correction factors which allows to take into account the<br />

crossed influence among free <strong>diagnosis</strong> variables (for example, influence <strong>of</strong> load on the<br />

effectiveness <strong>of</strong> a heat exchanger). It can be determined by modelling or, performing<br />

regression from plant data. The role <strong>of</strong> bd is explained later. It should be noted that most<br />

values <strong>of</strong> AD and bd are usually zero, because, in most cases crossed effects are not<br />

considered. Besides, elements <strong>of</strong> the diagonal <strong>of</strong> AD have to be zero.<br />

If increments are considered, the previous equation becomes:<br />

Δ x =Δ x + AD⋅Δx 3.97<br />

d d, int<br />

d<br />

which can be substituted in equation 3.11:<br />

Δe ≅ u ⋅ED⋅Δ x + u ⋅ED⋅AD⋅Δ x = u ⋅ED⋅Δ x + u ⋅EAD⋅Δx 3.98<br />

t t t t<br />

d, int d d, int<br />

d<br />

where each element EADij represents the variation <strong>of</strong> e induced by an increment <strong>of</strong><br />

one unit <strong>of</strong> the free <strong>diagnosis</strong> variable j through the free <strong>diagnosis</strong> variable i.<br />

If the vector <strong>of</strong> the variation <strong>of</strong> the free <strong>diagnosis</strong> variables is expressed as a<br />

product <strong>of</strong> a diagonal matrix and a unit vector, it is possible to write:<br />

Δe≅ u ⋅ED⋅Δ x + u ⋅EAD⋅ΔX ⋅ u= u ⋅Δ e + u ⋅EA⋅u 3.99<br />

t t t t<br />

d, int d<br />

int<br />

The previous equations show how the variation <strong>of</strong> e is decomposed into two terms;<br />

the first one is intrinsic, and the second is induced by the other free <strong>diagnosis</strong> variables.<br />

Each element EAij indicates the variation <strong>of</strong> e induced by the free <strong>diagnosis</strong> variable j<br />

through the free <strong>diagnosis</strong> variable i.<br />

59


Chapter 3<br />

The last point to be dealt with in this section is the determination <strong>of</strong> vector bd. It<br />

has been seen how it does not influence the <strong>diagnosis</strong> because it does not affect<br />

increments. However, it may be important in the interpretation because it determines the<br />

value <strong>of</strong> the vector <strong>of</strong> intrinsic free <strong>diagnosis</strong> variables (xd,int). It is suggested to<br />

calculate it by using the following expression:<br />

60<br />

bd =−AD⋅x d,ref<br />

3.100<br />

where the subscript ref means reference point, which can be an average point or<br />

the reference state <strong>of</strong> the <strong>diagnosis</strong>. This choice makes that, for a reference value <strong>of</strong> the<br />

free <strong>diagnosis</strong> variables, the value <strong>of</strong> induced effects is zero.<br />

Finally it should be noted that, although the development has been made for a<br />

linear dependence among free <strong>diagnosis</strong> variables, it is possible to consider any function<br />

linealized by applying Taylor series.<br />

3.5.3. Causality chains based on the dependence <strong>of</strong> free <strong>diagnosis</strong><br />

variables with other variables.<br />

The possibility <strong>of</strong> introducing dependence among free <strong>diagnosis</strong> variables,<br />

developed in the previous section, can be very interesting in some cases. However, to<br />

introduce models <strong>of</strong> the components, it is preferable to have the possibility to take into<br />

account the dependence <strong>of</strong> the free <strong>diagnosis</strong> variables with all the other variables. This<br />

task can be performed by a quite similar approach, as it is explained in this section.<br />

If there is a set <strong>of</strong> models which connect free <strong>diagnosis</strong> variables which the set <strong>of</strong><br />

variables, there is a matrix AX which contains the partial derivatives <strong>of</strong> free <strong>diagnosis</strong><br />

variables in relation to all the variables (For example, the dependence <strong>of</strong> condenser<br />

effectiveness with cooling water mass flow and heat to be exchanged)<br />

Δxd≅ AX ⋅Δx<br />

3.101<br />

It should be noted that matrix AX has to have zeros in the positions AXij if the i th free<br />

<strong>diagnosis</strong> variable is the j th variable <strong>of</strong> vector x. Besides, increment <strong>of</strong> x can be related<br />

−1<br />

to the increment <strong>of</strong> xd ( Δx≅ JD ⋅Δx d ). So, matrix AD, which contains the crossed<br />

influence among free <strong>diagnosis</strong> variables, can be calculated as:<br />

−1<br />

AD= AX ⋅JD<br />

3.102<br />

Once the matrix AD has been calculated, the procedure is exactly the same<br />

explained for <strong>causal</strong>ity chains based on the dependence among free <strong>diagnosis</strong> variables.


3.6. Calculation <strong>of</strong> fuel impact on a time span.<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Diagnosis can be defined as the comparison <strong>of</strong> an operation situation with a situation<br />

considered as reference; in other words, this is a static comparison. However,<br />

degradation phenomena are dynamic (although usually very slow), and not always is<br />

enough the comparison <strong>of</strong> two situations. Fortunately, the <strong>diagnosis</strong> method can be<br />

easily extended to deal with these situations.<br />

First, the global efficiency has to represent resources consumption per unit <strong>of</strong> time<br />

<strong>of</strong> the system (et). Given a free <strong>diagnosis</strong> variable i, its impact during a given time span<br />

from t1 to t2 can be calculated by integration:<br />

t2 t2<br />

∫ ∫ 3.103<br />

Δ e = Δe⋅dt ≅ ed ⋅Δx⋅dt t1−t2, i t, i t, i d, i<br />

t1 t1<br />

The previous integral can be calculated numerically by dividing the total time span.<br />

The detail <strong>of</strong> this division would depend on the information available.<br />

This approach can be useful for dynamic situations such as progressive degradation<br />

<strong>of</strong> a component (long time) or short time situations such as fouling.<br />

It can be directly used to evaluate the total impact due to a given cause during a past<br />

period. Besides, it can be applied to predict future situations if a degradation model is<br />

available. Neural networks can be very useful for this task.<br />

3.7. Conclusion<br />

In this chapter, the quantitative <strong>causal</strong>ity <strong>analysis</strong> for the <strong>diagnosis</strong> <strong>of</strong> energy<br />

systems has been developed. Although the method starts from a <strong>diagnosis</strong> algorithm,<br />

several features have been developed and analyses have been made in order to develop a<br />

complete <strong>diagnosis</strong> methodology.<br />

First <strong>of</strong> all, a new nomenclature has been introduced in order to clarify the notation<br />

and the concepts. Second, a methodology to quantify the influence <strong>of</strong> measurement<br />

errors on the <strong>diagnosis</strong> results has been developed. Then, the error due to linearization<br />

made in the method is analyzed, and techniques to reduce it are commented.<br />

Afterwards, some ideas on the choice <strong>of</strong> the free <strong>diagnosis</strong> variables are presented. If it<br />

is difficult to choose a definition for a free <strong>diagnosis</strong> variable which has wide<br />

acceptation and reproduces the physical behaviour <strong>of</strong> the system at the same time, it is<br />

61


Chapter 3<br />

possible to apply <strong>causal</strong>ity chains theory developed here. Finally, the idea <strong>of</strong> fuel<br />

impact during a time span has been introduced.<br />

A simple example has been used to clarify the basic ideas presented. Besides, most<br />

<strong>of</strong> the topics developed here are applied to a real example in Chapter 6.<br />

62


4. Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other<br />

<strong>diagnosis</strong> <strong>methods</strong><br />

In the previous chapter, a complete <strong>diagnosis</strong> methodology has been presented. This<br />

methodology is based on the linearization <strong>of</strong> equations <strong>of</strong> the thermodynamic<br />

representation <strong>of</strong> the thermal system to be diagnosed. In next chapters, it is going to be<br />

applied to the <strong>diagnosis</strong> <strong>of</strong> a three units coal fired power plant. However, as it has been<br />

seen in chapter two, there are other methodologies, several <strong>of</strong> them based on the<br />

thermoeconomic representation <strong>of</strong> the system. It would be interesting to analyze<br />

connections <strong>of</strong> both formulations. This is the aim <strong>of</strong> the first section <strong>of</strong> this chapter.<br />

In the second section, the capabilities <strong>of</strong> linear regression and neural networks for<br />

the <strong>diagnosis</strong> problem are explored<br />

4.1. Quantitative <strong>causal</strong>ity <strong>analysis</strong> and the fuel impact<br />

formula<br />

The fuel impact formula was developed to relate the variation <strong>of</strong> the exergy<br />

resources entering the plant to the variations <strong>of</strong> the independent variables <strong>of</strong> the<br />

thermoeconomic model (unit exergy consumptions and final products). Problems in the<br />

application <strong>of</strong> this formula appear because unit exergy consumptions are not actually<br />

independent variables, so that induced effects are present.<br />

On the other hand, the quantitative <strong>causal</strong>ity <strong>analysis</strong> method is able to relate a set <strong>of</strong><br />

independent variables to a user-defined efficiency indicator. However, this flexibility to<br />

be adapted to the real thermodynamic behaviour <strong>of</strong> the system makes impossible to<br />

obtain analytical formulas, so numerical calculations are used.<br />

63


Chapter 4<br />

The idea proposed here is to link both approaches by considering the independent<br />

variables <strong>of</strong> the thermoeconomic <strong>analysis</strong> (unit exergy cost and plant product) as the<br />

dependent efficiency indicators <strong>of</strong> quantitative <strong>causal</strong>ity approach. This allows to<br />

determine the value <strong>of</strong> both intrinsic and induced effects.<br />

4.1.1. From free <strong>diagnosis</strong> variables to unit exergy cost and final<br />

product.<br />

Given a thermodynamic description <strong>of</strong> an energy system, where a set <strong>of</strong> free<br />

<strong>diagnosis</strong> variables consistent with a set <strong>of</strong> restrictions has been chosen. The<br />

quantitative <strong>causal</strong>ity algorithm allows to relate the variation <strong>of</strong> all the variables<br />

describing the system to the variations <strong>of</strong> the free <strong>diagnosis</strong> variables.<br />

64<br />

0<br />

Δx ≅ JD ⋅⎢ ⎥<br />

−1 ⎡ ⎤<br />

⎣Δxd⎦ Once an efficiency indicator e has been chosen, the variation <strong>of</strong> this indicator can be<br />

related to the variation <strong>of</strong> the free <strong>diagnosis</strong> variables by using the vector ex containing<br />

the partial derivatives <strong>of</strong> e respect to the set <strong>of</strong> thermodynamic variables x:<br />

t t −1 ⎡ 0 ⎤<br />

Δe≅ex ⋅Δx≅ex ⋅JD ⋅⎢ Δ<br />

⎥<br />

4.2<br />

⎣ xd<br />

⎦<br />

Variable e is defined as a kind <strong>of</strong> efficiency indicator only for convenience, because<br />

<strong>of</strong> this is the interest <strong>of</strong> the <strong>diagnosis</strong> methodology. It is possible to relate other<br />

variables depending on the vector x only by replacing the vector ex by other vector<br />

containing the partial derivatives <strong>of</strong> the new variable to the components <strong>of</strong> the vector x.<br />

The same system can be represented by a thermoeconomic model. Once the model<br />

has been described and the purpose <strong>of</strong> each susbsystem introduced, the magnitudes <strong>of</strong><br />

this representation depend only on the set <strong>of</strong> thermodynamic variables x. So that, the<br />

variation <strong>of</strong> this magnitudes can be related to Δx only by multiplying times a suitable<br />

vector containing the corresponding partial derivatives. Thermoeconomic representation<br />

<strong>of</strong> a system includes exergy <strong>of</strong> flows, exergy and unit exergy costs <strong>of</strong> flows, and unit<br />

exergy consumptions <strong>of</strong> components. However, when the fuel impact formula is applied<br />

(for small deviations), the only free variables are unit exergy consumptions and exergy<br />

flows leaving the component (final products).<br />

n ⎛ n<br />

* 1<br />

0 * 1 ⎞<br />

Δ FT = ∑∑ ⎜ kP, j( x ) Δ κ jiPi( x ) + kP,<br />

i( x ) Δϖi⎟<br />

4.3<br />

i= 1⎝ j=<br />

0<br />

⎠<br />

4.1


Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

It should be noted that in this section the set <strong>of</strong> thermodynamic variables<br />

representing actual and reference states are represented as x1 and x 0 (nomenclature used<br />

in the quantitative <strong>causal</strong>ity <strong>analysis</strong>) instead on x and x0 (usual in thermoeconomic<br />

<strong>analysis</strong>).<br />

In this situation, it is interesting to consider κij and ωi as the indicators depending on<br />

x. Each specific consumption κij depends on x, so that:<br />

ij<br />

ij, t<br />

nt<br />

l<br />

∑ kxij l=<br />

1<br />

xl<br />

Δκ≅kx ⋅Δ x = ⋅Δ<br />

4.4<br />

Where kx ij is an nt x 1 vector containing the partial derivatives <strong>of</strong> the unit exergy<br />

consumption κij related to the set <strong>of</strong> nt variables <strong>of</strong> the thermodynamic model <strong>of</strong> the<br />

system. By using the vector xd, the variation <strong>of</strong> the unit exergy consumption can be<br />

finally related to the variation <strong>of</strong> the free <strong>diagnosis</strong> variables:<br />

nd<br />

ij, T −1<br />

⎡0⎤ r l r<br />

Δ κij = kx ⋅JD ⋅⎢ ⎥⋅Δ<br />

xd<br />

+ κij = ∑ kdij ⋅Δx<br />

d , l + κij<br />

4.5<br />

⎣U⎦ l=<br />

1<br />

Where the term r<br />

κ ij is the residual term, introduced to keep the equal sign. The<br />

previous relation can be extended to all the unit exergy cost <strong>of</strong> the system, including the<br />

κ :<br />

t<br />

unit exergy costs <strong>of</strong> flows coming from the outside = ( , k ,..., k )<br />

e<br />

k01 02 0n<br />

t n<br />

⎡ d Δ ⎤ e l r<br />

⎢ ⎥ = ⋅Δ xdl<br />

, +<br />

⎣Δ⎦ l=<br />

1<br />

∑<br />

k<br />

KD KD<br />

4.6<br />

KP<br />

Where each one <strong>of</strong> the nd matrices KD is a ((n + 1) x n) matrix containing the<br />

elements l<br />

ij<br />

r<br />

k or k :<br />

ij<br />

l l l<br />

⎡kd01 kd02 ... kd ⎤ 0n<br />

⎢ l l l ⎥<br />

l kd11 kd12 ... kd0n<br />

KD = ⎢ ⎥<br />

4.7<br />

⎢ . . ... . ⎥<br />

⎢ l l l ⎥<br />

⎢⎣kdn1 kdn2 ... kdnn⎥⎦<br />

r r r<br />

⎡kd01 kd02 ... kd ⎤ 0n<br />

⎢ r r r ⎥<br />

r kd11 kd12 ... kd0n<br />

KD = ⎢ ⎥<br />

4.8<br />

⎢ . . ... . ⎥<br />

⎢ r r r ⎥<br />

⎢⎣kdn1 kdn2 ... kdnn⎥⎦<br />

A similar development can be made for each one <strong>of</strong> the products <strong>of</strong> the plant:<br />

it ,<br />

Δωi≅ ⋅Δ<br />

wx x 4.9<br />

nd nd<br />

it , −1<br />

⎡0⎤ r l r l r<br />

Δ ωi = wx ⋅JD ⋅⎢ ⎥⋅Δ<br />

xd<br />

+ ωi = ∑wdi ⋅Δ xd,<br />

l + ωi = ∑Δ<br />

ωi + ωi<br />

⎣U⎦ l= 1 l=<br />

1<br />

4.10<br />

65


Chapter 4<br />

66<br />

Finally, the previous equation can be applied for all the products leaving the plant:<br />

nd ∑<br />

l<br />

xdl<br />

,<br />

r<br />

nd<br />

∑<br />

l r<br />

l= 1 l=<br />

1<br />

Δ ω = ω ⋅Δ + ω = Δ ω + ω 4.11<br />

If wastes (or residues) are considered in the thermoeconomic model, the fuel impact<br />

formula becomes:<br />

n n n<br />

⎛ * 1 0 * 1 0 * 1 ⎞<br />

Δ FT = ∑∑ ⎜ kP, j( x ) Δ κ jiPi( x ) + ∑kPR,<br />

j( x ) Δ θ jiPi( x ) + kP,<br />

i( x ) Δϖi⎟<br />

4.12<br />

i= 1⎝ j= 0 j=<br />

1<br />

⎠<br />

In this situation, θij are also independent variables <strong>of</strong> the thermoeconomic model,<br />

which have to be decomposed:<br />

nd<br />

ij, t −1<br />

⎡0⎤ r l r<br />

Δ θij = tx ⋅JD ⋅⎢ ⎥⋅Δ<br />

xd<br />

+ θij = ∑tdij⋅Δx<br />

d , l + θij<br />

⎣U⎦ l=<br />

1<br />

If all elements θij are considered together:<br />

r or θ ij .<br />

nd<br />

l=<br />

1<br />

4.13<br />

l r<br />

Δ Θ = ∑TD ⋅Δ xdl<br />

, + TD<br />

4.14<br />

l<br />

where each one <strong>of</strong> the nd matrices TD is a (n x n) matrix containing the elements θ ij<br />

Equations presented above allow to relate the independent variables <strong>of</strong> the<br />

thermoeconomic model <strong>of</strong> a system to the free <strong>diagnosis</strong> variables. This result is very<br />

useful to determine the effects induced by the variation <strong>of</strong> ambient conditions and set<br />

points and to malfunctions appearing in other components.<br />

4.1.2. Quantification <strong>of</strong> intrinsic and induced malfunctions<br />

According to the fuel impact formula, variations <strong>of</strong> specific exergy consumptions<br />

and final products are the source <strong>of</strong> the variation <strong>of</strong> fuel entering the plant. Besides,<br />

these variations are also the source <strong>of</strong> malfunctions and malfunction costs. In this<br />

section, expressions obtained previously for Δ κij<br />

, Δ ωi<br />

and Δ θij<br />

are substituted in the<br />

fuel impact formula and in the definitions <strong>of</strong> malfunctions and malfunction costs, in<br />

order to definitely link the <strong>diagnosis</strong> approach based on the thermoeconomic model and<br />

quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

If the decomposition <strong>of</strong> Δ κij<br />

, Δ ωi<br />

and Δ θij<br />

into elements corresponding to the free<br />

<strong>diagnosis</strong> variables is substituted in the fuel impact formula, it yields:


Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

T<br />

n ⎡ n<br />

∑∑ ⎢<br />

nd * 1 ⎛<br />

P, j ( x ) ⎜∑ l<br />

ji<br />

r ⎞<br />

ji ⎟<br />

0<br />

i ( x )<br />

n<br />

nd * 1 ⎛<br />

∑ PR, j ( x ) ⎜∑ l<br />

θ ji<br />

r ⎞<br />

θ ji⎟ 0<br />

i( x )<br />

nd<br />

* 1<br />

P, i( x ) ∑(<br />

l<br />

ωi ⎤ r<br />

ωi<br />

) ⎥<br />

i= 1 j= 0 l= 1 j= 1 l= 1 l=<br />

1<br />

Δ F = k ⋅ Δ k + k ⋅ P + k ⋅ Δ + ⋅ P + k ⋅ Δ +<br />

⎣ ⎝ ⎠ ⎝ ⎠<br />

⎦<br />

4.15<br />

In the previous equation, the impact in fuel due to each variation <strong>of</strong> unit exergy<br />

consumption and each fuel variation is decomposed into impacts due the free <strong>diagnosis</strong><br />

variables. It can be rearranged by introducing elements kd, wd and td:<br />

Δ F =<br />

⎡<br />

⎢⎣ ⎛<br />

⎝<br />

k ⋅kd ⋅ P + k ⋅td ⋅ P + k ⋅wd ⎞⎤<br />

⋅Δ x<br />

⎠⎥⎦<br />

4.16<br />

+ F<br />

The previous equation shows how the fuel increment can be shared into several<br />

nd n n n<br />

* 1 l 0<br />

l<br />

* 1 l 0 * l r<br />

T ∑∑∑ ⎢ ⎜ P, j( x ) ji i( x ) , ( ) ( )<br />

i ∑ PR j x ji i x P, i i ⎟⎥<br />

d, l T<br />

l= 1 i= 1 j= 0 j=<br />

1<br />

components each one due to an independent free <strong>diagnosis</strong> variable.<br />

term, calculated as:<br />

r<br />

F T is a residual<br />

n n n<br />

r ⎡ * 1 r 0 * 1 r 0 * 1 r⎤<br />

FT = ∑∑ ⎢ kP, j( x ) ⋅kji⋅ Pi( x ) + ∑kPR,<br />

j( x ) ⋅θji⋅ Pi( x ) + kP,<br />

i( x ) ⋅ωi<br />

⎥ 4.17<br />

i= 1 ⎣ j= 0 j=<br />

1<br />

⎦<br />

Malfunctions can also be decomposed in a summation <strong>of</strong> several terms each one<br />

corresponding to an impact due to a free <strong>diagnosis</strong> variable:<br />

nd nd<br />

0 ⎛ l r ⎞ 0<br />

l r<br />

MFji =Δkji ⋅ Pi ( x ) = ⎜∑Δ k ji + k ji ⎟⋅<br />

Pi ( x ) = ∑MFji<br />

+ MFji<br />

4.18<br />

⎝ l= 1 ⎠<br />

l=<br />

1<br />

n n nd nd<br />

⎛ l r ⎞ 0<br />

l r<br />

MFi = ∑MFji = ∑⎜∑Δ k ji + k ji⎟⋅ Pi( x ) = ∑MFi<br />

+ MFi<br />

4.19<br />

j= 0 j= 0⎝ l= 1 ⎠<br />

l=<br />

1<br />

where<br />

l<br />

MF ji and<br />

to MF ji and MF i :<br />

l l<br />

ji ji i<br />

0 ( )<br />

l<br />

MF i are respectively the contribution <strong>of</strong> the l th <strong>diagnosis</strong> variable<br />

MF =Δk ⋅P x 4.20<br />

n n<br />

0<br />

∑ ( ) ∑<br />

MF = Δk ⋅ P x = MF<br />

4.21<br />

l l l<br />

i ji i ji<br />

j= 0 j=<br />

0<br />

r<br />

MF ji and<br />

r<br />

MF ji are the residual term <strong>of</strong> malfunctions. The same decomposition can<br />

be made for malfunction associated with residues:<br />

nd nd<br />

0 ⎛ l r ⎞ 0<br />

l r<br />

MRji =Δθ ji ⋅ Pi( x ) = ⎜∑Δ θ ji + θ ji ⎟⋅<br />

Pi( x ) = ∑MRji<br />

+ MRji4.22<br />

⎝ l= 1 ⎠<br />

l=<br />

1<br />

n n nd nd<br />

⎛ l r ⎞ 0<br />

l r<br />

MRi = ∑MRji = ∑⎜∑Δ θ ji + θ ji⎟⋅ Pi( x ) = ∑MRi<br />

+ MRi<br />

4.23<br />

j= 1 j= 1⎝ l= 1 ⎠<br />

l=<br />

1<br />

67


Chapter 4<br />

Malfunction cost can also be decomposed in the summation <strong>of</strong> several components,<br />

each one due to a variation <strong>of</strong> a free <strong>diagnosis</strong> variable.<br />

68<br />

*<br />

ji =<br />

* 1<br />

P, j ( x ) ⋅<br />

nd ji = ∑<br />

* 1<br />

P, j ( x ) ⋅Δ<br />

l<br />

ji ⋅<br />

0<br />

i ( x ) +<br />

* 1<br />

P, j ( x ) ⋅<br />

r<br />

ji ⋅<br />

nd<br />

0<br />

i ( x ) = ∑<br />

*, l<br />

ji +<br />

*, r<br />

ji<br />

l= 1 l=<br />

1<br />

MF k MF k k P k k P MF MF<br />

4.24<br />

n<br />

*<br />

i = ∑<br />

n nd *<br />

ji = ∑∑<br />

* 1<br />

P, j( x ) ⋅Δ<br />

l<br />

ji⋅ n<br />

0 * 1<br />

i( x ) + ∑ P, j( x ) ⋅<br />

r<br />

ji⋅ nd<br />

0<br />

i( x ) = ∑<br />

*, l<br />

i +<br />

*, r<br />

i<br />

j= 0 j= 0 l= 1 j= 0 l=<br />

1<br />

MF MF k k P k k P MF MF<br />

where<br />

l<br />

MF ji<br />

*, and<br />

l<br />

MF i<br />

*, are the contributions <strong>of</strong> x d , l<br />

( ) ( )<br />

*, l * 1 l 0<br />

ji P, j ji i<br />

Δ to<br />

MF and MF :<br />

*<br />

ji<br />

*<br />

i<br />

4.25<br />

MF = k x ⋅Δk⋅Px 4.26<br />

n n<br />

∑ , ( ) ( ) ∑<br />

MF = k x ⋅Δk ⋅ P x = MF<br />

4.27<br />

*, l * 1 l 0 *, l<br />

i P j ji i ji<br />

j= 0 j=<br />

0<br />

*,r<br />

MF ji and<br />

*,r<br />

MF i are the residual term <strong>of</strong> malfunction costs. The malfunction cost<br />

due to the variation <strong>of</strong> the product<br />

variation <strong>of</strong> free <strong>diagnosis</strong> variables:<br />

*<br />

MF 0 can also be shared into impacts due to the<br />

n n ndnd * * 1 ⎛ * 1 l * 1 r⎞ *, l *, r<br />

MF0 = ∑kPi , ( x ) ⋅Δωi≅∑⎜∑kPi , ( x ) ⋅Δ ωi+ kPi , ( x ) ⋅ ωi⎟=<br />

∑MF0<br />

+ MF0<br />

i= 1 i= 1⎝ l= 1 ⎠ l=<br />

1<br />

4.28<br />

Where<br />

l<br />

MF *,<br />

0 is the contribution <strong>of</strong> xd,l to<br />

*<br />

MF 0 and<br />

MF is the residual.<br />

Finally, the decomposition can also be extended to the malfunction cost associated<br />

with the residues:<br />

*<br />

ji =<br />

* 1<br />

PR, j ( x ) ⋅<br />

nd ji = ∑<br />

* 1 l<br />

PR, j ( x ) ⋅Δθ ji ⋅<br />

0<br />

i ( x ) +<br />

* 1 r<br />

PR, j ( x ) ⋅θ ji ⋅<br />

nd<br />

0<br />

i ( x ) = ∑<br />

*, l<br />

ji +<br />

*, r<br />

ji<br />

l= 1 l = 1<br />

MR k MR k P k P MR MR<br />

4.29<br />

n<br />

*<br />

i = ∑<br />

n nd *<br />

ji= ∑∑<br />

* 1 l<br />

PR, j( x ) ⋅Δθ ji⋅ n<br />

0 * 1 r<br />

i( x ) + ∑ PR, j( x ) ⋅θ ji⋅ nd<br />

0<br />

i( x ) = ∑<br />

*, l<br />

i +<br />

*, r<br />

i<br />

j= 0 j= 0 l= 1 j= 0 l=<br />

1<br />

MR MR k P k P MR MR<br />

4.30<br />

The impact in fuel is the summation <strong>of</strong> the malfunction costs <strong>of</strong> all the components,<br />

included<br />

*<br />

MF0 and malfunctions associated with wastes:<br />

*, r<br />

0


n<br />

*<br />

n<br />

*<br />

T i i<br />

i= 0 i=<br />

1<br />

Δ F = MF + MR<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

∑ ∑ 4.31<br />

The previous equation can be expanded by introducing<br />

l<br />

MF i<br />

*, and<br />

*,l<br />

MR i :<br />

Δ F = MF + MR =<br />

⎛<br />

MF + MF<br />

⎞<br />

+<br />

⎛<br />

MR + MR<br />

⎞<br />

= MF + MF + MR + MR<br />

⎝ ⎠ ⎝ ⎠<br />

T<br />

n<br />

∑<br />

*<br />

i<br />

n<br />

∑<br />

*<br />

i<br />

n nd ∑⎜∑ *, l<br />

i<br />

*, r<br />

i ⎟<br />

n nd ∑⎜∑ *, l<br />

i<br />

*, r<br />

i ⎟<br />

nd ∑<br />

*, l *, r<br />

nd<br />

∑<br />

*, l *, r<br />

i= 0 i= 1 i= 0 l= 1 i= 1 l= 1 l= 1 l=<br />

1<br />

l<br />

Where MF *, *,l<br />

and MR are the contribution <strong>of</strong> the fuel consumption variation due<br />

to the free <strong>diagnosis</strong> variable xd,l, associated with productive flows and to residues:<br />

MF<br />

*, l<br />

=<br />

n<br />

∑<br />

i=<br />

0<br />

MF<br />

*, l<br />

i<br />

n<br />

*, l *, l<br />

i<br />

i=<br />

1<br />

4.32<br />

4.33<br />

MR = ∑ MR<br />

4.34<br />

MF<br />

MF<br />

*, 1<br />

0<br />

*, 1<br />

1<br />

n<br />

MF … d MF *,<br />

*, 2<br />

0<br />

n<br />

MF … d MF *,<br />

*, 2<br />

1<br />

0<br />

1<br />

MF<br />

MF<br />

*, r<br />

0<br />

*, r<br />

1<br />

*<br />

MF 0<br />

*<br />

MF 1<br />

… … … … … …<br />

MF<br />

*, 1<br />

n<br />

*,1<br />

MR 1<br />

n<br />

MF … d MF *,<br />

*, 2<br />

n<br />

*,2<br />

*,<br />

MR …<br />

1<br />

1 d n<br />

MR<br />

n<br />

MF<br />

MR<br />

*, r<br />

1<br />

*, r<br />

1<br />

*<br />

MF n<br />

*<br />

MR 1<br />

… … … … … …<br />

*,1<br />

MR n<br />

*,1 *,1<br />

MF + MR<br />

*,2<br />

*, n<br />

MR … d<br />

n<br />

MR n<br />

*,2 *,2<br />

MF + MR<br />

*, n *,<br />

… d nd<br />

MF + MR<br />

*,r<br />

MR n<br />

*<br />

MR n<br />

… T F Δ<br />

Table 4.1: Table <strong>of</strong> malfunctions and free <strong>diagnosis</strong> variables (MFD).<br />

As it is shown by the previous equations, the impact on fuel can be obtained by a<br />

double summation <strong>of</strong> the cost <strong>of</strong> malfunctions in each component due to each free<br />

<strong>diagnosis</strong> variable. So that,<br />

l<br />

MF i<br />

*, and<br />

*,l<br />

MRi elements can be represented in a table,<br />

where each column correspond to a free <strong>diagnosis</strong> variable xd,l and each component can<br />

have one or two rows depending on whether there are waste flows associated with it.<br />

The summation <strong>of</strong> elements in a column is<br />

l<br />

MF *, *,l<br />

+ MR and the summation <strong>of</strong><br />

69


Chapter 4<br />

elements in a row is MF or<br />

70<br />

*<br />

i<br />

*<br />

MR i . Finally, the total summation corresponds to ΔFT.<br />

This table shows how the variation <strong>of</strong> fuel consumption is due to the malfunctions<br />

induced by the free <strong>diagnosis</strong> variables in all the components and in the variation <strong>of</strong><br />

final product. So that, it can be named table <strong>of</strong> malfunctions induced by the free<br />

<strong>diagnosis</strong> variables (MFD).<br />

The previous <strong>analysis</strong> has been done by decomposing the variation <strong>of</strong> unit exergy<br />

consumptions into contributions due to each free <strong>diagnosis</strong> variable. This study is<br />

interesting but when the system is big and there are more than twenty or thirty variables<br />

this <strong>analysis</strong> may become rather complex. So that, it may be better to do a similar study<br />

but separating the variation <strong>of</strong> unit exergy consumptions into intrinsic variations (int)<br />

and variations induced by other components (oc), ambient conditions (ac), fuel quality<br />

(fq) and set points (sp). It should be noted that variations induced by ambient<br />

conditions, fuel quality and set points are calculated simply by a summation <strong>of</strong> the<br />

variations induced by the free <strong>diagnosis</strong> variables related to these categories. However,<br />

separation <strong>of</strong> intrinsic and induced effect depends not only on the free <strong>diagnosis</strong><br />

variables but also on the component. Each free <strong>diagnosis</strong> variable corresponding to a<br />

parameter <strong>of</strong> a component originates an intrinsic variation <strong>of</strong> κij and θij in that<br />

component and induced variations in the other components.<br />

Δ κ =Δ κ +Δ κ +Δ κ +Δ κ +Δ κ +Δ κ<br />

4.35<br />

int oc ac fq sp r<br />

ij ij ij ij ij ij ij<br />

Δ θ =Δ θ +Δ θ +Δ θ +Δ θ +Δ θ +Δ θ<br />

4.36<br />

int oc ac fq sp r<br />

ij ij ij ij ij ij ij<br />

Δ ω =Δ ω +Δ ω +Δ ω +Δ ω +Δ ω +Δ ω<br />

4.37<br />

int oc ac fq sp r<br />

i i i i i i i<br />

The fuel impact formula can be modified to consider separately intrinsic and<br />

induced effects:<br />

n n ⎡ * 1 int oc ac fq sp r<br />

0 ⎤<br />

Δ FT = ∑∑ ⎢ kP, j( x ) ⋅( Δ κ ji +Δ κ ji +Δ κ ji +Δ κ ji +Δ κ ji + κ ji) ⋅ Pi(<br />

x ) ⎥+<br />

i= 1 ⎣ j=<br />

0<br />

⎦<br />

n ⎡ n<br />

* 1 int oc ac fq sp r<br />

0 ⎤<br />

+ ∑∑ ⎢ kPR, j ( x ) ⋅( Δ θ ji +Δ θ ji +Δ θ ji +Δ θ ji +Δ θ ji + θ ji ) ⋅ Pi(<br />

x ) ⎥+<br />

i= 1 ⎣ j=<br />

1<br />

⎦<br />

oc ac fq sp r<br />

( ) ( )<br />

n<br />

* 1 int<br />

∑ ⎡kPi , ωiωiωiωiωiω ⎤<br />

⎣<br />

x i ⎦<br />

4.38<br />

i=<br />

1<br />

+ ⋅ Δ +Δ +Δ +Δ +Δ +<br />

Malfunctions can also be decomposed into intrinsic and induced:<br />

0 int oc ac fq sp r<br />

0<br />

( ) ( ) ( )<br />

MF =Δκ ⋅ P x = Δ κ +Δ κ +Δ κ +Δ κ +Δ κ + κ ⋅P<br />

x 4.39<br />

ji ji i ji ji ji ji ji ji i


int oc ac fq sp r<br />

ji ji ji ji ji ji ji<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

MF = MF + MF + MF + MF + MF + MF<br />

4.40<br />

n n<br />

int oc ac fq sp r<br />

∑ ∑ ( ) 4.41<br />

MF = MF = MF + MF + MF + MF + MF + MF<br />

i ji ji ji ji ji ji ji<br />

j= 0 j=<br />

0<br />

MF = MF + MF + MF + MF + MF + MF<br />

4.42<br />

Where<br />

int oc ac fq sp r<br />

i i i i i i i<br />

x<br />

MFi and<br />

x<br />

MF ji (x = int, oc, ac, fq and sp) are malfunctions intrinsic and<br />

induced by other components, ambient conditions, fuel quality and set points:<br />

x x<br />

ji ji i<br />

0 ( )<br />

MF =Δk ⋅P x 4.43<br />

n n<br />

0<br />

∑ ( ) ∑<br />

MF = Δk ⋅ P x = MF<br />

4.44<br />

x x x<br />

i ji i ji<br />

j= 0 j=<br />

0<br />

A similar development can be made with malfunctions associated with residues:<br />

0 int oc ac fq sp r<br />

0<br />

( ) ( ) ( )<br />

MR =Δθ ⋅ P x = Δ θ +Δ θ +Δ θ +Δ θ +Δ θ + θ ⋅P<br />

x 4.45<br />

ji ji i ji ji ji ji ji ji i<br />

MR = MR + MR + MR + MR + MR + MR<br />

4.46<br />

int oc ac fq sp r<br />

ji ji ji ji ji ji ji<br />

n n<br />

int oc ac fq sp r<br />

∑ ∑ ( ) 4.47<br />

MR = MR = MR + MR + MR + MR + MR + MR<br />

i ji ji ji ji ji ji ji<br />

j= 0 j=<br />

0<br />

MR = MR + MR + MR + MR + MR + MR<br />

4.48<br />

int oc ac fq sp r<br />

i i i i i i i<br />

The decomposition in intrinsic and induced effects can be extended to malfunction<br />

costs:<br />

oc ac fq sp r<br />

( x ) ( x ) ( ) ( x )<br />

MF = k ⋅ MF = k ⋅ Δ k +Δ k +Δ k +Δ k +Δ k + k ⋅P<br />

* * 1 * 1 int 0<br />

ji P, j ji P, j ji ji ji ji ji ji i<br />

* *,int *, oc *, ac *, fq *, sp *, r<br />

ji ji ji ji ji ji ji<br />

4.49<br />

MF = MF + MF + MF + MF + MF + MF<br />

4.50<br />

n n<br />

oc ac fq sp r<br />

∑ ∑ ( ) 4.51<br />

MF = MF = MF + MF + MF + MF + MF + MF<br />

* * *,int *, *, *, *, *,<br />

i ji ji ji ji ji ji ji<br />

j= 0 j=<br />

0<br />

MF = MF + MF + MF + MF + MF + MF<br />

4.52<br />

Where<br />

* *,int *, oc *, ac *, fq *, sp *, r<br />

i i i i i i i<br />

x<br />

MF i<br />

*, and<br />

MF *, (x = int, oc, ac, fq and sp) are the costs <strong>of</strong> malfunctions<br />

x<br />

ji<br />

intrinsic and induced by other components, ambient conditions, fuel quality and set<br />

points:<br />

( ) ( )<br />

MF = k x ⋅Δk⋅Px 4.53<br />

*, x * 1 x 0<br />

ji P, j ji i<br />

n n<br />

∑ , ( ) ∑<br />

MF = k Δk ⋅ P x = MF<br />

4.54<br />

*, x * x 0 *, x<br />

i P j ji i ji<br />

j= 0 j=<br />

0<br />

71


Chapter 4<br />

Again, a parallel distribution is made for the malfunction costs associated with the<br />

residues:<br />

72<br />

oc ac fq sp r<br />

( x ) ( x ) ( θ θ θ θ θ θ ) ( x )<br />

MR = k ⋅ MR = k ⋅ Δ +Δ +Δ +Δ +Δ + ⋅P<br />

* * 1 * 1 int 0<br />

ji PR, j ji PR, j ji ji ji ji ji ji i<br />

* *,int *, oc *, ac *, fq *, sp *, r<br />

ji ji ji ji ji ji ji<br />

4.55<br />

MR = MR + MR + MR + MR + MR + MR<br />

4.56<br />

n n<br />

oc ac fq sp r<br />

∑ ∑ ( ) 4.57<br />

MR = MR = MR + MR + MR + MR + MR + MR<br />

* * *,int *, *, *, *, *,<br />

i ji ji ji ji ji ji ji<br />

j= 0 j=<br />

0<br />

MR = MR + MR + MR + MR + MR + MR<br />

4.58<br />

* *,int *, oc *, ac *, fq *, sp *, r<br />

i i i i i i i<br />

Finally, fuel impact due to the variation <strong>of</strong> the final product can be separated into<br />

induced and intrinsic effects:<br />

n n<br />

oc ac fq sp r<br />

∑ P i( x ) ωi ∑ P i( x ) ( ωi ωi ωi ωi ωi ωi<br />

)<br />

MF = k ⋅Δ = k ⋅ Δ +Δ +Δ +Δ +Δ +<br />

* * 1 * 1 int<br />

0 , ,<br />

i= 1 i=<br />

1<br />

* *,int *, oc *, ac *, fq *, sp *, r<br />

0 0 0 0 0 0 0<br />

4.59<br />

MF = MF + MF + MF + MF + MF + MF<br />

4.60<br />

Where<br />

x<br />

MF *,<br />

0 can be the fuel impact caused by the variation <strong>of</strong> plant product due to<br />

intrinsic causes and causes induced by other components, ambient conditions, fuel<br />

quality and set points.<br />

Finally, the fuel impact can be expressed by a summation either <strong>of</strong> malfunction costs<br />

<strong>of</strong> each component or malfunction costs intrinsic and induced by other components,<br />

ambient conditions, fuel quality and set points:<br />

n n n<br />

* * *,int *, oc *, ac *, fq *, sp *, r<br />

∑ ∑ ∑ ( )<br />

Δ F = MF + MR = MF + MF + MF + MF + MF + MF +<br />

T i i i i i i i i<br />

i= 0 i= 1 i=<br />

0<br />

n<br />

i=<br />

1<br />

*,int *, oc *, ac *, fq *, sp *, r<br />

( MRi MRi MRi MRi MRi MRi<br />

)<br />

+ ∑ + + + + +<br />

4.61<br />

*,int *, oc *, ac *, fq *, sp *, r<br />

Δ FT= MF + MF + MF + MF + MF + MF +<br />

*,int *, oc *, ac *, fq *, sp *, r<br />

+ MR + MR + MR + MR + MR + MR<br />

4.62<br />

x<br />

Where MF *, *,x<br />

and MR are the fuel impacts due to effect intrinsic and induced by<br />

other components, ambient conditions, fuel quality and set points.<br />

MF<br />

*, x<br />

=<br />

n<br />

∑<br />

i=<br />

1<br />

MF<br />

*, x<br />

i<br />

n<br />

*, x *, x<br />

i<br />

i=<br />

1<br />

4.63<br />

MR = ∑ MR<br />

4.64


Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

As it can be seen, the separation <strong>of</strong> malfunctions into intrinsic and induced has been<br />

made in a parallel way as the separation <strong>of</strong> malfunctions into the different free <strong>diagnosis</strong><br />

variables. So that, result can also be summarized in a table <strong>of</strong> intrinsic and induced<br />

malfunctions (MFI). In this table, elements<br />

x<br />

MF i<br />

*, and<br />

*,x<br />

MRi are placed. Each column<br />

corresponds to a type <strong>of</strong> malfunction while each component has two rows: for<br />

malfunction associated with productive flows and with wastes. The summations <strong>of</strong><br />

columns are<br />

x<br />

MF *, *,x<br />

+ MR , and the summations <strong>of</strong> rows are elements<br />

Total summation is the fuel impact ΔFT.<br />

MF<br />

MF<br />

MF<br />

*, int<br />

0<br />

*, int<br />

1<br />

oc<br />

MF *,<br />

0<br />

oc<br />

MF *,<br />

1<br />

ac<br />

MF *,<br />

0<br />

ac<br />

MF *,<br />

1<br />

fq<br />

MF *,<br />

0<br />

fq<br />

MF *,<br />

1<br />

sp<br />

MF *,<br />

0<br />

sp<br />

MF *,<br />

1<br />

MF<br />

MF<br />

*, r<br />

0<br />

*, r<br />

1<br />

MF or MR .<br />

… … … … … … …<br />

*, int<br />

n<br />

*,int<br />

MR 1<br />

oc<br />

MF n<br />

*,<br />

MR<br />

*, oc<br />

1<br />

ac<br />

MF n<br />

*,<br />

MR<br />

*, ac<br />

1<br />

fq<br />

MF n<br />

*,<br />

MR<br />

*, fq<br />

1<br />

sp<br />

MF n<br />

*,<br />

MR<br />

*, sp<br />

1<br />

*,r<br />

MF n<br />

… … … … … … …<br />

*,int<br />

MR n<br />

*, int<br />

MF<br />

+ MR<br />

*,int<br />

*,oc<br />

MR n<br />

oc<br />

MF *,<br />

+ MR<br />

*,oc<br />

*,ac<br />

MR n<br />

ac<br />

MF *,<br />

+ MR<br />

*,ac<br />

*, fq<br />

MR n<br />

fq<br />

MF *,<br />

+ MR<br />

*, fq<br />

*,sp<br />

MR n<br />

sp<br />

MF *,<br />

+ MR<br />

*,sp<br />

MR<br />

*, r<br />

1<br />

*,r<br />

MR n<br />

*,r<br />

MF<br />

+ MR<br />

Table 4.2: Table <strong>of</strong> intrinsic and induced malfunctions (MFI).<br />

*,r<br />

*<br />

i<br />

*<br />

MF 0<br />

*<br />

MF 1<br />

*<br />

MF n<br />

*<br />

MR 1<br />

*<br />

MR n<br />

4.1.3. Determination <strong>of</strong> the suitability <strong>of</strong> a productive structure.<br />

The main source <strong>of</strong> inaccuracy in the application <strong>of</strong> thermoeconomic <strong>analysis</strong> to the<br />

<strong>diagnosis</strong> problem is due to the induced effects. To overcome it, several filtering<br />

strategies as well as variations <strong>of</strong> the productive structure have been proposed. The<br />

suitability <strong>of</strong> these techniques is usually determined by testing them in specific<br />

problems; so that, it is difficult to obtain general conclusions. In this way, when a new<br />

problem is tackled, it is very difficult to assure that a specific thermoeconomic model is<br />

suitable to diagnose it and to determine the error produced. Ideas presented in previous<br />

sections can help in this task. Equations developed to link the variations <strong>of</strong> free<br />

<strong>diagnosis</strong> variables and the variations <strong>of</strong> thermoeconomic parameters have been<br />

developed, but considering an example <strong>of</strong> <strong>diagnosis</strong> (a comparison <strong>of</strong> an actual and a<br />

Δ FT<br />

*<br />

i<br />

73


Chapter 4<br />

reference situation). So, they are not suitable to draw general conclusions about the<br />

connection <strong>of</strong> physical and thermoeconomic models. However, this general picture can<br />

be obtained by considering an artificial <strong>diagnosis</strong> example, where free <strong>diagnosis</strong><br />

variations have a representative value. This is the goal <strong>of</strong> this section.<br />

First <strong>of</strong> all, KD and TD matrices presented above can be very useful to determine the<br />

influence <strong>of</strong> induced effects in the unit exergy consumptions <strong>of</strong> a component due to:<br />

ambient conditions, fuel quality, set points and anomalies in other components. Each<br />

value<br />

74<br />

l<br />

l kd ij or td ij represents the sensibility <strong>of</strong> the unit exergy consumption κ ij or θ ij to<br />

the free <strong>diagnosis</strong> variable l. So that, the direct observation <strong>of</strong> KD l and TD l matrices<br />

allows to determine whether a variation in a free <strong>diagnosis</strong> variable induces a variation<br />

in an unit exergy consumption. However, the direct observation <strong>of</strong><br />

l kd and θ d values<br />

is not very useful for comparison because they are expressed in different units (free<br />

<strong>diagnosis</strong> variables have different units). This problem can be solved by multiplying<br />

them times ‘representative’ increments <strong>of</strong> the corresponding free <strong>diagnosis</strong> variables. If<br />

a large amount <strong>of</strong> plant data is available, a representative variation <strong>of</strong> the value <strong>of</strong> a<br />

given free <strong>diagnosis</strong> variable can be its standard deviation. If not, it can be calculated by<br />

dividing the difference <strong>of</strong> maximum minus minimum value into two.<br />

Δ x = σ Δ<br />

4.65<br />

rep<br />

di , xi<br />

Δ κ = kd ⋅Δ x<br />

4.66<br />

lrep ,<br />

l rep<br />

ij ij d , l<br />

rep<br />

ij<br />

nd l, rep<br />

ij<br />

nd<br />

l<br />

kdij rep<br />

xd<br />

, l<br />

l= 1 l=<br />

1<br />

∑ ∑ 4.67<br />

Δ κ = Δ κ = ⋅Δ<br />

Δ θ = td ⋅Δ x<br />

4.68<br />

lrep , l rep<br />

ij ij d , l<br />

rep<br />

ij<br />

nd l, rep<br />

ij<br />

nd<br />

l<br />

tdij rep<br />

xd<br />

, l<br />

l= 1 l=<br />

1<br />

∑ ∑ 4.69<br />

Δ θ = Δ θ = ⋅Δ<br />

Δ ω = wd ⋅Δ x<br />

4.70<br />

lrep ,<br />

l rep<br />

ij ij d , l<br />

rep<br />

ij<br />

nd l, rep<br />

ij<br />

nd<br />

l<br />

wdij rep<br />

xd<br />

, l<br />

l= 1 l=<br />

1<br />

∑ ∑ 4.71<br />

Δ ω = Δ ω = ⋅Δ<br />

It should be noted that the previous decomposition can be considered as exact<br />

(without the need <strong>of</strong> any residual term), because this is a fictitious situation. It is also<br />

possible to make the same decomposition into intrinsic variations (int) and variations<br />

induced by other components (oc), ambient conditions (ac), fuel quality (fq) and set<br />

points (sp):<br />

l<br />

ij<br />

ij


Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

rep<br />

nd<br />

l, rep int, rep oc, rep ac, rep fq, rep sp, rep<br />

ij<br />

l=<br />

1<br />

ij ij ij ij ij ij<br />

Δ κ = ∑ Δ κ =Δ κ +Δ κ +Δ κ +Δ κ +Δκ<br />

4.72<br />

rep<br />

nd<br />

l, rep int, rep oc, rep ac, rep fq, rep sp, rep<br />

ij<br />

l=<br />

1<br />

ij ij ij ij ij ij<br />

Δ θ = ∑ Δ θ =Δ θ +Δ θ +Δ θ +Δ θ +Δθ<br />

4.73<br />

rep<br />

nd<br />

l, rep int, rep oc, rep ac, rep fq, rep sp, rep<br />

i i i i i i i<br />

l=<br />

1<br />

Δ ω = ∑ Δ ω =Δ ω +Δ ω +Δ ω +Δ ω +Δω<br />

4.74<br />

By using the previous decompositions <strong>of</strong> κij, θij and ωi, it is possible to calculate<br />

representative values <strong>of</strong> malfunctions and malfunctions costs, by using expressions<br />

developed in section 4.1.2.<br />

4.1.4. Conclusion<br />

The aim <strong>of</strong> thermoeconomic <strong>diagnosis</strong> is to determine intrinsic malfunctions, so that,<br />

a good thermoeconomic model should be able to eliminate induced effects.<br />

Malfunctions induced by ambient conditions and fuel quality are usually not present<br />

because the reference condition is usually chosen with the same external conditions.<br />

Malfunctions induced by the set points are usually also eliminated, except in some<br />

approaches that are based in the filtration <strong>of</strong> effects induced by the control system<br />

(Verda et al., 2004). Finally, malfunctions induced by anomalies in other components<br />

are the most difficult to eliminate, so they constitute the main research topic in the<br />

application <strong>of</strong> the thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy systems. The<br />

formulation proposed here is able to quantify the importance <strong>of</strong> this induced terms, and<br />

thus to evaluate the ability <strong>of</strong> a given productive structure to reproduce the<br />

thermodynamic behaviour <strong>of</strong> the system. It is applied in Chapter 7.<br />

4.2. Application <strong>of</strong> neural networks and linear regression to<br />

the <strong>diagnosis</strong> <strong>of</strong> energy systems.<br />

Diagnosis <strong>methods</strong> previously presented are based on a representation <strong>of</strong> the thermal<br />

system analyzed by using either thermodynamic or thermoeconomic models, or the<br />

combination <strong>of</strong> both. However, when operation data are available, it is possible to use<br />

this information to build black-box models which can help to the <strong>diagnosis</strong> task. This<br />

idea is explored in this section and applied in Chapter 8 and in Section 6.6.<br />

75


Chapter 4<br />

First <strong>of</strong> all, fundamentals <strong>of</strong> linear regression and neural networks are sumarized.<br />

Afterwards, the possibilities <strong>of</strong> these tools to perform <strong>diagnosis</strong> <strong>of</strong> energy systems are<br />

presented.<br />

4.2.1. Linear regression fundamentals.<br />

In general, a linear regression model represents the relationship between a<br />

continuous response (y) and a continuous or categorical predictor x (x1,…xN) in the<br />

form:<br />

y = β ⋅ f ( x) + β ⋅ f ( x) + + β ⋅ f ( x ) + ε<br />

4.75<br />

76<br />

1 1 2 2 ... p p<br />

The response is modelled as a linear combination <strong>of</strong> (not necessarily linear)<br />

functions <strong>of</strong> the predictor, plus a random error ε. This error is assumed to be<br />

uncorrelated and distributed with mean 0 and constant (but unknown) variance. The<br />

expressions fj(x) (j=1,…,p) are the terms <strong>of</strong> the model. The βj (j=1,…,p) are the<br />

coefficients.<br />

The simplest examples <strong>of</strong> linear regression are linear additive models:<br />

y = β ⋅ x + β ⋅ x + + β ⋅ x + β + ε<br />

4.76<br />

1 1 2 2 ... N N N+<br />

1<br />

Given n independent observations (x 1 , y 1 ), (x 2 , y 2 ), …, (x n ,y n ) <strong>of</strong> the predictor x and<br />

the response y, the aim <strong>of</strong> regression is to estimate the model parameters βj (j=1,…,p).<br />

( ) ( )<br />

1<br />

1 1<br />

⎛ y ⎞ ⎛ f1x fpx ⎞ ⎛ β ⎞ 1 ⎛ε1⎞ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟<br />

⎜ ⎟= ⎜ ⎟⋅<br />

⎜ ⎟+ ⎜<br />

<br />

⎟<br />

⎜ n<br />

y ⎟ ⎜ n n ⎟<br />

⎜ f<br />

1 ( x ) fp( x<br />

⎜ ) ⎟ β ⎟ ⎜<br />

p ε ⎟<br />

⎝ ⎠ ⎝ n<br />

⎝<br />

<br />

⎠⎠<br />

⎝⎠ <br />

y<br />

ε<br />

X<br />

β<br />

4.77<br />

In general, n>p and the problem is overdetermined, so that coefficients ˆ<br />

β are<br />

calculated as estimators <strong>of</strong> β to minimise the sum <strong>of</strong> squares <strong>of</strong> the errors. ˆ<br />

β is<br />

unbiased with expected value β, and it has minimum variance among all unbiased<br />

estimators formed from linear combinations <strong>of</strong> the response data. Geometrically, to<br />

minimize the sum <strong>of</strong> squares <strong>of</strong> the errors,<br />

y−X ⋅ ˆ β must be perpendicular to the<br />

column space <strong>of</strong> X, which contains all the linear combinations <strong>of</strong> the model terms. This<br />

is summarized in the normal equations:<br />

( )<br />

t<br />

X ⋅ y−X ⋅ ˆ β =0<br />

4.78<br />

which allow to calculate ˆ<br />

β from X and y:


( ) 1<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

ˆ t<br />

−<br />

t<br />

β = X ⋅X ⋅X ⋅ y 4.79<br />

It should be noted that normal equations are <strong>of</strong>ten bad conditioned, so that it is usual<br />

not to solve them directly but by using a QR (orthogonal, triangular) decomposition <strong>of</strong><br />

X.<br />

The previous calculation <strong>of</strong> ˆ β as an estimator <strong>of</strong> β relies on the independence <strong>of</strong><br />

the model terms. When terms are correlated and the columns <strong>of</strong> the design matrix X<br />

have an approximate linear dependence, the matrix ( ) 1<br />

T<br />

−<br />

⋅<br />

X X becomes close to<br />

singular and the least squares estimate ˆ<br />

β becomes highly sensitive to random errors in<br />

the observed response y, producing a large variance. This situation <strong>of</strong> multicollinearity<br />

can arise, for example, when data are collected without an experimental design, which is<br />

quite usual in <strong>diagnosis</strong> <strong>of</strong> real operating systems. Multicollinearity can be detected by<br />

observing the correlation matrix R (p x p) defined as:<br />

R<br />

n<br />

l l<br />

∑(<br />

xi −xi) ⋅( xj −xj)<br />

l=<br />

1<br />

ij =<br />

n n<br />

1/2<br />

⎡ l<br />

2<br />

l<br />

2⎤<br />

⎢∑( xi −xi) ⋅∑( xj −xj)<br />

⎥<br />

⎣ l= 1 l=<br />

1 ⎦<br />

where x is the mean value:<br />

x<br />

i<br />

=<br />

4.80<br />

n<br />

∑<br />

l<br />

xi<br />

l=<br />

1<br />

4.81<br />

n<br />

R is symmetrical and contains ones in the diagonal. The other values vary between -<br />

1 and 1. An absolute value near to 1 outside the diagonal indicates high correlation<br />

between the variables.<br />

There are several <strong>methods</strong> to deal with multicollinearity. When a variable is highly<br />

linear dependent with others, it might be possible to eliminate it (assuming that<br />

information is lost) or substitute it by a combination <strong>of</strong> variables. Other option is the use<br />

<strong>of</strong> more advanced <strong>methods</strong> such as ridge regression, which are out <strong>of</strong> the scope <strong>of</strong> this<br />

work.<br />

For more details see MathWorks (2007), Montgomery and Peck (1992) and Jang et<br />

al. (1997).<br />

77


Chapter 4<br />

4.2.2. Neural network fundamentals.<br />

Neural networks are composed <strong>of</strong> simple elements (neurons) which operate in<br />

parallel. As in biological nervous systems, connections between neurons make possible<br />

to achieve complex behaviour <strong>of</strong> the whole system. A neural network can be trained to<br />

perform a function by adjusting the values <strong>of</strong> the connections (weights) between<br />

elements.<br />

4.2.2.1. Neurons, layers and networks.<br />

A neuron is represented in Figure 4.1. The input vector (with dimension r) is<br />

multiplied times the (single row) weight matrix W and then the bias (b) is added to form<br />

the value n:<br />

78<br />

n = W ⋅ p +b=w ⋅ p + w ⋅ p + ... + w ⋅ p + b<br />

4.82<br />

1,1 1 1,2 2 1, R R<br />

The value <strong>of</strong> n is modified by a transfer function f producing the output <strong>of</strong> the<br />

neuron (a).<br />

( ) ( )<br />

a = f n = f W ⋅ p + b<br />

4.83<br />

The behaviour <strong>of</strong> the neuron depends on the value <strong>of</strong> the parameters W and p, and <strong>of</strong><br />

the type <strong>of</strong> function f. The value <strong>of</strong> W and p for the different neurons in a given neural<br />

network are calculated in a training process. The transfer function f influences strongly<br />

the behaviour <strong>of</strong> the neurons.<br />

Figure 4.1. Single neuron. Source: Demuth and Beale (2002)<br />

There are several types <strong>of</strong> transfer functions. Figure 4.2. shows a hard-limit transfer<br />

function, which limits the output <strong>of</strong> the neuron to either 0, if n is less than 0; or 1, if n is<br />

greater than or equal to 0. This neuron allows to make classification decisions and is<br />

used in Perceptron networks.


Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

Figure 4.2. Hard-lim transfer function. Source: Demuth and Beale (2002)<br />

Figure 4.3 (Equation 4.84) represents a linear transfer function, which does not<br />

change the value <strong>of</strong> the input n. Figure 4.4 and 4.5 (and Equations 4.85 and 4.86)<br />

represent log-sigmoid and tan-sigmoid transfer functions, respectively.<br />

a = n<br />

4.84<br />

1<br />

a<br />

1 e − =<br />

+<br />

n<br />

Figure 4.3. Linear transfer function. Source: Demuth and Beale (2002)<br />

Figure 4.4. Log-sigmoid transfer function. Source: Demuth and Beale (2002)<br />

2<br />

a = −1<br />

2n<br />

( 1 e ) −<br />

+<br />

4.85<br />

4.86<br />

79


Chapter 4<br />

80<br />

Figure 4.5. Tan-sigmoid transfer function. Source: Demuth and Beale (2002)<br />

Two or more neurons can be combined in a layer. A layer with S neurons is<br />

represented in Figure 4.6. It should be noted that now W is a S x R matrix and b and a<br />

are S x 1 vectors.<br />

Figure 4.6. Layer <strong>of</strong> a neural network. Source: Demuth and Beale (2002)<br />

A network is formed by the connexion <strong>of</strong> several layers. There are several types <strong>of</strong><br />

networks according to the number, type and connexion <strong>of</strong> neurons and layers. It is also<br />

possible to include feedback or delays: a neural network containing one or more <strong>of</strong><br />

those elements is called dynamic; in other case it is called static.<br />

Once a network has been created (its structure has been defined), it has to be trained.<br />

Training consists in determining the value <strong>of</strong> weights and biases in order to obtain a<br />

network with a given behaviour by using a set <strong>of</strong> known inputs with its corresponding<br />

target outputs. In incremental training the weights and biases are updated each time an


Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

input is presented to the network. In batch training the weights and biases are only<br />

updated after all <strong>of</strong> the inputs are presented. It is more common to apply incremental<br />

training for dynamic networks and batch training to static networks, although other<br />

possibilities are also possible.<br />

4.2.2.2. Feedforward networks and backpropagation.<br />

A network composed <strong>of</strong> several layers connected in a way that the output <strong>of</strong> one is<br />

the input <strong>of</strong> the other is called feedforward; because information entering with the<br />

inputs is transformed in each layer and advances forward up to the network output. A<br />

feedforward network with a sigmoid layer and a linear output layer is capable <strong>of</strong><br />

approximating any function with a finite number <strong>of</strong> discontinuities.<br />

Feedforward networks are sometimes called backpropagation because this is a<br />

family <strong>of</strong> training algorithms used for them. The idea is to determine weight and biases<br />

to minimise the mean squared error between the network outputs and the target outputs.<br />

This is an exaple <strong>of</strong> batch training and is done iteratively. The simplest implementation<br />

<strong>of</strong> backpropagation learning updates the network weights and biases in the negative<br />

direction <strong>of</strong> the gradient. One iteration <strong>of</strong> this algorithm can be written as:<br />

x x α g 4.87<br />

k+ 1 k k k<br />

= − ⋅<br />

Where xk is a vector <strong>of</strong> current weights and biases, gk is the current gradient, and αk<br />

is the learning rate.<br />

Other more complex training algorithms have been developed in order to avoid<br />

problems appearing in different types <strong>of</strong> networks and to achieve faster convergence.<br />

For example, the Levenberg-Marquardt algorithm was designed to approach secondorder<br />

training speed without the need to compute the Hessian matrix. If the performance<br />

function has the form <strong>of</strong> a sum <strong>of</strong> squares, the Hessian matrix can be approximated as:<br />

t<br />

H ≅ J ⋅J<br />

4.88<br />

And the gradient can be computed as:<br />

t<br />

g = J ⋅e<br />

4.89<br />

Where J is the Jacobian matrix that contains first derivatives <strong>of</strong> the network errors<br />

with respect to the weights and biases, and e is a vector <strong>of</strong> network errors. One iteration<br />

<strong>of</strong> Levenberg-Marquardt algorithm can be written as:<br />

T<br />

1<br />

T<br />

k 1 k μ −<br />

x ⎡ ⎤<br />

+ = x −⎣J ⋅ J + I⎦ ⋅J ⋅e<br />

4.90<br />

81


Chapter 4<br />

μ is an scalar which can vary during the iterative process. Note that when μ is large,<br />

the method becomes gradient descendent with small step size.<br />

One <strong>of</strong> the problems that occur during neural network training is called overfitting.<br />

The error on the training set is driven to a very small value, but when new data is<br />

presented to the network the error is large. If the number <strong>of</strong> parameters in the network is<br />

much smaller than the total number <strong>of</strong> points in the training set, there is little chance <strong>of</strong><br />

overfitting. Fortunately, this is the common situation in <strong>diagnosis</strong> problems. In any case,<br />

it is advisable to use for training a fraction <strong>of</strong> the set <strong>of</strong> points available, and keeping the<br />

other points for checking.<br />

More details on neural networks can be seen in Demuth and Beale (2002) and in<br />

Jang et al. (1997).<br />

4.2.3. Application for <strong>diagnosis</strong>.<br />

Once the fundamentals <strong>of</strong> linear regression and neural networks have been<br />

presented, their capabilities for <strong>diagnosis</strong> are analyzed in this section. It should be noted<br />

that a sufficient high amount <strong>of</strong> plant data is needed to determine the coefficients <strong>of</strong> the<br />

regression model and the weights and biases <strong>of</strong> the network. It might be possible to<br />

obtain this information from simulation; but in most cases, if a simulator is available, it<br />

is more useful to use the simulator itself or quantitative <strong>causal</strong>ity <strong>analysis</strong> instead <strong>of</strong><br />

these <strong>methods</strong>.<br />

4.2.3.1. Diagnosis <strong>of</strong> the complete system.<br />

One possibility is to consider the whole system to be diagnosed as a black box, and<br />

to build a pure experimental model <strong>of</strong> it. Inputs <strong>of</strong> these models should be the free<br />

<strong>diagnosis</strong> variables and the output is the global efficiency indicator.<br />

If linear regression with linear additive models is used, coefficients β are directly the<br />

elements <strong>of</strong> vector ed. These factors are constant, so that this method is not able to deal<br />

with non-linear behaviour. It should be noted that this point could be solved by<br />

including quadratic terms, but this possibility is not considered in this work.<br />

When a complex system is analyzed as a whole, the problem <strong>of</strong> multicollinearity<br />

<strong>of</strong>ten appears. To avoid it, it is useful to perform a variable change in order to eliminate<br />

the most important dependences. If two variables xi and xj are correlated:<br />

82<br />

x ≅ a + a ⋅x<br />

4.91<br />

i i0ij j


The first <strong>of</strong> them is substituted by:<br />

i i 0i<br />

ij j<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong> and other <strong>diagnosis</strong> <strong>methods</strong><br />

xˆ = x −a −a ⋅x<br />

4.92<br />

Then, linear regression is applied to the new sets <strong>of</strong> variables (which now are<br />

independent). Finally, inverse variable change is used in order to obtain the impact<br />

factors corresponding to the initial variables, which is the objective <strong>of</strong> the procedure.<br />

If neural networks are used, it is advisable to use a feedforward net with two layers:<br />

one hidden layer with sigmoid neurons and one output layer with linear neurons. Since<br />

the functions to be approximated have few singularities, it is convenient to use few<br />

neurons in the hidden layer.<br />

Once the network is trained, impacts on the global efficiency indicator due to<br />

variations <strong>of</strong> free <strong>diagnosis</strong> variables can only be obtained by simulation with the<br />

network.<br />

Linear regression (with linear additive models) has the advantage <strong>of</strong> providing<br />

directly the impact factors ed. On the other hand, it is only suitable for linear relations<br />

and variable change is <strong>of</strong>ten needed. Neural networks are able to deal with non-linear<br />

dependences, but they do not provide information explicitly. It should be kept in mind<br />

that both techniques are based on the minimisation <strong>of</strong> the error in the determination <strong>of</strong><br />

the global efficiency indicator, so that they tend to quantify precisely only the effect <strong>of</strong><br />

variables with high impact. This effect can be seen in the practical application<br />

developed in Chapter 8.<br />

4.2.3.2. Diagnosis <strong>of</strong> sub-systems. Hybrid configurations.<br />

Sometimes, applicability <strong>of</strong> black box based approaches for complete systems with a<br />

lot <strong>of</strong> free <strong>diagnosis</strong> variables is limited because they are able to quantify precisely only<br />

those with high influence.<br />

One possiblity would be a tree approach. First <strong>of</strong> all, the whole system should be<br />

decomposed into subsystems, each one with several input variables (free <strong>diagnosis</strong><br />

variables) and one or more output variables. So, it is possible to build local models <strong>of</strong><br />

the subsystems. Afterwards, a global model has to be developed with the output<br />

variables <strong>of</strong> the local models (and, perhaps, some free <strong>diagnosis</strong> variables) as input<br />

variables, and the global efficiency indicator as the output variable. This approach can<br />

provide good results but suitable decomposition is needed, which it is not<br />

straightforward.<br />

83


Chapter 4<br />

Another option is to use a hybrid approach, combining quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong> and linear regression and/or neural networks. Causality chains theory provides<br />

the possibility to introduce local models on quantitative <strong>causal</strong>ity <strong>analysis</strong>. These<br />

models can be built by using linear regressions or neural networks. This approach can<br />

be very convenient because the use <strong>of</strong> experimental models is limited to those parts <strong>of</strong><br />

the <strong>analysis</strong> which are most difficult to study analytically. It is applied in Section 6.6 to<br />

filter the influence <strong>of</strong> ambient temperature in a cooling tower.<br />

4.2.3.3. Other issues.<br />

In Section 3.6 it has been introduced the idea <strong>of</strong> <strong>diagnosis</strong> <strong>of</strong> a time span, what needs<br />

a determination <strong>of</strong> the degradation along time. This evolution can be very complex, so<br />

that neural networks can be useful.<br />

4.2.4. Conclusion<br />

Fundamentals <strong>of</strong> linear regression and neural networks as well as some possibilities<br />

<strong>of</strong> application <strong>of</strong> them for <strong>diagnosis</strong> purposes have been presented in this section. A<br />

practical application can be seen in Chapter 8 and in Section 6.6.<br />

These approaches can be useful when plant information is available. They are able<br />

to quantify the main <strong>diagnosis</strong> causes <strong>of</strong> a complete system degradation. Besides, they<br />

can help to diagnose complex sub-systems and could be integrated in a <strong>diagnosis</strong><br />

structure based on quantitative <strong>causal</strong>ity <strong>analysis</strong> by using <strong>causal</strong>ity chains.<br />

84


5 Case study: 3x 350 MW coal-fired power<br />

plant.<br />

The methodology previously described has been applied to the <strong>diagnosis</strong> <strong>of</strong> a<br />

pulverized coal-fired power plant. Although <strong>diagnosis</strong> comprises the boiler, steam cycle<br />

and cooling system; the aim <strong>of</strong> the system is to diagnose the cycle and the main<br />

parameters <strong>of</strong> the boiler, so that complex issues related to heat transfer have been<br />

simplified.<br />

In the fist part, the plant is described. Then the plant components’ are reviewed,<br />

presenting the models and parameters proposed to analyze their behaviour as well as<br />

their degradation mechanisms, which are the aspects to be diagnosed, and the choice <strong>of</strong><br />

suitable <strong>diagnosis</strong> variables. Afterwards, available plant measurements are presented,<br />

which determines the depth <strong>of</strong> the achievable <strong>diagnosis</strong>; lack <strong>of</strong> information should be<br />

substituted by suitable models. According to the components’ characteristics and the<br />

available plant data, the list <strong>of</strong> <strong>diagnosis</strong> variables can be built.<br />

5.1 Teruel power plant description.<br />

The Teruel power plant has been chosen to apply the <strong>diagnosis</strong> methodologies<br />

presented in Chapter 3 and Chapter 4. It is a conventional pulverised coal-fired plant<br />

that consists <strong>of</strong> three units <strong>of</strong> 350 MW each one, built around 1980. It is located in<br />

Andorra, in the Teruel province (Spain) and owned by Endesa Generación S.A. In this<br />

section, the main characteristics <strong>of</strong> this power plant are presented: the cycle, the cooling<br />

system and the boiler.<br />

85


Chapter 5<br />

5.1.1 Steam cycle description<br />

The steam cycle <strong>of</strong> the plant was built by Mitsubishi and is depicted in Figure 5.1,<br />

along with the cooling system. Main steam leaving the boiler at 160 bar and 540 ºC<br />

enters the high pressure turbine where it is expanded down to 40 bar. Then it goes back<br />

to the boiler (actually to the reheater) in order to increase its temperature up to 540 º C<br />

again. Finally, it is expanded in two intermediate pressure turbines and in the low<br />

pressure turbine. These turbines are separated in two casings: one for high and medium<br />

pressure ones and the other for the low pressure. Due to the high volumetric flow, the<br />

low pressure turbine is double flow type, and has four stages (in each side). At the end<br />

<strong>of</strong> each turbine stage, part <strong>of</strong> the steam is extracted for the feedwater heaters, the<br />

deareator and the turbo-pump.<br />

Steam leaving the low pressure turbine is condensed in the condenser. Water leaving<br />

the condenser is pumped by an electric pump up to a pressure <strong>of</strong> about 8 bars. Then it<br />

goes trough two components <strong>of</strong> minor importance that are not depicted in the diagram:<br />

ejector and sealing steam condenser, where it is slightly heated. Afterwards, the feeding<br />

water is heated in the low pressure feedwater heaters, by using bleed steam from the low<br />

pressure turbine. These heaters are shell and tubes-type, with the water flowing inside<br />

the tubes and the steam condensing outside. Then, the water enters the deaerator, which<br />

is a direct contact heater (the fourth heater). Its aim is not only to heat the water but<br />

mainly to lead it to the saturation point, in order to release incondensable gases<br />

dissolved in it. Water leaving the deaerator is pumped again up to 180 bar. This second<br />

pump is not electrically-driven but it has its own steam turbine that uses bleed steam<br />

from the end <strong>of</strong> the second medium pressure turbine. Finally, the water is heated in two<br />

high pressure heaters (5 and 6). It should be noted that once it has condensed, the steam<br />

used in each heater is fed into the previous one, except the 4 th that does not have<br />

drainage and the 1 st that drains in the condenser.<br />

There are a lot <strong>of</strong> secondary flows that have not been included in figure 5.1, most <strong>of</strong><br />

them corresponding to the steam seals system. Besides, there are other flows entering<br />

the deaerator, such as the condensed water from the steam coil heaters.<br />

Some details on the cycle are included in Table 5.1.<br />

86


Case study: 3x350 MW coal-fired power plant<br />

RH<br />

HP IP 1 IP 2 LP<br />

COND<br />

CT<br />

TP<br />

1<br />

2<br />

3<br />

5<br />

6<br />

4<br />

BOILER<br />

Figure 5.1. Steam cycle.<br />

87


Chapter 5<br />

Parameter Characteristics<br />

Type Mitsubishi ‘Reheat Condensing Two Casing Double Flow’<br />

Inlet regulation by eight sequential valves<br />

Number <strong>of</strong> casings Three: high pressure (HP), intermediate pressure (IP 1 and 2) and low pressure<br />

(LP)<br />

Peak power 360 000 kW (maximum in generator)<br />

Nominal power 350 000 kW (optimum)<br />

Load levels 350 MW; 262.5 MW; 175 MW; 87.5 MW<br />

Rotation 3000 rpm, clockwise sense (from governing stage towards LP)<br />

Technical limits: 48.5 Hz < frequency < 50.5 Hz<br />

Cycle Water-steam Feeding water (boiler inlet):<br />

254 º C<br />

conditions<br />

Steam pressure prior to HP closing valve:<br />

162 kg/cm<br />

Steam temperature prior to HP closing valve:<br />

Steam temperature prior to IP closing valve:<br />

Absolute condenser pressure<br />

2<br />

538 ºC<br />

538 ºC<br />

0.069 kg/cm 2<br />

Steam extractions: 1<br />

maximum values and<br />

turbine section<br />

st :<br />

2 nd :<br />

3 rd :<br />

BFPT:<br />

4 th :<br />

5 th :<br />

6 th 26.2 kg/s, 354 ºC, 10<br />

18.6 kg/s, 445 ºC,<br />

17.2 kg/s, 318 ºC,<br />

11.9 kg/s, 318 ºC,<br />

15.4 kg/s, 207 ºC,<br />

9.94 kg/s, 94 ºC,<br />

:<br />

8.81 kg/s, 70 ºC,<br />

th stage,<br />

15 th stage,<br />

20 th stage,<br />

20 th stage,<br />

22 nd stage,<br />

24 th stage,<br />

25 th 10<br />

stage,<br />

th (HP)<br />

5 th (IP-1)<br />

10 th (IP-2)<br />

10 th (IP-2)<br />

2 nd (LP-1)<br />

4 th (LP-2)<br />

5 th (LP-3)<br />

Pressure losses Reheater<br />

< 10%<br />

Extractions (HP)<br />

< 3 %<br />

Other extractions<br />

< 5 %<br />

Stages in turbine High pressure<br />

Curtis double impulsion wheel<br />

Reaction: 9<br />

Intermediate pressure Reaction: 10<br />

Low pressure<br />

Reaction: 6 x 2 (double flow)<br />

Low pressure last Blade length<br />

850.9 mm<br />

wheel<br />

Pitch diameter<br />

2552.7 mm<br />

Peripheral velocity 534.0 m/s<br />

Mechanical losses < 1000 kW<br />

Radial clearance Blading fins: > 1.0 mm Gland fins: > 0.5 mm<br />

Generator Hydrogen cooling system<br />

0.987 design efficiency<br />

88<br />

Table 5.1. Main characteristics <strong>of</strong> the Teruel Power plant steam cycle. (Zaleta, 1997).


Case study: 3x350 MW coal-fired power plant<br />

5.1.2 Cooling system<br />

The cooling system is a closed loop type, based on a natural draft cooling tower. It is<br />

composed <strong>of</strong> condenser, cooling tower, pumps and water ducts.<br />

The condenser is located just below the low pressure turbine. Steam leaving the<br />

pressure goes downwards becoming condensed by releasing heat to the cooling water<br />

that flows inside tubes that cross the condenser. As it is usual when cooling towers are<br />

used, the condenser has two pass, which means that cooling water crosses it twice.<br />

Some design characteristics <strong>of</strong> the condenser are summarized in table 5.2.<br />

Condenser pressure 0.069 kg/cm 2 absolute<br />

Steam flow rate 202.8 kg/s<br />

Cooling water flow rate 9.78 m 3 /s<br />

Inlet cooling water 24.4 ºC<br />

Outlet cooling water 35.3 ºC<br />

Number <strong>of</strong> tubes 23700<br />

Cooling surface 17650 m 2<br />

Table 5.2. Condenser characteristics (Lazaro, 2001)<br />

Heated cooling water leaving the condenser goes to a cooling tower through a duct<br />

<strong>of</strong> 2.5 meters <strong>of</strong> diameter. There, water is dispersed over a fill media located in order to<br />

favour the matter and heat transfer between the water going downwards and the air<br />

going upwards. The evaporation <strong>of</strong> part <strong>of</strong> the water causes the cooling <strong>of</strong> the rest <strong>of</strong> the<br />

water. The air flow is originated by density difference (natural-draft type). Some<br />

characteristics <strong>of</strong> the tower appear in Table 5.3.<br />

Finally, water from the bottom <strong>of</strong> the tower is again pumped towards the condenser.<br />

Water is added in order to compensate the amount evaporated.<br />

89


Chapter 5<br />

90<br />

Cooling water flow rate 10.56 m 3 /s<br />

Hot water temperature 35.3 ºC<br />

Cold water temperature 24.4 ºC<br />

Ambient wet bulb temperature 12.9 ºC<br />

Relative humidity 79 %<br />

Tower height 107.3 m<br />

Bottom diameter 81.22 m<br />

Top diameter 50.70 m<br />

Fill volume 84000 m 3<br />

Table 5.3. Cooling water design characteristics (Lazaro, 2001).<br />

5.1.3 Boiler<br />

Boilers at Teruel power plant are pulverized coal-fired, front-fired, natural<br />

circulation and balanced draft. They have three superheating stages and one reheating<br />

stage. The main nominal characteristics <strong>of</strong> each one <strong>of</strong> them are summarized in table<br />

5.4.<br />

Main steam flow rate 302.8 kg/s<br />

Main steam temperature 540 ºC<br />

Main steam pressure 169 kg/cm 2<br />

Reheated steam flow rate 266.7 kg/s<br />

Reheated steam temperature 540 ºC<br />

Reheated steam pressure 40 kg/cm 2<br />

Table 5.4. Teruel power plant boilers, nominal production (Díez, 2002).<br />

Air-gases circuit is composed <strong>of</strong> two parallel lines, called ‘side A’ and ‘side B’.<br />

Each one <strong>of</strong> them is depicted in Figure 5.2. Air entering the boiler is pressurized by the<br />

forced draft fan (FDF) and then is divided into two flows: secondary and primary air.<br />

Secondary air goes to the secondary air preheater (SAH) where it increases its<br />

temperature by using sensible heat from the flue gases leaving the boiler. These heaters<br />

are regenerative-type, which means that they work based on heating and cooling <strong>of</strong><br />

heating elements, due to an alternative contact with hot and cold flow. There are two


Case study: 3x350 MW coal-fired power plant<br />

ways to achieve this: by rotation <strong>of</strong> the basket where heating elements are located<br />

(Ljungstrom), or by rotation <strong>of</strong> the bells that conduct the air and gas flows through fixed<br />

baskets (Rothemuhle). In 1999, Rothemuhle type secondary air preheaters at Teruel<br />

power plant were substituted by Ljungstrom type. One important consequence <strong>of</strong> a<br />

working principle based on moving parts is the presence <strong>of</strong> air leakages, which<br />

constitute an important efficiency reduction cause. Heated secondary air goes to the<br />

wind box and then to the furnace through tangential holes located at the burners.<br />

TO BURNERS<br />

TO FGD<br />

TO MILLS<br />

FROM EZ1 FROM BY-PASS<br />

SAH<br />

EP<br />

IDF<br />

PAH<br />

SCH1<br />

Figure 5.2. Fans and preheaters.<br />

Pressure <strong>of</strong> primary air is further increased by the primary air fan (PAF). Then, this<br />

flow is heated by the primary air heaters (PAH). These preheaters were originally<br />

regenerative-type but in 1999 they were substituted by tubular-recuperative heaters. In<br />

these heat exchangers, hot flue gases flow inside tubes that are surrounded by the cold<br />

air. They do not present leakages, or at least they can be neglected compared to the<br />

recuperative-type ones. Heated primary air goes to the mills where it is used to dry and<br />

transport the pulverized coal.<br />

Both primary and secondary air pre-heaters can be damaged due to acid<br />

condensation if surface temperature is too low. To prevent this situation, manufacturer<br />

advises to keep the average cold side temperatures (mean <strong>of</strong> air inlet and flue gases<br />

outlet temperatures) above a certain limit (about 110 ºC). To assure that this condition is<br />

fulfilled also in cold weather, steam coil air heaters are located before the pre-heaters in<br />

SCH2<br />

PAF<br />

FDF<br />

91


Chapter 5<br />

both primary and secondary air flows. These heaters work by condensing steam<br />

extracted from the drum, whose pressure has been previously reduced to 16 bar.<br />

A part <strong>of</strong> the primary air is devised after the primary air fan an goes to the mill<br />

without being heated. This tempering air is used to control the temperature <strong>of</strong> primary<br />

air entering the mills. There is also a derivation <strong>of</strong> heated primary air that can be fed in<br />

the secondary air flow before the preheater, in order to increase its temperature without<br />

using steam in the coil heaters. Finally, it is possible to extract hot air from both primary<br />

and secondary flows to heat-up flue gases leaving the sulphur removal unit in order to<br />

get suitable stack entering temperature. It should be noted that only primary air is<br />

usually used for this issue.<br />

Figure 5.3 shows a schematic side view <strong>of</strong> the boiler, excluded the air pre-heaters<br />

area previously commented. There are 6 mills, each one corresponding to a row <strong>of</strong><br />

burners and named with a letter from A to F. The flow <strong>of</strong> primary air and pulverised<br />

coal goes from each mill to the four burners <strong>of</strong> each row where secondary air is also<br />

introduced and coal is burnt. Sometimes, natural gas is also burnt to assure good flame<br />

characteristics in part load or start-up processes.<br />

The union <strong>of</strong> up to 24 flames forms a fire ball where a great amount <strong>of</strong> heat is<br />

released, mainly by radiation. This radiation is mainly used to boil the water in the walls<br />

<strong>of</strong> the furnace: water goes downwards from the drum through external tubes and<br />

evaporates while going upwards back to the drum through tubes covering all the furnace<br />

walls. Density difference <strong>of</strong> water and steam is enough to assure enough flow rate<br />

(natural circulation). Furnace radiation also contributes notably to heat transfer in the<br />

second superheater (also called radiative superheater).<br />

Hot gases leaving the furnace are derived by the nose and goes trough the secondary<br />

and tertiary superheaters. Then, they arrive to the plenum and are separated in two<br />

flows: one flow going to the reheater and one part <strong>of</strong> the first economizer and other<br />

goint through primary superheater, second economizer and the second part <strong>of</strong> the first<br />

economizer. This flow distribution is controlled by dampers and is adjusted in order to<br />

maintain the reheated temperature set-point. Heat exchanger surfaces are cleaned by a<br />

sootblowing system that uses steam taken from the steam cycle in the boiler, before the<br />

final superheater. Exchange surface <strong>of</strong> heat exchangers located in the boiler are<br />

summarized in table 5.5.<br />

92


D<br />

C<br />

E<br />

B<br />

F<br />

A<br />

SH 2<br />

FURNACE<br />

Case study: 3x350 MW coal-fired power plant<br />

SH 3<br />

SH 1<br />

EZ 2<br />

Figure 5.3. Schematic side view <strong>of</strong> the boiler.<br />

EZ 1<br />

RH<br />

93


Chapter 5<br />

94<br />

Heat exchanger Exchange area (m 2 )<br />

Primary economizer (EZ1) 3360<br />

Secondary economizer (EZ2) 1970<br />

Primary superheater (SH1) 2545<br />

Secondary superheater (SH2) 1296<br />

Final superheater (SH3) 1668<br />

Reheater (RH) 8440<br />

Table 5.5. Exchange area <strong>of</strong> the boiler heat exchangers. (Díez, 2002)<br />

Boiler can also be explained from the water-steam point <strong>of</strong> view. Water coming<br />

from the cycle (actually from the 6 th feedwater heater), is first heated in the primary and<br />

secondary economizers and goes into the drum. As it has been previously said, there is a<br />

secondary water-steam loop in which water goes downwards from the drum and goes<br />

back to it after being evaporated in the furnace walls. Dry steam is taken from the upper<br />

part <strong>of</strong> the drum and goes to the plenum, boiler ceiling and division walls <strong>of</strong> the<br />

convective area where increases its temperature. Afterwards, it goes to the primary,<br />

radiant and final superheaters, where it reaches the suitable temperature <strong>of</strong> 540 ºC. This<br />

set-point, as well as the temperature <strong>of</strong> steam exiting the radiant superheater, are<br />

controlled by introducing water at the entering <strong>of</strong> the final and radiant superheaters,<br />

respectively. This water is taken from the impulsion <strong>of</strong> the turbo-pump (see figure 5.1).<br />

Flue gases flow leaving the first economizer goes to the secondary and primary air<br />

preheaters (Figure 5.2). Part <strong>of</strong> flue gases leaving the reheater can be deviated directly<br />

to the primary air preheater in order to increase the temperature <strong>of</strong> gases entering this<br />

heat exchanger (by-pass <strong>of</strong> economizer). Gases leaving the primary and secondary air<br />

preheaters go to the electrostatic precipitator (EP) that removes the particulates that the<br />

gas flow has, and to the induced draft fan (IDF) which moves them towards the sulphur<br />

removing unit and the stack.<br />

More details on this boiler can be seen in Diez (2002).


Case study: 3x350 MW coal-fired power plant<br />

5.2 Diagnosis model for a pulverized-coal power plant.<br />

A main task <strong>of</strong> this section is to review the models and parameters available to<br />

describe the behaviour <strong>of</strong> the elements constituting a conventional coal-fired power<br />

plant. The problem <strong>of</strong> modelling and the choice <strong>of</strong> suitable <strong>diagnosis</strong> variables are<br />

related topics: a <strong>diagnosis</strong> variable close to the observable behaviour <strong>of</strong> a system that<br />

can be calculated from data directly measured or obtained by a performance test does<br />

not need any modelling; on the other hand, it is possible to choose a <strong>diagnosis</strong> variable<br />

closer to the physical state and degradation level <strong>of</strong> the system, which probably will<br />

need a more or less complex model to be related to the observable parameters. In<br />

general, the previous option will be preferred, and the influence <strong>of</strong> the operation point<br />

on the variable will be filtered by using <strong>methods</strong> such as anamnesis (repeated <strong>diagnosis</strong><br />

along time) or the use <strong>of</strong> <strong>causal</strong>ity chains which allow to integrate crossed dependence<br />

and models.<br />

Besides, modelling is sometimes needed to include in the performance tests flows<br />

and systems for which measurement is not available (e.g. seals flows). Of course, it may<br />

be possible to include simulation as an additional capability for what-if analyses <strong>of</strong> a<br />

<strong>diagnosis</strong> system installed in a power plant, by using almost the same set <strong>of</strong> equations <strong>of</strong><br />

the <strong>diagnosis</strong> but substituting identities corresponding to <strong>diagnosis</strong> variables by a<br />

suitable model.<br />

Another important issue tackled in this section is a review <strong>of</strong> the main anomalies or<br />

causes <strong>of</strong> efficiency decrease <strong>of</strong> the different plant components. Although the <strong>diagnosis</strong><br />

system is based on a thermodynamic representation <strong>of</strong> the system and the<br />

characterization <strong>of</strong> its components by using parameters relating pressures, temperatures<br />

and so on, it should be kept in mind that the final goal is to detect physical degradations<br />

in components and to quantify their effects on the global system.<br />

These two aspects, modelling and <strong>diagnosis</strong> variables, and degradation mechanisms<br />

are studied for the main plant components, ordered by plant zones: steam cycle, cooling<br />

system and boiler. The idea is to start with the general ideas and then to descend to the<br />

case <strong>of</strong> study. However, the most specific points on the implementation at Teruel power<br />

plant are summarized later, after having presented the available plant data.<br />

95


Chapter 5<br />

5.2.1 Steam cycle<br />

The main components <strong>of</strong> this area are steam turbines. Then, the <strong>diagnosis</strong> model for<br />

regenerative water heaters is explained. Finally, other components such as pumps,<br />

generator and steam seals are considered.<br />

5.2.1.1 Steam turbine<br />

A detailed study <strong>of</strong> the steam turbine would strictly require an aerodynamic model<br />

<strong>of</strong> each stage based on the speed triangle (see Salisbury, 1974). However, this approach<br />

is neither practical (due to the lack <strong>of</strong> information) nor necessary, because for the<br />

purpose <strong>of</strong> this work, turbines can be considered as black boxes. So that, only two<br />

relations are needed for each turbine section. In the case study there are seven turbine<br />

sections (one high pressure, two intermediate pressure and four low pressure), because<br />

the two flows <strong>of</strong> the low pressure are symmetrical and considered as only one. These<br />

two parameters are isoentropic efficiency and flow coefficient.<br />

Isoentropic efficiency <strong>of</strong> a turbine is defined as the quotient between the enthalpy<br />

drop and the enthalpy drop to the same isobar if the expansion were isoentropic.<br />

96<br />

η<br />

h − h<br />

=<br />

in out<br />

st ,<br />

hin−hout, s<br />

Where hin and hout are the enthalpy <strong>of</strong> steam entering and exiting the turbine and<br />

hout,s is the enthalpy <strong>of</strong> a fictitious point that has the same entropy <strong>of</strong> the inlet and the<br />

same pressure <strong>of</strong> the outlet: hout,s = h(sin, pout). Once known the inlet conditions and the<br />

discharge pressure <strong>of</strong> a turbine section, isoentropic efficiency allows to determine the<br />

temperature <strong>of</strong> the exiting flow (or its quality, if it is located inside the vapor dome).<br />

Spencer et al. (1974), provide a methodology to calculate the isoentropic efficiency<br />

<strong>of</strong> steam turbines <strong>of</strong> above 16500 kW working with or without reheating. These<br />

relations are essentially empirical, but they are supported by several cases <strong>of</strong> practical<br />

application. For a steam cycle with reheating, these authors propose the use <strong>of</strong> two<br />

expansion lines. The first one corresponds to the regulation stage and the high pressure<br />

turbine. The isoentropic efficiency <strong>of</strong> this section is calculated by considering a<br />

reference value that depends on the turbine characteristics and two correction factors<br />

depending on the volumetric flows and the pressure ratio.<br />

The second expansion line comprises the two intermediate pressure and the low<br />

pressure turbine. It is calculated by modifying a base value by using three correcting<br />

factors related to volumetric flow and temperature and pressure in the inlet <strong>of</strong> this<br />

5.1


Case study: 3x350 MW coal-fired power plant<br />

section. This expansion line should be corrected for the last section, taking into account<br />

the exhaust losses originated by the kinetic energy <strong>of</strong> the steam leaving the turbine. This<br />

correction depends on volumetric flow and quality <strong>of</strong> steam leaving the turbine.<br />

It may be possible to choose the reference values <strong>of</strong> these correlations as free<br />

<strong>diagnosis</strong> variables. They would be quite close to physical behaviour <strong>of</strong> turbines and<br />

might be more suitable to filter malfunctions induced by variable operation points.<br />

However, they would be more difficult to calculate and to interpret, because they could<br />

not be directly calculated from properties <strong>of</strong> fluids entering and exiting the equipment.<br />

The other parameter needed to determine the steam turbines behaviour is the flow<br />

coefficient, which is defined as:<br />

m<br />

φ =<br />

5.2<br />

pin<br />

υ<br />

in<br />

where m is the flow rate and pin and υ in are the pressure and specific volume <strong>of</strong><br />

steam entering the turbine stage. Under the hypothesis <strong>of</strong> ideal gas, the previous<br />

equation can be expressed as:<br />

m⋅T φ =<br />

p<br />

<br />

in<br />

in<br />

(ideal gas) 5.3<br />

where Tin is the temperature <strong>of</strong> steam entering the turbine. According to the Stodola<br />

ellipse applied to a group <strong>of</strong> stages (see Cooke, 1985), for very high pressure ratio sonic<br />

flow is reached and the flow coefficient is constant. In general, actual and design flow<br />

coefficients are related by:<br />

φ = φ ⋅<br />

d<br />

1−<br />

r<br />

1−<br />

r<br />

2<br />

p<br />

2<br />

pd ,<br />

where the sub-index d means design conditions and rp is the pressure relation:<br />

r<br />

p<br />

out<br />

p = 5.5<br />

pin<br />

By using these equations, it is possible to determine the value <strong>of</strong> the pressure along<br />

all the expansion line, except in two points: the condenser and the inlet <strong>of</strong> the high<br />

pressure turbine. Stodola ellipse refers to turbine inlet, so that it cannot be applied to the<br />

condenser, whose pressure appears as an equilibrium between the heat to be removed to<br />

condense the steam leaving the turbine and the capacity <strong>of</strong> evacuate this heat, which is<br />

5.4<br />

97


Chapter 5<br />

determined by the cooling system (condenser, pumps and cooling tower) and mainly by<br />

the ambient conditions (temperature and relative humidity). If the high pressure turbine<br />

would operate at sliding pressure, the Stodola ellipse should be applied. However, it<br />

operates at constant pressure that is maintained by opening and closing the regulation<br />

valves. This variation <strong>of</strong> the regulation system makes the model useless, because steam<br />

pressure is directly a set-point.<br />

Zaleta (1997) shows a review <strong>of</strong> the causes <strong>of</strong> turbine degradation, which he calls<br />

intrinsic malfunction. They can be classified into i) erosion and breakings, ii) sediments,<br />

iii) rugosity, iv) damage in steam way and seals and v) control valves.<br />

Erosion and breakings is one <strong>of</strong> the most common malfunctions, and it is mainly<br />

due to the presence <strong>of</strong> solid particles in the steam. Erosion is a function <strong>of</strong> particle size,<br />

impact angle and particle speed. Breaking is similar to erosion in the sense that it entails<br />

a lost <strong>of</strong> material. However its effect is more severe and can produce high vibrations and<br />

lead to turbine shut-<strong>of</strong>f.<br />

Effects produced by erosion are: i) increment <strong>of</strong> inlet areas (and reduction <strong>of</strong> inlet<br />

pressure), ii) changes in incidence angles and iii) efficiency losses and changes in steam<br />

properties. Corrective <strong>methods</strong> (after erosion) are very costly and <strong>of</strong>ten entail the<br />

replacement <strong>of</strong> wheels <strong>of</strong> blades. For damages not very severe, material injection by<br />

using spray plasma technology may be used. Some preventive <strong>methods</strong> are the use <strong>of</strong><br />

filters to remove solid particulates, special water treatment and the use <strong>of</strong> anti-oxidant<br />

materials.<br />

Sediments are caused by inadequate water treatment and for high particulate content<br />

in water. They occur usually in turbine areas where speed is relatively low. Sediments<br />

produce increment in rugosity and reduction in effective areas. Preventive <strong>methods</strong> are<br />

the same used for avoiding erosions. Corrective <strong>methods</strong> use conventional cleaning<br />

processes based on air jet and particulates.<br />

Some parts <strong>of</strong> a turbine usually have a special surface treatment in order to assure<br />

low rugosity. However, due to the flow <strong>of</strong> steam and particulates, surfaces increase its<br />

rugosity and losses increase.<br />

Damage in steam way is caused by an excess <strong>of</strong> rugosity and the erosion <strong>of</strong> the<br />

initial and final parts <strong>of</strong> the blade: ‘Leading Edge Erosion’ and ‘Trailing Edge Erosion’<br />

Besides, a damage in seals will produce a derivation <strong>of</strong> a fraction <strong>of</strong> steam (‘By-pass<br />

leakage’). Seals are studied later.<br />

98


Case study: 3x350 MW coal-fired power plant<br />

Control valves, when are completely open, have pressure losses <strong>of</strong> 1-2 %. However,<br />

this figure can also be substantially increased due to erosion and sediments. Besides,<br />

damages in valve seats origin steam leakages.<br />

Along with these intrinsic malfunctions, Zaleta (1997) refers to induced<br />

malfunctions: changes in admission temperature, outlet pressure and extraction flow<br />

rate.<br />

5.2.1.2 Steam turbine seals and other secondary flows.<br />

The presence <strong>of</strong> holes to hold the turbine axis connecting several turbine sections at<br />

different pressure among themselves and with the atmosphere requires the use <strong>of</strong> a<br />

sealing system to avoid steam leakages and, in the low pressure sections, air infiltration.<br />

Sealing in steam turbines is achieved by using labyrinth seals through which a small<br />

amount <strong>of</strong> steam flows in a controlled manner. In the middle <strong>of</strong> these seals, chambers<br />

working at suitable pressure are located in order to keep reduced these leakages.<br />

Pressure <strong>of</strong> these chambers are controlled by quite complex systems that also comprise<br />

a device called seals steam regulator and a seals steam condenser, as well as a grid <strong>of</strong><br />

ducts that connects suitably all these parts. In this condenser, steam containing<br />

infiltrated air is condensed while this air is released.<br />

To calculate the amount <strong>of</strong> steam flowing through each seal, the Martin equation can<br />

be applied (Salisbury, 1974):<br />

p<br />

in<br />

i = ⋅ t ⋅ ⋅<br />

5.6<br />

ν in<br />

m k A β<br />

Where p and υ are pressure and specific volume <strong>of</strong> the flow entering the seal, At is<br />

the area <strong>of</strong> the clearance and β is a geometric factor depending on the teeth number <strong>of</strong><br />

the seal (z):<br />

β =<br />

2<br />

1−<br />

rp<br />

z−ln r<br />

p<br />

k is a constant determined by the shape <strong>of</strong> teeth and rp is the pressure ratio:<br />

r<br />

p<br />

out<br />

p = 5.5<br />

pin<br />

If geometric information is not available, it is possible to adjust a constant<br />

comprising k·At·β by using plant design information, neglecting the small variation <strong>of</strong><br />

pressure ratio. Hussain (1984) proposes another expression to be used when critical<br />

conditions are not reached:<br />

5.7<br />

99


Chapter 5<br />

100<br />

2 2<br />

pin − pout<br />

m i = At<br />

⋅<br />

5.8<br />

z⋅p ⋅ν<br />

in in<br />

In the case <strong>of</strong> study, not only leakages had to be calculated, but also flows<br />

connecting turbine seals, seal steam regulator and seal steam condenser. So that design<br />

information available for 100, 75 and 50 % load (Auxiesa, 1980) has been used to adjust<br />

second order polynomials that relate these flows to the load or to the nearest main flow.<br />

As it has been commented previously, damages in seals can originate an increment<br />

<strong>of</strong> leakages that implies additional losses. However, these flows are not measured and<br />

are not considered as <strong>diagnosis</strong> variables.<br />

The same approach (use <strong>of</strong> polynomials adjusted to design and not consideration as<br />

<strong>diagnosis</strong> variable) is applied to other secondary flows <strong>of</strong> the cycle: steam flow rate to<br />

ejector and seal flows in turbo-pump.<br />

5.2.1.3 Alternator<br />

Walsh and Fletcher (1998) propose curves to determine the efficiency <strong>of</strong> a generator<br />

depending on its active power and its power factor. These curves can be summarized in<br />

the following expression, which relates the actual to the maximum efficiency:<br />

k2+ k3⋅W gen<br />

( 1 ( 0.85) ) ( 1 )<br />

ηgen = ηgen<br />

max ⋅ + k ⋅ PF − ⋅ −e<br />

, 1<br />

where PF is the power factor, Wgen is the electric power and k1, k2 and k3 are<br />

constants. In the case <strong>of</strong> study, only information for three load levels were available, so<br />

that, a second order polynomial relation has been adjusted, considering not only the<br />

generator efficiency but also mechanical losses in the shaft. Since mechanical power<br />

provided to the generator cannot be measured nor determined, alternator efficiency is<br />

not considered as a <strong>diagnosis</strong> variable.<br />

5.2.1.4 Feedwater heaters<br />

The aim <strong>of</strong> these components is to increase the temperature <strong>of</strong> water before it enters<br />

the boiler by condensing steam that is extracted from the turbines at several pressures<br />

(bleed steam). These heaters are shell and tube type; water to be heated flows inside the<br />

tubes while steam is cooled down to the saturation temperature, condensed, and finally<br />

it is sub cooled. This process is shown in Figure 5.4. It should be noted that in the first<br />

heaters, corresponding to the last turbine stages, steam extracted is saturated, so that the<br />

first stage <strong>of</strong> steam cooling does not take place.<br />

5.9


T<br />

TTD<br />

Case study: 3x350 MW coal-fired power plant<br />

TDCA<br />

Figure 5.4. Temperature pr<strong>of</strong>ile in a feedwater heater.<br />

Thermal behaviour <strong>of</strong> regenerative heaters is characterized by two parameters:<br />

terminal temperature difference (TTD) and drain cooling approach temperature<br />

(TDCA). TTD is the difference between the saturation temperature <strong>of</strong> the extracted<br />

steam and the temperature <strong>of</strong> water leaving the heater. It should be noted that it can be<br />

positive, zero or negative, except when steam is saturated, and it can be only positive.<br />

TDCA is the difference between the temperatures <strong>of</strong> water entering the heater and<br />

drainage (water resulting from the condensing <strong>of</strong> steam). This value can only be<br />

positive.<br />

To model these heat exchangers, it is possible to use the classical approach based on<br />

the global heat transfer coefficient U, combined with the use <strong>of</strong> <strong>methods</strong> such as ε-NTU<br />

or Log-mean temperature difference (LMTD) (Mills, 1992). The problem with these<br />

approaches is that U coefficient varies in the three zones <strong>of</strong> the exchanger: steam<br />

cooling, condensing and subcooling.<br />

In this work, TTD and TDCA have been chosen as <strong>diagnosis</strong> variables. They can<br />

also be supposed constants for simulation purposes.<br />

Anomalies that may appear in these heaters are mainly fouling and duct blocking,<br />

which originate a decrease in the heat transfer capability and an increase in the value <strong>of</strong><br />

TTD. Sometimes, due to the necessity <strong>of</strong> plant operation and maintenance, a heater<br />

should be by-passed. This means that the equipment is isolated by closing the steam<br />

inlet and drainage and by connecting directly water inlet and outlet. This situation is<br />

correctly diagnosed by the TTD parameters, which increases dramatically because water<br />

at the exit has the same temperature as at the inlet.<br />

x<br />

101


Chapter 5<br />

Deaerator can also be considered as a water heater, because it heats the water.<br />

However, its main purpose is to release gases dissolved in the water. For this reason,<br />

water and steam flows are not separated but they are mixed, so that no drainage is<br />

present. Besides, water leaves the deaerator as saturated liquid (and at the extraction<br />

pressure), so that TTD equals zero. Due to these fixed operation conditions determined<br />

for the deaerator working principle, no <strong>diagnosis</strong> variable is defined for this component.<br />

5.2.1.5 Pumps<br />

Pump manufacturers provide curves that relate the pressure increment provided by<br />

the pump, and its volumetric efficiency to the flow rate that is pumped:<br />

102<br />

out in Δp<br />

( )<br />

p − p =Δ p= f m<br />

5.10<br />

( )<br />

η v = fη m<br />

5.11<br />

The first curve, called characteristic curve <strong>of</strong> the pump, should be compared with the<br />

characteristic curve <strong>of</strong> the installation to determine the operation point <strong>of</strong> the system<br />

(flow rate and pressure increment). Power required by the pump can be calculated by<br />

the following expression:<br />

m ν p<br />

W<br />

⋅ ⋅Δ<br />

=<br />

5.12<br />

η ⋅η<br />

mec v<br />

Where W is the mechanical power required by the pump, ν is the specific volume <strong>of</strong><br />

the liquid, ηmec is the mechanical efficiency and ηv is volumetric efficiency. Equations<br />

presented above constitute the model <strong>of</strong> a pump working at constant speed. To simulate<br />

a pump working at different speed (e. g. when a variable speed hydraulic coupling<br />

system is used), relations provided by dimensional <strong>analysis</strong> can be used. According this,<br />

flow rate relation is proportional to speed relation, pressure increment is proportional to<br />

the square <strong>of</strong> speed relation, and power varies with the cube:<br />

m n n<br />

1 1 =<br />

5.13<br />

m n<br />

n2<br />

2<br />

2<br />

n1<br />

= ⎜<br />

1<br />

⎟<br />

n2<br />

2<br />

Δp ⎛ n ⎞<br />

Δp ⎝n⎠ W n<br />

W n<br />

3<br />

n ⎛ ⎞<br />

1 1 = ⎜ ⎟<br />

n2<br />

⎝ 2 ⎠<br />

5.14<br />

5.15


Case study: 3x350 MW coal-fired power plant<br />

The parameter suitable to diagnose the pumps depends on the type <strong>of</strong> pump and<br />

plant measurement available. In the case <strong>of</strong> electric pumps, a good option is to use a<br />

global efficiency comprising electric, mechanical and volumetric efficiencies:<br />

m⋅ν⋅Δp ηp = ηe⋅ηmec⋅ ηv<br />

=<br />

W<br />

<br />

e<br />

5.16<br />

where ηe and We are respectively the efficiency <strong>of</strong> the electric motor and the<br />

electric power consumed by it.<br />

Big pumps for feeding water in power plants are usually driven by a small turbine<br />

fed with extraction steam, coupled to a hydraulic transmission. The group <strong>of</strong> turbine,<br />

transmission and pump is called turbo-pump. This system can be diagnosed by<br />

considering separately its three components and defining the transmission efficiency as:<br />

W<br />

η tr =<br />

W<br />

p<br />

<br />

t<br />

5.17<br />

where p W and t W are respectively the power delivered to the pump and the power<br />

provided by the turbine. The problem <strong>of</strong> this approach is that these two values are<br />

seldom available. For this reason, it is preferable to define a global efficiency<br />

comprising turbine, transmission and pump. By combining equations <strong>of</strong> isoentropic<br />

efficiency and energy balance <strong>of</strong> a turbine, efficiency <strong>of</strong> a transmission and efficiency <strong>of</strong><br />

a pump, it yields:<br />

mp⋅ν⋅Δp ηtp = ηt⋅ηtr ⋅ ηp=<br />

m⋅ h −h<br />

( , , , )<br />

t in t s out t<br />

The previous equation shows that turbo-pump efficiency is the quotient between the<br />

minimum power to pump the water and the maximum power obtained by expanding the<br />

steam, considering ideal processes in both cases. This efficiency is very suitable for<br />

<strong>diagnosis</strong> purposes. However, it varies with load, and this dependence should be taken<br />

into account for simulation (Alconchel, 1988).<br />

5.2.1.6 Pressure losses<br />

Pressure losses in water-steam circuit appear in turbine inlet valves, cross-over pipe,<br />

regenerative water heaters, steam extraction and heat exchangers and ducts in the boiler.<br />

Zaleta (1997) provides expressions for pressure drops in turbine inlet valves:<br />

5.18<br />

103


Chapter 5<br />

104<br />

v<br />

Sa = m ⋅<br />

a<br />

5.20<br />

Where m is the steam flow rate, ν is the specific volume and a is the sound speed for<br />

steam in those conditions, k varies from 0.3 and 2.5 and D is the valve diameter.<br />

For the cross-over pipe:<br />

2<br />

c<br />

Δ px−over=<br />

0.25⋅<br />

5.21<br />

ν<br />

Where c is the steam speed and ν is the specific volume. Alconchel (1988) proposes<br />

the following expression for pressure losses in ducts connecting turbine and feedwater<br />

heaters:<br />

Δ p<br />

= 0.03 if extraction comes from the final <strong>of</strong> a turbine 5.22<br />

p<br />

Δ p<br />

= 0.05 if extraction comes from an intermediate point <strong>of</strong> a turbine 5.23<br />

p<br />

These values have obtained from the values observed in different plants. For a given<br />

unit, it is possible to adjust them by using design plant data.<br />

Pressure losses in regenerative water heaters, reheater and other parts <strong>of</strong> the boiler<br />

can be calculated by potential or polynomial equations adjusted to plant balance:<br />

n<br />

Δ p = k⋅m 5.24<br />

Δ p = a+ b⋅ m+ c⋅m 2 5.25<br />

Where k, a, b and c are generic coefficients and n is an exponent that can vary from<br />

1.6 to 2. Alconchel (1988) has determined that, for the case <strong>of</strong> study, pressure losses are<br />

almost constant in reheater and linear with the load in the other heat exchangers.<br />

Although the previous equations should be used for simulation purposes, pressure<br />

drops can be used directly for <strong>diagnosis</strong> purposes. In order to simplify the modelling<br />

and <strong>diagnosis</strong>, pressure losses in ducts can be supposed to occur in one <strong>of</strong> the<br />

components that the duct connects.<br />

5.2.2 Cooling system<br />

The cooling system is composed <strong>of</strong> a condenser, a cooling tower and pumps and<br />

ducts system that is responsible for the water circulation among them.


Case study: 3x350 MW coal-fired power plant<br />

5.2.2.1 Condenser<br />

To model the condenser, a transfer equation should be included in order to take into<br />

account the capacity <strong>of</strong> the equipment to condense the steam while heating the cooling<br />

water. This equation can be based on the mean logarithmic temperature:<br />

Q = F⋅U⋅A⋅ΔT 5.26<br />

lm<br />

Where Q is the amount <strong>of</strong> heat exchanged, U is the global transfer coefficient, A is<br />

the area and F is a factor which takes into account the heat exchanger configuration (if<br />

tubes are parallel or not). When one <strong>of</strong> the fluids changes its phase, the value <strong>of</strong> F is 1.<br />

ΔTlm is the log-mean temperature difference:<br />

Δ T =<br />

lm<br />

( Th, in −Tc, out ) −( Th, out −Tc,<br />

in )<br />

⎛T −T<br />

⎞<br />

ln<br />

h, in c, out<br />

⎜ ⎟<br />

Th, out T ⎟<br />

⎝ − c, in ⎠<br />

Where subscripts h and c means hot and cold, respectively. For the particular<br />

example <strong>of</strong> a condenser, where there are a flow <strong>of</strong> water (w) and condensing steam that<br />

enters and exits at the same temperature (s), the previous expression becomes:<br />

Tw, in −Tw,<br />

out<br />

Δ Tlm,<br />

c =<br />

5.28<br />

⎛Ts−T ⎞ w, out<br />

ln ⎜ ⎟<br />

Ts T ⎟<br />

⎝ − w, in ⎠<br />

Another way to introduce the characteristic equation is given by the ε-NTU method.<br />

The effectiveness <strong>of</strong> a heat exchanger (ε), relates the heat transferred to the maximum<br />

heat that could be achievable in an ideal infinite exchanger:<br />

Q = ε ⋅Q<br />

5.29<br />

Where<br />

max<br />

( hin cin)<br />

max min , ,<br />

5.27<br />

Q = C ⋅ T −T<br />

5.30<br />

( h p h c p c)<br />

C = min m ⋅c , m ⋅c<br />

5.31<br />

min , ,<br />

The previous equations show that the maximum heat would correspond to a situation<br />

when one <strong>of</strong> the flows exits the heat exchanger at the same temperature at which the<br />

other enter. That flow would be the one for which the product <strong>of</strong> flow rate and specific<br />

heat capacity is minimum. In university texts on heat transfer (Mills, 1992) are available<br />

equations that relate the efficiency to two parameters, for several heat exchangers<br />

configurations. These two parameters are heat capacity rates relation, Cr, and the<br />

number <strong>of</strong> transfer units, NTU:<br />

105


Chapter 5<br />

106<br />

C<br />

C<br />

min<br />

r = 5.32<br />

Cmax<br />

U⋅A NTU = 5.33<br />

C<br />

min<br />

For heat exchangers where one <strong>of</strong> the flows suffers a change <strong>of</strong> phase (Cr tends to<br />

zero), effectiveness equation becomes quite simple:<br />

− NTU<br />

ε = 1−<br />

e<br />

5.34<br />

This second method is preferable when numerical <strong>methods</strong> are used to solve the set<br />

<strong>of</strong> equations in simulation because the equation to calculate ΔTlm may difficult<br />

convergence when cold and hot side temperatures are quite closed, as occurs in a<br />

condenser. In other situations, the log-mean temperature difference method has the<br />

advantage that it is not needed to asses previously which one <strong>of</strong> the two flows has lower<br />

value <strong>of</strong> C.<br />

In both <strong>methods</strong>, it is necessary to obtain the global transfer coefficient, which can<br />

be calculated by thermal resistances composition, taking into account convection and<br />

fouling in the two sides <strong>of</strong> each tube, and conduction in the tube wall.<br />

1 1 Aext Aext<br />

1<br />

= ⋅ + Fint + + Fext<br />

+<br />

5.35<br />

U D<br />

ext hint Aint ext<br />

2⋅π⋅kln hext<br />

w ⋅<br />

D<br />

int<br />

Where h the convection coefficient, A is area, F is fouling resistance and kw is<br />

thermal conductivity <strong>of</strong> the wall. Subscripts int and ext stand for interior and exterior. It<br />

should be noted that the value <strong>of</strong> U calculated should be multiplied by the exterior area<br />

in equations 5.26 and 5.33. It would be equivalent to calculate the value <strong>of</strong> U related to<br />

the interior by modifying slightly the equation and then to multiply it times the interior<br />

area.<br />

The convection coefficient is calculated from the dimensionless number <strong>of</strong> Nusselt<br />

(Nu):<br />

Nu ⋅k<br />

h = 5.36<br />

D<br />

Where k is the thermal conductivity <strong>of</strong> the substance and D is the inner or outer<br />

diameter. Suitable correlations for calculating the value <strong>of</strong> Nu depending <strong>of</strong> the situation<br />

can be found in university textbooks on Heat Transfer such as Mills (1992).<br />

As a <strong>diagnosis</strong> variable to summarize the condenser behaviour, effectiveness has<br />

been used, because <strong>of</strong> its universality and easiness to be calculated from plant data:


ε<br />

( )<br />

( )<br />

Q<br />

m ⋅c ⋅ T −T T −T<br />

c w p, w w, out w, in w, out w, in<br />

c =<br />

Q = =<br />

c,max mw⋅cp, w⋅ Tsat −Tw,<br />

in Tsat −Tw,<br />

in<br />

Case study: 3x350 MW coal-fired power plant<br />

5.2.2.2 Cooling tower<br />

To model the behaviour <strong>of</strong> a cooling tower, it is possible to use a polynomial adjust<br />

<strong>of</strong> curves provided by tower manufacturer. These curves provide the tower outlet<br />

temperature as a function <strong>of</strong> the cooling water flow rate, temperature difference between<br />

water entering and exiting the tower, ambient temperature and relative humidity. An<br />

expression quadratic for the flow rate and linear for the other variables is enough.<br />

The most extended theoretical model for cooling towers is based on the Merkel<br />

equation. It is derived by considering that the heat exchange in the tower depends on the<br />

enthalpy difference <strong>of</strong> saturated air and air in the conditions <strong>of</strong> the tower. By<br />

considering a differential <strong>of</strong> height in a section <strong>of</strong> the tower, the following equation<br />

applies:<br />

( )<br />

air, sat air w p, w w<br />

5.37<br />

dq = β ⋅ h −h ⋅a⋅ dV = m ⋅c ⋅dT<br />

5.38<br />

where β is a mass transfer coefficient and a is the area <strong>of</strong> tower filling per unit <strong>of</strong><br />

volume. By integrating and rearranging the previous equation, the Merkel number<br />

appears, which is a parameter associated to the tower design.<br />

Twin<br />

,<br />

β ⋅ A c ⋅ dT<br />

Me = =<br />

m∫ h − h<br />

pw , w<br />

w Twout<br />

, air, sat air<br />

where A is the total area <strong>of</strong> the tower filling. To calculate the previous integral, the<br />

Chebyshev method is usually used (Perry, 1984):<br />

( )<br />

c ⋅ T −T<br />

5.39<br />

Tw,in<br />

cpw , ⋅ dTwpw<br />

, win , wout , ⎛ 1 1 1 1 ⎞<br />

∫ ≈ ⋅ ⎜ + + + ⎟<br />

5.40<br />

h<br />

, , 4<br />

Twout<br />

air sat −hair ⎝Δh1 Δh2 Δh3 Δh4<br />

⎠<br />

Δh1 is the value <strong>of</strong> hair,sat – hair at a temperature [Tw,out + 0.1·(Tw,in – Tw,out)]<br />

Δh2 is the value <strong>of</strong> hair,sat – hair at a temperature [Tw,out + 0.4·(Tw,in – Tw,out)]<br />

Δh3 is the value <strong>of</strong> hair,sat – hair at a temperature [Tw,out + 0.6·(Tw,in – Tw,out)]<br />

Δh4 is the value <strong>of</strong> hair,sat – hair at a temperature [Tw,out + 0.9·(Tw,in – Tw,out)]<br />

Finally, in a natural-draft tower, the amount <strong>of</strong> air should accomplish the flotability<br />

equation, which states that the pressure drop <strong>of</strong> air crossing the tower is equal to the<br />

force experimented by the air due to the density differences (Muñoz et al., 1986):<br />

( )<br />

Δ p = g⋅H⋅ ρ − ρ<br />

5.41<br />

air, tower air, in air, out<br />

107


Chapter 5<br />

Where g is the gravity constant, H is the tower height and ρair is the air density.<br />

Better results can be obtained by using more complex models. For example, Al-<br />

Waked and Behnia (2006) have developed a CFD model <strong>of</strong> cooling towers which<br />

allows to consider even influence <strong>of</strong> wind.<br />

The previous models can be used to simulate the tower. However, neither the<br />

Merkel number nor the cooled water temperature are suitable free <strong>diagnosis</strong> variables.<br />

The first one is not commonly used by plant operators and does not have a simple<br />

definition. The second is simple, but vary at least an order <strong>of</strong> magnitude more for<br />

ambient conditions than for tower degradation. The parameter proposed is a tower<br />

effectiveness related to ambient conditions, which consider not only air temperature but<br />

also its humidity.<br />

108<br />

ε<br />

T −T<br />

=<br />

5.42<br />

w, in w, out<br />

tower<br />

Twin , −Twb<br />

Where Tw is the temperature <strong>of</strong> water entering or exiting the tower and Twb is the<br />

wet-bulb temperature <strong>of</strong> air.<br />

5.2.2.3 Pumps and ducts<br />

Expressions to model the pumps have been presented previously, in the paragraph<br />

corresponding to the cycle. Besides, the pressure drop in ducts can be modelled by<br />

parabolic equations adjusted to plant data or by equations available in textbooks on<br />

Fluid Mechanics (White, 1979). However, to diagnose separately pump and ducts it<br />

would be necessary the value <strong>of</strong> water pressure at pump inlet and outlet, which is not<br />

usually available. Besides, pumps operation strategy is to work at full load in order to<br />

keep the water mass flow at design conditions. For these reasons, the <strong>diagnosis</strong> variable<br />

chosen is the cooling water flow rate. This parameter can be calculated by using only<br />

water temperatures (and heat rejected by the cycle), without any pressure losses model<br />

<strong>of</strong> the system.<br />

5.2.3 Boiler<br />

Once the steam cycle and cooling system have been presented, this last section<br />

describes the models and <strong>diagnosis</strong> variables <strong>of</strong> the boiler. First, the separated lossesmethod<br />

to calculate the boiler efficiency and coal demand is described. Then, the<br />

combustion model is presented. Afterwards, heat transfer models for the boiler and air<br />

heaters are described, along with some comments on the blowing system. It should be


Case study: 3x350 MW coal-fired power plant<br />

noted that heat transfer in a boiler is a very complex issue that is going to be simplified<br />

in this work. Finally, some comments on ancillary devices are included.<br />

5.2.3.1 Boiler efficiency and coal consumption<br />

The method to calculate coal consumption and boiler efficiency has been adapted<br />

from that implemented in a monitoring system installed in the plant (Uche, 2001). This<br />

fact allows plant staff to compare easily results provided by the two systems. It is based<br />

in the ASME PTC 4.1 standard for calculating the thermal efficiency <strong>of</strong> boilers (ASME,<br />

1964).<br />

The main differences come from the different definition <strong>of</strong> the system boundary,<br />

which is a key issue in the calculation <strong>of</strong> the efficiency, because it affects strongly to the<br />

flows to be used to calculate this parameter. ASME PTC 4.1 standard suggest that<br />

forced draft fans, recycling air fans, electrostatic precipitators or other ash elimination<br />

devices downstream the air heaters, induced draft fans, economizers for using residual<br />

gases heat (downstream the induced draft fans) and stack should be located outside the<br />

control volume. So that, the only ancillary systems whose consumption is considered by<br />

the standard are: air-gas preheaters, primary air fans (if there are one per each mill, they<br />

are located after the preheaters), mills, fans for recycling hot gases towards the furnace<br />

and pumps to assist water circulation in the furnace. This criterion is not strictly adopted<br />

here. First, induced draft fans and electrostatic precipitators are included in the control<br />

volume, because reliable temperature measurement is located in induced draft fans<br />

impulsion. Second, primary air fans and steam coil heaters and mills are considered<br />

outside. More details on these differences can be found in Uche (2001) and in Rangel<br />

(2005).<br />

The calculation <strong>of</strong> coal consumption is done by considering the energy balance <strong>of</strong><br />

the boiler:<br />

E = U + L<br />

5.43<br />

Where E is the total amount <strong>of</strong> energy entering the boiler, U is the utile heat<br />

delivered to the cycle and other parts <strong>of</strong> the power plant and L are the loses. Efficiency<br />

is calculated by:<br />

U L<br />

η boiler = = 1−<br />

5.44<br />

E E<br />

According to the previous expression, there are two ways to calculate the efficiency.<br />

The first consists directly in dividing the utile energy into the energy entering the boiler<br />

109


Chapter 5<br />

(direct method), which is not possible because the amount <strong>of</strong> fuel is not measured. On<br />

the other hand, is also possible to determine the efficiency by considering the losses<br />

(indirect method). This second method has been widely used traditionally because it is<br />

easy to implement by expressing losses in energy units per kilogram <strong>of</strong> coal. When the<br />

calculation is integrated in a global system to calculate all the plant, both energy balance<br />

and a form <strong>of</strong> the efficiency equation are solved simultaneously, so that both <strong>methods</strong><br />

are actually merged.<br />

Energy entering the boiler is composed <strong>of</strong> two terms, the enthalpy <strong>of</strong> the fuel and<br />

other incomes:<br />

110<br />

E = H f + B<br />

5.45<br />

Energy <strong>of</strong> the fuel includes the coal and, sometimes, other fuels such as natural gas.<br />

It should be noted that the amount <strong>of</strong> gas can be easily measured, so that the amount <strong>of</strong><br />

coal can be calculated by difference. Fuel energy can be approximated by the product <strong>of</strong><br />

flow rate and high heating value (Díez, 2001):<br />

H = m ⋅ HHV + m ⋅HHV<br />

5.46<br />

f f f ng ng<br />

Other incomes are due to forced draft fan, primary air fan, secondary and primary<br />

coil heaters:<br />

B = B + B + B + B<br />

5.47<br />

fdf paf 1ch 2ch<br />

Each one <strong>of</strong> the terms <strong>of</strong> the previous equation is calculated by multiplying the<br />

amount <strong>of</strong> air times the enthalpy increment suffered by it:<br />

( )<br />

B = m ⋅ h −h<br />

5.48<br />

i air, i air, out air, in<br />

The previous equation shows that only the effective energy transferred to the air is<br />

considered, without taking into account losses produced for example in the electric<br />

motor. This definition allows to keep the energy balance <strong>of</strong> the control volume, but is<br />

not suitable to diagnose malfunctions in these ancillary devices.<br />

Utile heat provided by the boiler is divided in two terms: utile heat to the cycle and<br />

utile heat to ancillaries:<br />

U = U + U<br />

5.49<br />

boiler −cycle boiler −ancillaries<br />

Heat delivered by the boiler to the steam cycle should take into account the<br />

following flows: main steam, reheated steam (cold and hot), boiler feeding water and<br />

tempering water:<br />

Uboiler− cycle = mms ⋅hms−mcrs ⋅ hcrs + mhrs ⋅hhrs −m fw ⋅hfw−m tw ⋅htw<br />

5.50


Case study: 3x350 MW coal-fired power plant<br />

Heat to ancillaries includes sootblowing, uncontrolled leakages, drum purge, and<br />

primary and secondary coil heaters:<br />

U − = U + U + U + U + U<br />

5.51<br />

boiler ancillaries sb ul dp 1ch 2ch<br />

Terms <strong>of</strong> the previous equations can be calculated by using the following<br />

expressions:<br />

( ( ) )<br />

( ( ) ( ) )<br />

( ) ( )<br />

U = m ⋅ h −h<br />

T<br />

5.52<br />

sb sb sb sl ref<br />

U = m ⋅ h p −h<br />

T<br />

5.53<br />

ul ul ss drum sl ref<br />

( )<br />

( ( ) ( ) )<br />

( ( ) ( ) )<br />

U = m ⋅ h p −h<br />

T<br />

5.54<br />

dp dp sl drum sl ref<br />

U = m ⋅ h p −h<br />

T<br />

5.55<br />

1ch 1ch<br />

ss drum sl ref<br />

U = m ⋅ h p −h<br />

T<br />

5.56<br />

2ch 2chss<br />

drum sl ref<br />

Where hsl and hss are the enthalpies <strong>of</strong> saturated liquid and saturated steam,<br />

respectively, pdrum is the drum pressure and Tref is the reference temperature (ambient<br />

temperature).<br />

After considering the boiler incomes and the utile heat, only boiler losses have to be<br />

studied prior to finishing the paragraph. Losses considered are due to: unburned carbon<br />

in ashes (uc), coal moisture (mf), hydrogen (H), air moisture (mA), CO in gases (CO),<br />

radiation and convection (B), sensible heat in ashes (d), sensible heat in dry flue gases<br />

(gs), air to sulphur removing unit (sru) unit and other losses (ol):<br />

L= Luc + Lmf+ LH+ LmA+ LCO+ LB+ Ld+ Lgs+ Lsru + Lol<br />

5.57<br />

Losses due to unburned carbon in ashes appears due to incomplete combustion that<br />

leads to residual carbon present in ashes and slag:<br />

L = 4.1868⋅8055.7⋅m ⋅unb<br />

[kW] 5.58<br />

uc f<br />

Where 8055.7 kcal/kg is the energy released in the combustion <strong>of</strong> one kilogram <strong>of</strong><br />

carbon, 4.1868 is the conversion factor from kcal to kJ, m f is the coal flow rate and unb<br />

is the number <strong>of</strong> kilograms <strong>of</strong> unburned carbon in residues per kilogram <strong>of</strong> coal entering<br />

the boiler:<br />

unbres<br />

unb = Z ⋅<br />

1−<br />

unbres<br />

5.59<br />

unbres = fashes ⋅ unbresashes + fslag ⋅ unbresslag<br />

5.60<br />

111


Chapter 5<br />

Where Z is the mass fraction <strong>of</strong> ash in coal, unbres is the number <strong>of</strong> kilograms <strong>of</strong><br />

carbon per kilogram <strong>of</strong> residues, which is calculated by taking into account the fraction<br />

<strong>of</strong> residues which is ashes and the fraction which is slag.<br />

Losses for fuel moisture are calculated by:<br />

112<br />

( ( , , ) ( ) )<br />

2<br />

L = W ⋅m ⋅ h T p −h<br />

T<br />

5.61<br />

mf f s go g H O sl ref<br />

Where w is the mass fraction <strong>of</strong> moisture in fuel and Tgo and pg,H2O are the<br />

temperature and partial pressure <strong>of</strong> steam in the gases leaving the control volume.<br />

Reference enthalpy corresponds to saturated liquid at reference temperature, because<br />

high heating value is used for the fuel.<br />

Losses due to hydrogen in fuel correspond to the water formed in the combustion <strong>of</strong><br />

the hydrogen contained in the fuel, so that they are calculated in a parallel way as fuel<br />

moisture losses:<br />

( )<br />

2<br />

( ) ( , ) ( )<br />

L = 8.936 ⋅ H⋅ m + H ⋅m ⋅ h T , P −h<br />

T<br />

5.62<br />

H f ng ng s go g H O sl ref<br />

Where 8.936 are the kilograms <strong>of</strong> water produced in the combustion <strong>of</strong> one kilogram<br />

<strong>of</strong> hydrogen and H and Hng are the mass fractions <strong>of</strong> hydrogen in coal and natural gas,<br />

respectively.<br />

Humidity <strong>of</strong> air also contributes to the losses. However, saturated steam is<br />

considered as reference because this water enters the boiler as steam:<br />

( ( , ) ( ) )<br />

2<br />

L = W ⋅m ⋅ h T p −h<br />

T<br />

5.62<br />

mA mA a, dry s go g, H O ss ref<br />

Where WmA is the number <strong>of</strong> kilograms <strong>of</strong> water in air per kilogram <strong>of</strong> dry air. It can<br />

be calculated from ambient temperature and pressure and relative humidity (Φ):<br />

W<br />

mA<br />

( )<br />

( )<br />

psat Tamb<br />

= 0.622⋅Φ⋅<br />

p −Φ⋅p<br />

T<br />

amb sat amb<br />

Where 0.622 is the relation between the molar masses <strong>of</strong> water and dry air.<br />

Losses per unburned gases take into account the effect produced by the fraction <strong>of</strong><br />

carbon that has been partially oxidized to CO:<br />

CO CO, g<br />

5.63<br />

L = 23631⋅12⋅n [kW] 5.64<br />

Where 23631 kJ/kg is the difference <strong>of</strong> energy released by burning 1 kilogram <strong>of</strong><br />

carbon producing CO2 and producing CO, 12 is the carbon molar mass and nCO,g is the<br />

number <strong>of</strong> kmol <strong>of</strong> CO per second produced in the combustion.


Case study: 3x350 MW coal-fired power plant<br />

Although the standard ASME PTC 4.1 (ASME, 1964) provides curves to calculate<br />

boiler losses due to convection and radiation, for the case <strong>of</strong> study there were empirical<br />

correlations available (Uche, 2001), which have been used here:<br />

4 4<br />

( ( ) ( ) )<br />

Q = S⋅ ⋅ ⋅ ⋅T − − ⋅T − [kcal/h] 5.65<br />

−9<br />

rad 4.48286 10 1.8 wall 491.68 1.8 amb 491.68<br />

( ( ) )<br />

Q = S⋅4.8816⋅ 1+ 0.73818⋅v<br />

⋅ T − T [kcal/h] 5.66<br />

conv wind wall amb<br />

4.1868<br />

LB= ⋅ ( Qrad + Qconv<br />

) [kW] 5.67<br />

3600<br />

Where S is the boiler outer surface in square meters (4158 m 2 ), Tamb and Twall are<br />

respectively temperatures <strong>of</strong> ambient and boiler wall in ºC (the second one has been<br />

fixed to 40ºC) and vwind is the wind speed in m/s.<br />

Losses in ashes and slag are calculated as:<br />

( ( ) ( ) )<br />

( , ) , ( )<br />

L = f ⋅Z⋅m ⋅ c T ⋅T −c T ⋅T<br />

5.68<br />

d ash f ash go go ash ref ref<br />

( )<br />

L = f ⋅Z⋅m ⋅ c T ⋅T −c T ⋅T<br />

5.69<br />

p slag f slag slag o slag o slag ref ref<br />

Where cash and cslag are the specific heat <strong>of</strong> ash and slag, respectively, and Tslag,o is<br />

the temperature <strong>of</strong> slag leaving the furnace. It is considered that flying ashes have the<br />

same temperature as the gas.<br />

The main contribution to the losses corresponds to the sensible heat <strong>of</strong> dry flue gases<br />

leaving the boiler:<br />

( ( ) ( ) )<br />

L = m ⋅ h T −h<br />

T<br />

5.70<br />

gs gdry , gdry , go gdry , ref<br />

A flow <strong>of</strong> hot air is extracted from the boiler to heat-up flue gases leaving the flue<br />

gas desulfuration unit. This energy flow leaving the control volume is also considered<br />

as losses:<br />

( ( ) ( ) )<br />

L = m ⋅ h T −h<br />

T<br />

5.71<br />

fgd air, fgd air fgd air ref<br />

Finally, a 0.7 % corresponding to other losses is also added, which takes into<br />

account other issues not considered previously.<br />

L = 0.007⋅<br />

E<br />

5.72<br />

ol<br />

It should be noted that sootblowing steam has not been considered a loss but a useful<br />

product. Due to this choice, variation in this magnitude will not affect boiler efficiency.<br />

113


Chapter 5<br />

5.2.3.2 Combustion<br />

The aim <strong>of</strong> this paragraph is to explain the <strong>diagnosis</strong> variables, equations and models<br />

related to combustion.<br />

To determine the composition <strong>of</strong> gases leaving the furnace, first it is necessary to<br />

consider the fuel composition. Coal contains carbon, hydrogen, nitrogen, oxygen,<br />

sulphur, ashes and moisture. All these mass fractions minus one (oxygen, which is<br />

calculated by difference) are degrees <strong>of</strong> freedom and, subsequently, free <strong>diagnosis</strong><br />

variables. Natural gas contains CH4, C2H6, C3H8, CO2 and other components. Again, the<br />

molar fractions <strong>of</strong> all components minus one are degrees <strong>of</strong> freedom and can be free<br />

<strong>diagnosis</strong> variables. Besides, relation between the amounts <strong>of</strong> coal and natural gas<br />

should be specified. A good variable to this relation is the fraction <strong>of</strong> energy provided<br />

by natural gas. High heating values <strong>of</strong> coal and gas are also free <strong>diagnosis</strong> variables.<br />

Once the fuel composition is determined, gas composition is calculated by matter<br />

balances. If complete and stoichiometric combustion is achieved, gas contains: CO2,<br />

H2O, SO2, N2 and Ar. The amount and composition <strong>of</strong> gases and the amount <strong>of</strong> air<br />

needed are calculated by performing matter balances for carbon, hydrogen, sulphur,<br />

oxygen, nitrogen and argon.<br />

In practical applications, stoichiometric combustion is not possible, so that an<br />

additional amount <strong>of</strong> air is introduced (air excess), which implies an additional degree<br />

<strong>of</strong> freedom. The variable chosen is the oxygen concentration <strong>of</strong> gases leaving the boiler<br />

(before air preheaters), which is usually a set point to control the air introduced in the<br />

boiler.<br />

Besides, real combustion is not complete, so that unburned carbon appears in ash<br />

and CO is present in gases, which adds two more degrees <strong>of</strong> freedom and <strong>diagnosis</strong><br />

variables: mass fraction <strong>of</strong> carbon in ashes and molar fraction <strong>of</strong> CO in flue gases. Flue<br />

gases have minor amounts <strong>of</strong> other components such as NOx, which are <strong>of</strong> great<br />

importance from the environmental point <strong>of</strong> view, but its influence <strong>of</strong> energy efficiency<br />

is negligible, so that they are not considered. It should be noted that all these variables<br />

are not actually independent, because they depend strongly <strong>of</strong> the type <strong>of</strong> coal and boiler<br />

operation (excess air, choice <strong>of</strong> burners rows…). However, due to the difficulty to<br />

determine these relations, they are going to be considered as independent.<br />

A lot <strong>of</strong> research has been done in order to predict unburned coal in ashes, some <strong>of</strong><br />

them applied to the case <strong>of</strong> study <strong>of</strong> this thesis. Influence <strong>of</strong> CO in boiler efficiency is<br />

about one order <strong>of</strong> magnitude lower than effect <strong>of</strong> carbon in ashes (Díez, 2002).<br />

114


Case study: 3x350 MW coal-fired power plant<br />

A method to predict unburned carbon in ashes has to calculate the combustion time<br />

<strong>of</strong> coal particles in the furnace, which is the time necessary to burn all the carbon in the<br />

particle. If it is lower than the residence time, all particle is burned; if not, an amount <strong>of</strong><br />

residual carbon appears. Combustion <strong>of</strong> a carbon particle in a utility boiler can be<br />

divided in three stages (Spalding, 1983):<br />

1. Devolatilisation: is the phase when volatiles contained in the particle are released.<br />

2. Homogeneous volatile combustion.<br />

3. Heterogeneous combustion <strong>of</strong> the carbonaceous residue (fixed carbon + ashes).<br />

Time necessary for the first and second stages are very reduced compared to the<br />

time for the last one, so that this third stage is going to determine the combustion time.<br />

Two models have been proposed to simulate it. The first one is the progressive<br />

conversion model, which considers that the gas surrounding the particle enters inside it<br />

and all particle reacts simultaneously.<br />

The unreacted core shrinking model considers that the reaction starts in the outer<br />

part <strong>of</strong> the particle and goes towards the particle centre. Díez (2002) applies this<br />

methodology to one boiler <strong>of</strong> Teruel power plant.<br />

CFD techniques have also been applied to simulate combustion in this boiler (Iranzo<br />

et al., 1999, 2001).<br />

5.2.3.3 Heat transfer in boilers<br />

To model heat transfer in a boiler, it has been proposed to decompose it in a set <strong>of</strong><br />

heat exchangers and to apply to each one <strong>of</strong> them either the ε-NTU or the ΔTlm <strong>methods</strong><br />

presented in the condenser section.<br />

An important issue <strong>of</strong> heat transfer in the boiler that was negligible in the condenser<br />

is radiation. If radiation absorbed by a heat exchanger is considered and heat emission is<br />

neglected (which is a good approximation due to the low temperature <strong>of</strong> the surfaces<br />

and the high temperature <strong>of</strong> the gases or the flame (Díez, 2002)), the energy balance<br />

becomes:<br />

Q + α ⋅G⋅ A = Q<br />

5.73<br />

h w h c<br />

where h Q and c Q are the heat corresponding to hot and cold sides, αw is the wall<br />

absorptivity, G is the irradiation and Ah is the area <strong>of</strong> the hot side <strong>of</strong> the heat exchanger.<br />

The problem <strong>of</strong> the radiative term appears when the transference equation is written,<br />

because two options are possible:<br />

FUA ⋅ ⋅Δ T + α ⋅GA ⋅ = Q<br />

5.74a<br />

lm w h c<br />

115


Chapter 5<br />

116<br />

Q + α ⋅G⋅ A = F⋅UA⋅Δ T<br />

5.74b<br />

h w h lm<br />

There is no reason for choosing one or the other, and this choice can lead to different<br />

solutions. A method has been developed to solve this problem (Díez, 2002). The<br />

mathematical development is similar to the one used to develop the ΔTlm equation: to<br />

consider a section <strong>of</strong> differential length and constant temperature <strong>of</strong> both hot and cold<br />

fluids. However, the contribution <strong>of</strong> the radiation <strong>of</strong> this section is also considered. By<br />

integrating that balance, the following equation is obtained for a counterflow heat<br />

exchanger with radiation from outside:<br />

⎛Th, out − Tc, in + k ⎞ ⎛Tc, out −Tc, in Th, in −Th,<br />

out ⎞<br />

ln ⎜ = UA⋅<br />

−<br />

⎜ ⎟ ⎜ ⎟<br />

Th, in − Tc, out + k ⎟ Q ⎝ c Q<br />

5.75<br />

⎝ ⎠<br />

h ⎠<br />

Where k is a factor with temperature dimensions that take into account the effect <strong>of</strong><br />

radiation:<br />

α w⋅G⋅Ah αw⋅G<br />

k = +<br />

5.76<br />

⎛C⎞ h<br />

h<br />

h<br />

UA⋅⎜ −1⎟<br />

⎝Cc⎠ where hh is the convection coefficient <strong>of</strong> the hot side and Cc and Ch are the relations<br />

between the energy and temperature variations <strong>of</strong> cold and hot fluids:<br />

( , , )<br />

Q = C ⋅ T −T<br />

5.78<br />

c c c out c in<br />

( , , )<br />

Q = C ⋅ T −T<br />

5.79<br />

h h h in h out<br />

The transference equation obtained can be completed by adding a factor F<br />

multiplying U·A, in order to take into account effect due to non parallel fluxes. The use<br />

<strong>of</strong> this factor in this equation is an approximation, but at least it can be made from an<br />

excact equation. As it can be seen, when radiation is not present, the value <strong>of</strong> k is zero<br />

and the transfer equation becomes the classical ΔTlm expression.<br />

Once the transfer equation has been presented, the global heat transfer coefficient<br />

should be calculated, by using the thermal resistances composition equation previously<br />

presented (5.2.2.1).<br />

Due to the high temperatures <strong>of</strong> the hot fluid (combustion gases), convection<br />

coefficient in the hot side should take into account not only convective but also radiant<br />

contributions:<br />

hh= hconv + hrad<br />

5.80


( )<br />

h g wall<br />

Case study: 3x350 MW coal-fired power plant<br />

hconv can be calculated by using suitable correlations available in bibliography<br />

(Mills, 1992) The apparent radiation coefficient hrad is defined according to the heat<br />

transferred by radiation from the gases inside the heat exchanger to the tubes:<br />

Q<br />

rad , g→wall hrad<br />

=<br />

5.81<br />

A ⋅ T −T<br />

It should be noted that the apparent radiation coefficient takes into account the<br />

radiation from the gas located inside the exchanger (next to the tubes) and does not<br />

affect the energy balance <strong>of</strong> the heat exchanger. Radiation from other areas <strong>of</strong> the boiler<br />

(mainly the furnace and the plenum <strong>of</strong> gases) is introduced by the flux G which has<br />

originated the modification <strong>of</strong> the transference equation. For example, radiation from<br />

the furnace can be calculated by:<br />

4 4<br />

( )<br />

Q = F ⋅A ⋅σ ⋅ε ⋅ T −T<br />

5.82<br />

rad , furnace ij i gas gas, furnace dwall<br />

Where Fij is the view factor between two surfaces, which represents the fraction <strong>of</strong><br />

radiation leaving the furnace (surface i) that arrives at the heat exchanger (surface j),<br />

and depends on the geometry, σ is the Stefan-Boltzman constant and Tdwall is the<br />

temperature <strong>of</strong> “dirty wall” <strong>of</strong> the tubes. This expression can also be used to calculate<br />

radiation from the plenum, by considering the suitable temperatures and view factors.<br />

To model the furnace, Kakaç (1991) cites the standard method developed by Blokh<br />

(1984). It uses experimental coefficients obtained in test performed in different boiler<br />

types, mainly in China and the former USSR, where this method was adopted as a<br />

standard for boiler manufacturers. The basic equation <strong>of</strong> this method is the Gurvich<br />

formula, which provides the value <strong>of</strong> the temperature <strong>of</strong> gases leaving the furnace:<br />

T = T ⋅<br />

Bo<br />

0.6<br />

go, furnace g, ac 0.6 0.6<br />

Bo + M ⋅εF<br />

Where Tg,ac is the adiabatic combustion temperature, M is a coefficient which takes<br />

into account temperature distribution in the furnace, εF is the average emissivity <strong>of</strong> the<br />

furnace and Bo is the Boltzman number, which represents the relation between heat<br />

transferred by enthalpy variation and by radiation. Actually, the previous equation<br />

represents a global energy balance in the furnace.<br />

Ideas previously presented summarized a heat transfer model developed by Díez<br />

(2002) for Teruel power plant. Boiler has been discretized in: furnace, primary and<br />

secondary economizers, reheaters, primary, secondary and radiant superheaters, gases<br />

5.83<br />

117


Chapter 5<br />

plenum and boiler walls next to the radiant superheater and the plenum. The following<br />

radiation flows have been considered: furnace to secondary and tertiary superheaters<br />

and to walls, secondary superheater to walls and tertiary superheater, and plenum to<br />

tertiary superheater, primary superheater, reheater and walls.<br />

This model has been simplified and included in this work. By using a simulator,<br />

relations between radiation flows and heat exchangers effectiveness with load have been<br />

obtained. This simplification may be strong but it is justified because interest is focused<br />

in the global behaviour <strong>of</strong> the boiler, and models for every individual heat exchanger are<br />

considered <strong>of</strong> secondary importance.<br />

The <strong>diagnosis</strong> variable proposed for heat exchangers is its effectiveness. However,<br />

due to the lack <strong>of</strong> information, it is not possible to diagnose individually all heat<br />

exchangers. The aim is to determine whether the boiler is able to use efficiently the heat<br />

<strong>of</strong> flue gases, so that a good parameter may be the average temperature <strong>of</strong> gases leaving<br />

the boiler (prior to air preheaters):<br />

118<br />

T<br />

go, av<br />

m ⋅ T + m ⋅T<br />

=<br />

m + m<br />

g, bypass g , bypass go, ez1 go, ez1<br />

g, bypass go, ez1<br />

In the previous equation, flow <strong>of</strong> gases leaving the primary economizer and by pass<br />

gases extracted after the reheater have been considered. Specific heat variation has been<br />

neglected. This variable can be transformed into an effectiveness by considering<br />

reference temperatures for adiabatic combustion temperature and water entering the<br />

boiler:<br />

ε<br />

=<br />

T −T<br />

5.84<br />

5.85<br />

g, ac, ref go, av<br />

boiler<br />

Tg, ac, ref −Tw,<br />

in, ref<br />

where Tg,ac,ref and Tw,in,ref are reference values for temperatures <strong>of</strong> adiabatic<br />

combustion and water entering the boiler. They have been fixed to 2000 and 200 ºC,<br />

although it may be also possible to use the specific values <strong>of</strong> these temperatures instead<br />

<strong>of</strong> these constant values.<br />

The standard method for furnaces calculation allows easy implementation and fast<br />

response that allows to implement the heat transfer models for real time monitoring.<br />

However, it entails a strong simplification. So that, it has been proposed to use a more<br />

sophisticated method to model the furnace (Díez, 2002; Díez et al., 2005; Cortés et al.,<br />

2001). This method is based on the use <strong>of</strong> the Hottel method to modelise radiation,<br />

which implies the discretization <strong>of</strong> the furnace in several sections. Fluid flow and heat


Case study: 3x350 MW coal-fired power plant<br />

release patterns are needed, so that a combustion model and a CFD simulation are used.<br />

CFD calculations are time expensive, so that they are not suitable for some important<br />

applications such as real time monitoring. These authors propose to pre-calculate the<br />

fluid flow for several representative scenarios, and then use this information in real-time<br />

calculations. Hottel method has also been applied to the gases plenum.<br />

The main problem related to heat transfer degradation in boilers is fouling, or<br />

mineral matter deposits in heat transfer surfaces. Fouling in a pulverized coal-fired<br />

boiler can be classified in slagging and fouling (Valero and Cortés, 1996):<br />

-Slagging, referring to the deposition taking place in sections <strong>of</strong> the boiler where<br />

radiant heat transfer is dominant. Most commonly, slagging means ash fouling in the<br />

water walls making up the furnace. Also deposits in radiant platens and first rows <strong>of</strong><br />

pendant superheaters are occasionally included in this category.<br />

-Fouling, defined as deposits on the surface <strong>of</strong> convective (i.e. not seen from the<br />

flame) tube bundles.<br />

These problems can increase when design fuel is modified. To clean the surfaces,<br />

sootblowing is used. However, when slagging has started, sootblowing is not effective<br />

because slagging produces a reduction in heat transfer with leads to higher temperature<br />

at walls surface and, subsequently, more slagging. Due to this situation, load should be<br />

reduced in order to decrease the radiation and to enable a solidification <strong>of</strong> deposits,<br />

which allows its removal. The problem <strong>of</strong> slagging in coal-fired boilers, and specifically<br />

at Teruel power station has been tackled by Cortés et. al. (1989).<br />

Sootblowing usually succeeds in fouling cleaning, but inadequate use <strong>of</strong> this<br />

technique has serious disadvantages. First, if steam is applied to a clean surface, it can<br />

be damaged. Second, steam consumption originates an important reduction <strong>of</strong> power<br />

plant efficiency. For this reason, it is very important to use sootblowing in an optimum<br />

way.<br />

The first step is to monitor the heat transfer in the boiler, which can be used by<br />

applying models such as those previously presented. Fouling monitoring can lead to<br />

improve sootblowing: do not blow according to pre-fixed schedules but taking into<br />

account the state <strong>of</strong> the surfaces. The next step is sootblowing optimization, which looks<br />

for an equilibrium between the efficiency losses produced by sootblowing and<br />

efficiency losses caused by heat transfer reduction. To perform this optimization,<br />

predictive models should be used, capable to determine the probability <strong>of</strong> blow success<br />

for a given fouling conditions and to predict the evolution <strong>of</strong> the fouling along time.<br />

119


Chapter 5<br />

Neural networks are a suitable tools to build-up these models. More details on these<br />

topics and its application to Teruel power plant can be seen in: Cortés et al. (1993),<br />

Bella et al. (1994), and Teruel et al. (2005).<br />

In this thesis, the amount <strong>of</strong> steam has been considered as independent free<br />

<strong>diagnosis</strong> variable. This will allow to see the high influence <strong>of</strong> it on unit efficiency.<br />

5.2.3.4 Air preheaters<br />

As it has been commented in the plant description, at Teruel power plant there are<br />

two pairs <strong>of</strong> preheaters: primary air preheaters (recuperative type) and secondary air<br />

preheaters (regenerative type). In this section, aspects on modelling, working and<br />

suitable free <strong>diagnosis</strong> definitions are commented for these plant components.<br />

The main factors defining a right operation <strong>of</strong> a preheater are fouling and air<br />

leakages. As in boiler heat exchangers, fouling appears when flying ashes are deposed<br />

on heat exchanger surfaces reducing the heat exchange capability. In regenerative type<br />

preheaters, this effect may lead to significant reduction <strong>of</strong> channel areas with important<br />

pressure losses. Besides, chemical fouling may appear if gases reach temperatures lower<br />

than the acid condensation point. To avoid this, cold side temperature should be kept<br />

above a value fixed by the preheater manufacturer.<br />

Air leakages are caused by pressure difference between air and gases circuits.<br />

Although they can appear in all types <strong>of</strong> preheaters, they are relevant in regenerative<br />

preheaters <strong>of</strong> rotating type, where the working principle based on a rotary basket makes<br />

it difficult to achieve a good sealing. An excessive value <strong>of</strong> air leakages implies a<br />

reduction in preheater efficiency and additional work <strong>of</strong> forced and induced fans.<br />

Air leakages also affect the definition <strong>of</strong> thermal efficiency:<br />

120<br />

ε<br />

preheater<br />

( ) ( )<br />

( ) ( )<br />

m ⋅h T −m ⋅h<br />

T<br />

=<br />

m ⋅h T −m ⋅h<br />

T<br />

a, o a a, o a, in a a, in<br />

ao , a gin , ain , a ain ,<br />

For simulation purposes, the calculation <strong>of</strong> this parameter depends on the type <strong>of</strong><br />

preheaters.<br />

For recuperative preheaters, the conventional ε-NTU method is usually applied. The<br />

case <strong>of</strong> regenerative heat exchangers is a bit different, because these types <strong>of</strong> heat<br />

exchangers do not work in stationary state, but in a cyclic operation: an intermediate<br />

heating matrix is alternatively heated by a flow <strong>of</strong> hot fluid and cooled by the cold<br />

stream. To simulate this type <strong>of</strong> heat exchangers, the ε – NTU0 method can be used.<br />

5.86


Case study: 3x350 MW coal-fired power plant<br />

It is supposed that conduction along the matrix axis is negligible in the matrix and in<br />

circulating fluids, and transversal conductivity in the matrix is considered to be very<br />

high. These assumptions are equivalent to suppose that matrix temperature is constant in<br />

each normal section considered. With these hypotheses it is possible to write<br />

expressions similar to those <strong>of</strong> the ε – NTU method but considering the exchange<br />

periods and the heat capacity <strong>of</strong> the matrix material:<br />

⎛ ma cp, a Pa hc Pa C ⎞<br />

mat<br />

ε preheater ε NTU0 , , ,<br />

mgcpg , PghhPgmacpa , Pa<br />

⋅ ⋅ <br />

⋅<br />

= ⎜<br />

⎟<br />

⋅ ⋅ ⋅ ⋅ ⋅ ⎟<br />

⎝<br />

<br />

⎠<br />

5.87<br />

Where Pa and Pg are the exchange periods for air and gases, hc and hh are the<br />

convection coefficients for cold and hot sides, Cmat is the thermal capacity <strong>of</strong> the<br />

material <strong>of</strong> the matrix and NTU0 is the modified number <strong>of</strong> transfer units:<br />

NTU<br />

0<br />

⎛ ⎞<br />

1<br />

⎜<br />

1<br />

⎟<br />

= ⋅⎜<br />

⎟<br />

m 1 1<br />

a⋅cp, a ⎜ ⎟<br />

⎜<br />

+<br />

hh⋅Ah hc⋅A ⎟<br />

⎝ c ⎠<br />

It should be noted that heat exchanger configuration does not appear, because it is<br />

supposed that regenerative preheaters always work at counterflow. Besides, matrix<br />

thermal conductivity is not present because it has been supposed to be zero in<br />

longitudinal direction and infinite in transversal direction. These previous equations<br />

have been developed for static preheaters. However, they can be adapted to rotary type<br />

exchangers by discretizing them in angular sectors. More details on the simulation <strong>of</strong><br />

regenerative preheaters can be seen in Díez (2002).<br />

The parameter proposed as free <strong>diagnosis</strong> variable is a simplified expression <strong>of</strong><br />

efficiency:<br />

ε<br />

preheater<br />

T −T<br />

=<br />

T −T<br />

ao , ain ,<br />

g, in a, in<br />

It should be noted that this equation is the same as that one presented previously if<br />

infiltration is neglected and specific heat capacity <strong>of</strong> air is considered as constant. This<br />

simplified equation has been used because it can be calculated directly from plant data<br />

and is used by plant staff. For a simple modelling <strong>of</strong> heat exchangers, it is possible to<br />

correlate it with plant load.<br />

The other aspect to be considered in preheaters is infiltration. The <strong>diagnosis</strong> variable<br />

to take into account this phenomenon is air leakages. They can be quantified by<br />

measuring oxygen molar fraction in flue gases entering and exiting the preheater:<br />

5.88<br />

5.89<br />

121


Chapter 5<br />

122<br />

r − r<br />

O2, o O2, in<br />

a, leakage = g, in ⋅<br />

5.90<br />

0.21−<br />

rO2,<br />

o<br />

m m<br />

where rO2 are the oxygen molar fractions in flue gases entering and exiting the<br />

preheater.<br />

The previous equations allow to determine leakages in a preheater if oxygen fraction<br />

is known at the outside <strong>of</strong> it. However, in some cases, such as the working example,<br />

oxygen content at outside is only known in one point where flows leaving all the<br />

preheaters. In this case, equation 5.90 can be applied but considering total gas flow and<br />

total leakages. To determine leakages separately in all preheaters, additional<br />

assumptions should be made, such as neglecting leakages in recuperative preheaters.<br />

For these reasons, chosen <strong>diagnosis</strong> variable is the relation between total leakages and<br />

total gas entering all preheaters, which can be decomposed in two terms corresponding<br />

to primary and secondary preheaters:<br />

m m + m<br />

a, leakages, total a, leakages,1 ph a, leakages,2ph Leakages = = = Leakages1ph + Leakages2ph<br />

mgas, in, total mgas,<br />

in, total<br />

5.2.3.5 Ancillary devices<br />

Equipments included in this last section are fans, mills and electrostatic precipitator.<br />

Fans behaviour are represented in characteristic curves provided by the<br />

manufacturer. These curves, usually relate the static pressure increment to the<br />

volumetric flow and an angle:<br />

st st v<br />

( , )<br />

5.91<br />

Δ p =Δp Q β<br />

5.92<br />

Where β is the angle <strong>of</strong> the director blades in radial fans, and the angle <strong>of</strong> fan blades<br />

in axial fans. In radial fans, a set <strong>of</strong> curves indicates the power consumed in function <strong>of</strong><br />

the same parameters:<br />

W = W Q β<br />

5.93<br />

( <br />

v,<br />

)<br />

In axial fans, curves provide the efficiency as function <strong>of</strong> the operation point:<br />

( Q , p )<br />

η = η Δ<br />

5.94<br />

fan fan v st<br />

Operation point depends on both the fan and the installation, and is determined by<br />

crossing their characteristic curves (such as in pumps). ηfan, power consumed and<br />

operation point are related by the following expression:


Case study: 3x350 MW coal-fired power plant<br />

Δptot ⋅Q<br />

v<br />

η fan =<br />

5.95<br />

W<br />

Ancillary devices have not been studied in this work, because gross electric power<br />

and efficiency are considered. The impact in fuel <strong>of</strong> each additional kW consumed by<br />

them is directly the gross heat rate <strong>of</strong> the unit. These extra consumptions can be caused<br />

not only by a malfunction in the component itself but also by a malfunction in other<br />

parts <strong>of</strong> the plant that make these devices to work more. For example, additional<br />

electricity consumptions by fans can be due to a degradation in the fans or to a higher<br />

pressure loss in the boiler. In the same way, increment in electricity consumed by<br />

electrostatic precipitators can be caused not only by a failure in them but also to an<br />

increment in flying ashes in flue gases. Finally, electricity consumed by mills depends<br />

on both internal causes and coal properties.<br />

Analysis <strong>of</strong> these relations requires <strong>analysis</strong> <strong>of</strong> the behaviour and modelling <strong>of</strong> these<br />

components. This problem has been tackled in general and specifically for Teruel<br />

problem plant by Arauzo et al. (Arauzo et al., 1993; Arauzo and Cortés 1995a, 1997).<br />

Díez (2002) has adapted the methodology <strong>of</strong> fan modelling for the new primary air fans.<br />

Relation between ancillary devices operation and boiler has also been studied by<br />

applying a symptom-problem scheme (Arauzo and Cortés, 1995b).<br />

5.3 Monitoring and <strong>diagnosis</strong> model for Teruel power plant.<br />

In the first part <strong>of</strong> this section, used measurements are reviewed, because they are<br />

going to determine the depth <strong>of</strong> the <strong>diagnosis</strong> achievable. According to this information<br />

and to the guidelines presented in the previous paragraph, free <strong>diagnosis</strong> variables are<br />

presented. Finally, global efficiency indicators defined are presented.<br />

5.3.1 Measurement review.<br />

In table 5.6, a list <strong>of</strong> measurements used to monitor the power plant is presented.<br />

They are ordered by areas: ambient conditions, water-steam circuit (according to fluid<br />

flow and included cooling system) and boiler.<br />

123


Chapter 5<br />

Number Description<br />

124<br />

1 Ambient temperature<br />

2 Dew point<br />

3 Ambient pressure<br />

4 Wind speed<br />

5 Main steam temperature right side<br />

6 Main steam temperature left side<br />

7 Main steam pressure right side<br />

8 Main steam pressure left side<br />

9 Cold reheated steam temperature right side<br />

10 Cold reheated steam temperature left side<br />

11 Cold reheated steam pressure right side<br />

12 Cold reheated steam pressure left side<br />

13 Hot reheated steam temperature right side<br />

14 Hot reheated steam temperature left side<br />

15 Hot reheated steam pressure right side<br />

16 Hot reheated steam pressure left side<br />

17 Temperature <strong>of</strong> steam extraction for 5 th heater<br />

18 Pressure <strong>of</strong> steam extraction for 5 th heater<br />

19 Temperature <strong>of</strong> steam extraction for deaerator<br />

20 Pressure <strong>of</strong> steam extraction for deaerator<br />

21 Gross power<br />

22 Condenser absolute pressure<br />

23 Temperature <strong>of</strong> cold cooling water right side<br />

24 Temperature <strong>of</strong> cold cooling water left side<br />

25 Temperature <strong>of</strong> hot cooling water right side<br />

26 Temperature <strong>of</strong> hot cooling water left side<br />

27 Temperature <strong>of</strong> water entering the deaerator<br />

28 Flow rate <strong>of</strong> water entering the deaerator<br />

Table 5.6.A. List <strong>of</strong> plant measurements used.


Number Description<br />

29 Mass flow <strong>of</strong> water from dribbling tank to deaerator<br />

30 Pressure <strong>of</strong> water entering the turbo-pump<br />

31 Pressure <strong>of</strong> water leaving the turbo-pump<br />

32 Temperature <strong>of</strong> the drain <strong>of</strong> 5 th heater<br />

33 Temperature <strong>of</strong> water leaving the 5 th heater<br />

34 Temperature <strong>of</strong> the drain <strong>of</strong> 6 th heater<br />

35 Temperature <strong>of</strong> water leaving the 6 th heater<br />

36 Steam for sootblowing flow rate<br />

37 Temperature <strong>of</strong> air entering primary preheater side A<br />

38 Temperature <strong>of</strong> air entering primary preheater side B<br />

39 Temperature <strong>of</strong> air leaving primary preheater side A<br />

40 Temperature <strong>of</strong> air leaving primary preheater side B<br />

41 Temperature <strong>of</strong> gases entering primary preheater side A<br />

42 Temperature <strong>of</strong> gases entering primary preheater side B<br />

43 Temperature <strong>of</strong> gases leaving primary preheater side A<br />

44 Temperature <strong>of</strong> gases leaving primary preheater side B<br />

45 Temperature <strong>of</strong> air leaving secondary coil heater side A<br />

46 Temperature <strong>of</strong> air leaving secondary coil heater side B<br />

47 Temperature <strong>of</strong> air leaving secondary preheater side A<br />

48 Temperature <strong>of</strong> air leaving secondary preheater side B<br />

Case study: 3x350 MW coal-fired power plant<br />

49 Temperature <strong>of</strong> gases entering secondary preheater side A<br />

50 Temperature <strong>of</strong> gases entering secondary preheater side B<br />

51 Temperature <strong>of</strong> gases leaving secondary preheater side A<br />

52 Temperature <strong>of</strong> gases leaving secondary preheater side B<br />

53 Temperature <strong>of</strong> gases leaving induced draft fans<br />

54 Primary air mass flow side A<br />

55 Primary air mass flow side B<br />

56 Secondary air mass flow side A<br />

Table 5.6.B. List <strong>of</strong> plant measurements used.<br />

125


Chapter 5<br />

Number Description<br />

126<br />

57 Secondary air mass flow side B<br />

58 Tempering air mass flow side A<br />

59 Tempering air mass flow side B<br />

60 Air to flue gas desulfuration mass flow side A<br />

61 Air to flue gas desulfuration mass flow side B<br />

62 Oxygen in flue gases leaving the boiler side A<br />

63 Oxygen in flue gases leaving the boiler side B<br />

64 Oxygen in flue gases leaving induced draft fans<br />

65 Natural gas flow<br />

66 Coal high heating value<br />

67 Carbon mass fraction in coal<br />

68 Hydrogen mass fraction in coal<br />

69 Oxygen mass fraction in coal<br />

70 Nitrogen mass fraction in coal<br />

71 Sulphur mass fraction in coal<br />

72 Ash mass fraction in coal<br />

73 Moisture mass fraction in coal<br />

74 Carbon in ash<br />

Table 5.6.C. List <strong>of</strong> plant measurements used.<br />

As it can be seen, there is an important lack <strong>of</strong> information in the low pressure zone<br />

<strong>of</strong> the steam cycle and in the boiler. This fact is going to limit the depth <strong>of</strong> the <strong>diagnosis</strong><br />

in these areas, and is going to lead to the use <strong>of</strong> aggregated variables and models. On the<br />

other hand, there are local redundancies in some areas: temperatures in air heaters, mass<br />

flows <strong>of</strong> primary, secondary and tempering air (one is redundant), and regenerative<br />

water heaters. Although several techniques are available to solve this fact, for example<br />

by minimizing a global uncertainty function (Correas, 2001), here, these variables have<br />

been used to pre-calculate several parameters that are going to be introduced in the<br />

resolution <strong>of</strong> the plant: for example, instead <strong>of</strong> introducing the flows <strong>of</strong> primary and<br />

secondary air, the relation between them is used, thus eliminating one variable.


Case study: 3x350 MW coal-fired power plant<br />

All these variables are measured by plant instruments and are available in the plant<br />

information system, except results coming from laboratory (coal and ash <strong>analysis</strong>),<br />

which corresponds to daily analyses.<br />

5.3.2 Free <strong>diagnosis</strong> variables<br />

The aim <strong>of</strong> this section is to present the <strong>diagnosis</strong> variables used and how they are<br />

calculated in the performance test and in the simulation problems, in order to clarify the<br />

differences and common points between the three problems.<br />

Free <strong>diagnosis</strong> variables appear in table 5.7. They are ordered by categories: ambient<br />

conditions (A), fuel (F), set points (SP) and component parameters (CP). It should be<br />

noted that this classification is only indicative, and it is subject to discussion. For<br />

example, the amount <strong>of</strong> natural gas is fixed by a set-point but is included in fuel, and air<br />

for desulfuration unit is not strictly a set point because the set point is the temperature <strong>of</strong><br />

flue gases to stack.<br />

The column <strong>of</strong> “calculation in performance test” summarizes the differences<br />

between the <strong>diagnosis</strong> and performance test problems. Most <strong>of</strong> variables are<br />

measurements or can be calculated directly from them. In this case, the equation is the<br />

same for <strong>diagnosis</strong> and performance test problems. It should be noted that prior to the<br />

main equations system, a module to pre-calculate parameters directly from plant signals<br />

is used. On the other hand, there are other variables that cannot be easily obtained from<br />

plant data. In this case, another parameter is fixed instead <strong>of</strong> the value <strong>of</strong> the <strong>diagnosis</strong><br />

variable considered. For simulation purposes, equations are the same as for <strong>diagnosis</strong>,<br />

but the value <strong>of</strong> free <strong>diagnosis</strong> variables corresponding to components parameters is<br />

determined by suitable correlations that should be tuned to the component<br />

characteristics. Formally, this is a minor difference, but in practice this is the more<br />

complex and time consuming issue.<br />

In table 5.8, more <strong>diagnosis</strong> variables are included. They are formulated in the same<br />

way, however, due to the lack <strong>of</strong> information, they have a constant value, so that they<br />

do not provide any impact.<br />

127


Chapter 5<br />

Num. Description units Calculation in<br />

performance<br />

test<br />

128<br />

Category<br />

1 Ambient temperature ºC Direct A<br />

2 Relative humidity % Direct A<br />

3 Wind speed m/s Direct A<br />

4 Coal high heating value kJ/kg Direct F<br />

5 Carbon mass fraction in coal % Direct F<br />

6 Hydrogen mass fraction in coal % Direct F<br />

7 Moisture mass fraction in coal % Direct F<br />

8 Ash mass fraction in coal % Direct F<br />

9 Sulphur mass fraction in coal % Direct F<br />

10 Nitrogen mass fraction in coal % Direct F<br />

11 Energy provided by natural gas % Natural gas<br />

flow rate<br />

12 Main steam temperature ºC Direct SP<br />

13 Reheated steam temperature ºC Direct SP<br />

14 Main steam pressure bar Direct SP<br />

15 Gross electric power MW Direct SP<br />

16 Oxygen in flue gases leaving the boiler % Direct SP<br />

17 Average cold-side temperature in<br />

secondary air preheaters<br />

18 Average cold-side temperature in<br />

primary air preheaters<br />

ºC Direct SP<br />

ºC Direct SP<br />

19 Sootblowing steam flow rate kg/s Direct SP<br />

20 Air for flue gas desulfuration unit kg/s Direct SP<br />

21 Tempering and primary air relation kg/kg Direct SP<br />

22 Primary air-coal ratio kg/kg Primary and<br />

secondary air<br />

relation<br />

Table 5.7.A. Free <strong>diagnosis</strong> variables.<br />

F<br />

SP


Case study: 3x350 MW coal-fired power plant<br />

Num. Description units Calculation in<br />

performance<br />

test<br />

23 Temperature difference <strong>of</strong> flue gases<br />

entering preheaters<br />

Category<br />

ºC Direct SP<br />

24 Fraction <strong>of</strong> flue gases through PAH - Relation <strong>of</strong><br />

flue gases<br />

temperature<br />

drop in PAH<br />

and SAH<br />

25 High pressure turbine isoentropic<br />

efficiency<br />

26 Intermediate pressure 1 isoentropic<br />

efficiency<br />

27 Intermediate pressure 2 isoentropic<br />

efficiency<br />

SP<br />

% Direct CP<br />

% Direct CP<br />

% Direct CP<br />

28 Low pressure isoentropic efficiency % Water entering<br />

the deaerator<br />

flow rate.<br />

29 Intermediate<br />

coefficient<br />

pressure 1 flow<br />

30 Intermediate<br />

coefficient<br />

pressure 2 flow<br />

kg 1/2 ·m 3/2 ·<br />

s -1 ·bar -1/2<br />

kg 1/2 ·m 3/2 ·<br />

s -1 ·bar -1/2<br />

31 Low pressure flow coefficient kg 1/2 ·m 3/2 ·<br />

s -1 ·bar -1/2<br />

CP<br />

Pressure CP<br />

Pressure CP<br />

Pressure CP<br />

32 TTD 6 th water heater ºC Direct CP<br />

33 TTD 5 th water heater ºC Direct CP<br />

34 TTD 3 rd water heater ºC Direct. CP<br />

Table 5.7.B. Free <strong>diagnosis</strong> variables.<br />

129


Chapter 5<br />

Num. Description units Calculation in<br />

performance<br />

test<br />

130<br />

Category<br />

35 TDCA 6 th water heater ºC Direct CP<br />

36 TDCA 5 th water heater ºC Direct CP<br />

37 Pressure drop in reheater bar Direct CP<br />

38 Pressure increment in turbo-pump bar Direct. CP<br />

39 Water losses kg/s Direct CP<br />

40 Condenser effectiveness - Condenser CP<br />

41 Cooling water flow rate kg/s pressure, CP<br />

42 Cooling tower effectiveness - Cooling water<br />

inlet and outlet<br />

temperatures<br />

CP<br />

43 Secondary air preheater effectiveness - Direct CP<br />

44 Primary air preheater effectiveness - Direct CP<br />

45 Aggregated boiler effectiveness - Temperature<br />

<strong>of</strong> flue gases<br />

leaving IDF<br />

46 Air infiltration in preheaters % Oxygen in flue<br />

gases leaving<br />

IDF<br />

47 Carbon in ashes % Direct CP<br />

Table 5.7.C. Free <strong>diagnosis</strong> variables.<br />

CP<br />

CP


Case study: 3x350 MW coal-fired power plant<br />

Num. Description Value Units Category<br />

48 Natural gas high heating value 43488 kJ/Nm 3 F<br />

49 Methane in natural gas 91.54 % F<br />

50 Ethane in natural gas 5.27 % F<br />

51 Propane in natural gas 1.99 % F<br />

52 Temperature <strong>of</strong> steam leaving<br />

secondary superheater<br />

53 Air from the output <strong>of</strong> primary<br />

heater entering secondary heater<br />

480 ºC SP<br />

0 kg/s SP<br />

54 Air leakage in primary air heater 0 kg/s CP<br />

55 CO in flue gases leaving the boiler 130 ppm CP<br />

Table 5.8. Free <strong>diagnosis</strong> variables with constant value.<br />

5.3.3 Global efficiency indicators.<br />

Although the theoretical development has been made for only one global efficiency<br />

indicator, it is possible to define as much indicators as the analyst considers. In this<br />

work, three parameters have been considered, according to the usual practice in the<br />

power plant: gross unit heat rate, gross cycle heat rate and boiler efficiency.<br />

Gross unit heat rate is the relation <strong>of</strong> energy <strong>of</strong> fuel entering the plant and power<br />

produced by the alternator:<br />

H<br />

f<br />

HRu =<br />

W<br />

e<br />

Gross cycle heat rate is the relation <strong>of</strong> energy provided to the cycle and electric<br />

power produced:<br />

HR<br />

c<br />

U + U<br />

=<br />

W<br />

boiler− cycle others<br />

e<br />

where Uothers takes into account other minor energy sources: water from dribbling<br />

tank and vent from blow down tank entering the deaerator, and make-up water entering<br />

the condenser. Finally, boiler efficiency is calculated as:<br />

5.96<br />

5.97<br />

U L<br />

η boiler = = 1−<br />

5.44<br />

E E<br />

131


Chapter 5<br />

although it is usually expressed in %. More details can be seen in section 5.2.3,<br />

Rangel (2005) or Uche (2001).<br />

5.4 Conclusion<br />

This chapter should be considered as an example <strong>of</strong> methodology to be followed in<br />

order to define a <strong>diagnosis</strong> system. First <strong>of</strong> all, the power plant has to be analyzed.<br />

Second, information regarding models, degradation mechanisms and operational<br />

problems has to be collected for the main plant components. Afterwards, available<br />

measurements should be considered, in order to determine the depth <strong>of</strong> <strong>diagnosis</strong><br />

achievable. Finally, all this information has to be used to define free <strong>diagnosis</strong> variables<br />

and global efficiency indicator(s).<br />

It should be noted that the process is not finished until the system is installed and<br />

tested with plant data. The detail <strong>of</strong> the <strong>diagnosis</strong> required is not only determined by<br />

previous experiences and theory but also by the actual situation <strong>of</strong> the system to be<br />

analyzed, which is not known until the <strong>diagnosis</strong> is performed. In other words, this<br />

design is not usually straightforward but revisions are needed.<br />

Information on the data retrieval system and instrumentation accuracy can be seen in<br />

Rangel (2005). Results <strong>of</strong> the application <strong>of</strong> a <strong>diagnosis</strong> system in the working example<br />

are reported in next chapter.<br />

132


6 Application <strong>of</strong> quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong><br />

In this chapter, results <strong>of</strong> anamnesis or the repeated <strong>diagnosis</strong> <strong>of</strong> the Teruel power<br />

plant are presented. This technique allows to see the evolution <strong>of</strong> degradation indicators<br />

with its dispersion, and helps to overcome errors induced by instrumentation uncertainty<br />

and the <strong>diagnosis</strong> method itself. After a brief introduction to present the main<br />

assumptions made, evolution <strong>of</strong> the three efficiency indicators considered (unit and<br />

cycle heat rate and boiler efficiency) is shown. Then, the influence <strong>of</strong> all free <strong>diagnosis</strong><br />

variables is analyzed, demonstrating the capability <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to<br />

quantify the impact caused by each one <strong>of</strong> them. Afterwards, error produced by the<br />

method is studied by analysing the residual term (the part <strong>of</strong> the deviation in the global<br />

efficiency indicator which is not assigned to free <strong>diagnosis</strong> variables variations). Results<br />

show that this term is negligible. Finally, techniques such as impact addition and<br />

<strong>causal</strong>ity chain <strong>analysis</strong> are applied to facilitate the interpretation <strong>of</strong> the <strong>diagnosis</strong> results<br />

and to filtrate crossed relation between free <strong>diagnosis</strong> variables.<br />

6.1 Introduction<br />

The <strong>diagnosis</strong> system presented in the previous chapter has been used to diagnose<br />

the three units <strong>of</strong> Teruel power plant for a time span <strong>of</strong> more than six years: from 1 st<br />

January 1999 to 20 th March 2005. Information needed to perform the <strong>diagnosis</strong> has been<br />

obtained from the data base <strong>of</strong> the plant information system, where average daily<br />

measurements classified by load levels were available. Only full load level (more than<br />

320 MW) has been considered. In total, 4570 operation days have been studied: 1539 <strong>of</strong><br />

unit 1, 1495 <strong>of</strong> unit 2 and 1536 <strong>of</strong> unit 3.<br />

133


Chapter 6<br />

Since the interest <strong>of</strong> the study is the <strong>analysis</strong> <strong>of</strong> the evolution <strong>of</strong> free <strong>diagnosis</strong><br />

variables and their impacts, the choice <strong>of</strong> a reference state is a secondary issue, which<br />

depends on the preferences <strong>of</strong> the analyst. In fact, results obtained by using one<br />

reference state or another, would very similar but the graphs <strong>of</strong> impacts would be<br />

biased. When a new plant is studied, design conditions are usually used as reference.<br />

However, the age <strong>of</strong> the plant considered advises to use a more recent reference, such as<br />

one <strong>of</strong> the days analyzed. According to plant staff and to the results obtained by<br />

preliminary analyses, unit 3 during September 2003 represents a good period to be used<br />

as reference, because most free <strong>diagnosis</strong> variables have a value which can be<br />

considered as representative <strong>of</strong> the plant, neither too high nor too low. For this reason,<br />

the state <strong>of</strong> this unit on 17 th September 2003 has been chosen.<br />

For each one <strong>of</strong> the 47 free <strong>diagnosis</strong> variables considered, seven graphs may be<br />

studied: variable evolution, the evolution <strong>of</strong> the impact on the three efficiency indicators<br />

considered and relation between these three impacts and the variation <strong>of</strong> the free<br />

variable. In each one <strong>of</strong> these graphs, results for the three groups are plotted by using<br />

different colours. This multiplicity <strong>of</strong> graphs makes non practical to include all <strong>of</strong> them<br />

in this document. Information contained in them can be summarized by using several<br />

parameters.<br />

Besides the use <strong>of</strong> indicators to quantify the variability <strong>of</strong> the free <strong>diagnosis</strong><br />

variables and their impacts is needed. Standard deviations are proposed for this task.<br />

They allow to resume the variation <strong>of</strong> free <strong>diagnosis</strong> variables and the importance <strong>of</strong> the<br />

impacts caused by them. An impact with high standard deviation indicates that its<br />

corresponding free <strong>diagnosis</strong> variable has influenced strongly the evolution <strong>of</strong> the<br />

efficiency indicator during the time span studied. The use <strong>of</strong> these parameters has the<br />

advantage <strong>of</strong> avoiding the influence <strong>of</strong> the reference state chosen. They are defined as<br />

follows:<br />

134<br />

σ<br />

σ<br />

xdi<br />

,<br />

Ii<br />

, ηb<br />

=<br />

=<br />

n<br />

p<br />

j<br />

∑(<br />

xdi , − xdi<br />

, )<br />

j=<br />

1<br />

n<br />

n<br />

p<br />

p<br />

j<br />

∑(<br />

Ii, η − I , )<br />

b iηb<br />

j=<br />

1<br />

n<br />

p<br />

2<br />

2<br />

6.1<br />

6.2


σ<br />

σ<br />

IiHR<br />

, c<br />

IiHR<br />

, u<br />

=<br />

=<br />

n<br />

p<br />

j<br />

∑(<br />

IiHR , − I , )<br />

c iHRc<br />

j=<br />

1<br />

n<br />

n<br />

p<br />

j<br />

∑(<br />

IiHR , − I , )<br />

u iHRu<br />

j=<br />

1<br />

n<br />

p<br />

p<br />

2<br />

2<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

where σ is the standard deviation <strong>of</strong> a free <strong>diagnosis</strong> variable ( x di , ), or <strong>of</strong> its impacts<br />

on boiler efficiency ( Ii, η ), steam cycle heat rate ( I<br />

b<br />

iHR , ) or unit heat rate (<br />

c<br />

iHR , u<br />

6.3<br />

6.4<br />

I ), which<br />

are calculated by using the individual (j) and average values <strong>of</strong> these parameters in a set<br />

<strong>of</strong> n p points. These sets can be the points <strong>of</strong> a unit, or all the points <strong>of</strong> the three units<br />

together. When no unit is specified, parameters are calculated by using all the units <strong>of</strong><br />

the plant.<br />

Other interesting aspect is the relation between the variation <strong>of</strong> a free <strong>diagnosis</strong><br />

variable ( Δ xdi<br />

, ) and the impact which it produces on the boiler efficiency and cycle and<br />

unit heat rates ( Ii, η , I<br />

b iHR , and<br />

c iHR , u<br />

correspond to the elements <strong>of</strong> vector ed.<br />

I<br />

edi,<br />

η b x<br />

= Δ<br />

ed<br />

ed<br />

iHR , c<br />

iHR , u<br />

i,<br />

ηb<br />

di ,<br />

I<br />

=<br />

Δx<br />

iHR , c<br />

I<br />

=<br />

Δx<br />

di ,<br />

iHR , u<br />

di ,<br />

I ). For a simple <strong>diagnosis</strong>, these impact factors<br />

These factors are very important because they constitute the relation between free<br />

<strong>diagnosis</strong> variable increment and impact, which is the key <strong>of</strong> the <strong>diagnosis</strong><br />

methodology. However, they are associated to only one point. Since now there are<br />

several points, it is preferable to calculate average values by obtaining the slopes <strong>of</strong> the<br />

graphs which represent impacts versus variation <strong>of</strong> free <strong>diagnosis</strong> variables. These<br />

slopes are obtained by calculating the linear regression <strong>of</strong> the graph. To remark this fact,<br />

6.5<br />

6.6<br />

6.7<br />

135


Chapter 6<br />

av av<br />

av<br />

the superscript av can be added. In this way, edi, η , ed<br />

b iHR , and<br />

c<br />

iHR , u<br />

136<br />

ed stand for<br />

calculated slope <strong>of</strong> the curves that relate the impacts on boiler efficiency and cycle and<br />

unit heat rates produced by the variation <strong>of</strong> the free <strong>diagnosis</strong> variable x di , to the<br />

increment <strong>of</strong> this variable.<br />

In graphs relating impacts to free variables increments, not only the slope presented<br />

above has to be considered. Besides, the degree <strong>of</strong> non linearity or dispersion <strong>of</strong> the<br />

points is also important. To take into account this fact, the linear regression coefficients<br />

(R 2 ) are also considered. These parameters reach the value 1 when the correlation is<br />

perfectly lineal, and decrease when the graphs curve or when dispersion appears. If all<br />

<strong>of</strong> them were 1, relations would be perfectly known and constant values <strong>of</strong> ed could be<br />

used. However, these relations are not always so perfect, and it is worth to use the<br />

<strong>diagnosis</strong> methodology to calculate ed for each point. This task is detailed in section<br />

6.4.<br />

Once both assumptions made for performing the anamnesis and definitions and<br />

importance <strong>of</strong> some parameters have been presented, in the next section the evolution <strong>of</strong><br />

the three efficiency indicators is shown.<br />

6.2 Evolution <strong>of</strong> efficiency indicators<br />

The three efficiency indicators considered are unit and steam cycle gross heat rate<br />

and boiler efficiency. Units <strong>of</strong> heat rate are sometimes kcal/kWh, although in this work<br />

are kJ/kJ (non-dimensional). Boiler efficiency is expressed in %.<br />

Figure 6.1 shows the evolution <strong>of</strong> unit gross heat rates. During the first three<br />

months, unit heat rates are high. Then they decrease dramatically and then increase<br />

steadily for three years and stand more or less flat for other three. Finally, they decrease<br />

sharply again and tend to increase. Apart from these common trends, some differences<br />

among the three units can be observed. During the first two years, unit two has higher<br />

heat rate. Then, in the last three years, unit three is the best and then comes unit two.<br />

However, in the last months, unit one heat rate is reduced substantially, so that the<br />

values <strong>of</strong> the three units become quite close. Although some general points can be<br />

observed, the graph is characterized by a high short-term dispersion (width <strong>of</strong> the band<br />

<strong>of</strong> points) <strong>of</strong> about 0.05, which makes difficult to detect tendencies. Besides, some


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

points too high or too low (outliers) are present. A first step towards a suitable<br />

decomposition <strong>of</strong> unit heat rate evolution is the use <strong>of</strong> cycle heat rate and boiler<br />

efficiency.<br />

3.10<br />

3.05<br />

3.00<br />

2.95<br />

2.90<br />

2.85<br />

2.80<br />

2.75<br />

Unit heat rate<br />

2.70<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.1. Unit heat rate evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Evolution <strong>of</strong> steam cycle gross heat rate is shown in figure 6.2. This variable<br />

oscillates with yearly period around the value <strong>of</strong> 2.4 due to the effect <strong>of</strong> ambient<br />

temperature. In summer, due to higher temperatures condenser pressure increases and<br />

cycle heat rate also does. Besides this stationary oscillation <strong>of</strong> about 0.1, there is a short<br />

term dispersion <strong>of</strong> less than 0.05. Some outliers are also present, probably caused by<br />

measurement errors or corresponding to days in which the unit affected has operated in<br />

the full load range only for a short time. An example <strong>of</strong> these out <strong>of</strong> range values are the<br />

points <strong>of</strong> unit 2 in the summer <strong>of</strong> 1999. Fortunately, due to the big amount <strong>of</strong> days<br />

available, the presence <strong>of</strong> these points does not constitute a big problem.<br />

Apart from this stationary oscillation and dispersion, a long time evolution can be<br />

observed. During the first months, the heat rate is quite high, and then efficiency<br />

improves suddenly. Afterwards, due to component degradation, heat rate increases for<br />

two years and tends to remain constant for other two. Finally, around the end <strong>of</strong> 2003, it<br />

decreases again. Some differences among the three units can also be observed. During<br />

137


Chapter 6<br />

the first three years, all units show similar efficiency. However, in the last three years it<br />

can be seen how unit three works better and unit one works worse.<br />

138<br />

2.60<br />

2.55<br />

2.50<br />

2.45<br />

2.40<br />

2.35<br />

2.30<br />

2.25<br />

Steam cycle heat rate<br />

2.20<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.2. Steam cycle heat rate evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

In figure 6.3, the evolution <strong>of</strong> boiler efficiency for the time span considered is shown<br />

for the three units <strong>of</strong> the power plant. It can be seen how the graph oscillates around<br />

83% with amplitude <strong>of</strong> almost 2 points. This oscillation corresponds to the seasons <strong>of</strong><br />

summer and winter: efficiency is higher in summer because losses associated with hot<br />

gases leaving the boiler decrease. There is also a short-term dispersion with amplitude<br />

<strong>of</strong> around 1 point. This dispersion makes difficult to identify whether one unit works<br />

better than the others. It seems that all boilers work worse in the first two years, but,<br />

without a <strong>diagnosis</strong> procedure it is difficult to assure this fact and, <strong>of</strong> course, to detect<br />

its cause.<br />

As it has been seen, evolution <strong>of</strong> the global efficiency indicators is complex and, due<br />

to crossed influence <strong>of</strong> the variation in ambient conditions, sometimes is difficult even<br />

to determine whether a unit tends to work worse or to remain steadily. The use <strong>of</strong> cycle<br />

heat rate and boiler efficiency entails a first improvement compared to unit heat rate, but<br />

obviously they are not enough. In conclusion, a <strong>diagnosis</strong> methodology is needed in<br />

order to decompose these complex patterns into the causes that originate them. This is<br />

the task <strong>of</strong> the next section.


%<br />

87<br />

86<br />

85<br />

84<br />

83<br />

82<br />

81<br />

80<br />

6.3 Diagnosis results <strong>analysis</strong><br />

Boiler efficiency<br />

Figure 6.3. Boiler efficiency evolution.<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

79<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

In this section, evolution <strong>of</strong> the free <strong>diagnosis</strong> variables and their impacts on the<br />

three global efficiency indicators is analyzed, as well as relations among free variable<br />

increments and the impacts they produce. Due to the multiplicity <strong>of</strong> graphs available (7<br />

per each one <strong>of</strong> the 47 free <strong>diagnosis</strong> variables), not all <strong>of</strong> them are shown. Instead, the<br />

value <strong>of</strong> the indicators presented in the previous sections is <strong>of</strong>ten given.<br />

First, free <strong>diagnosis</strong> variables related to ambient conditions are shown. Second,<br />

variables related to fuel quality. Afterwards, importance <strong>of</strong> set points is studied, first in<br />

steam cycle and second in boiler. Finally, variables related to equipment efficiency are<br />

presented: steam cycle, cooling system and boiler. At the end, indicators corresponding<br />

to all the variables are summarized in a table.<br />

6.3.1 Ambient conditions.<br />

Ambient conditions are temperature, relative humidity and wind speed. Their main<br />

characteristic is that they are not controllable.<br />

139


Chapter 6<br />

The main ambient condition is ambient temperature. It affects both steam cycle<br />

(through cooling system and condenser pressure) and boiler (due to the variation <strong>of</strong><br />

losses associated with flue gases). It has another important effect related to the steam<br />

consumption <strong>of</strong> the coil heaters that preheat air entering the preheaters: in cold weather,<br />

more steam is needed to keep the temperature set point.<br />

Figure 6.4 shows the evolution <strong>of</strong> this variable for the three units, as well as the<br />

value <strong>of</strong> ambient temperature at the reference condition. The climatic characteristics <strong>of</strong><br />

the plant emplacement makes this variable to oscillate from less than 5 degrees in winter<br />

to more than 30 in summer, which is a strong variation above all taking into account<br />

that we are working with daily averages. The average value <strong>of</strong> this variable is 15.6 ºC<br />

and its standard deviation is 7.30 ºC.<br />

ºC<br />

140<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Ambient temperature evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.4. Ambient temperature evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Variation <strong>of</strong> ambient temperature induces a strong impact on steam cycle heat rate,<br />

as it can be seen in figure 6.5. Importance <strong>of</strong> this impact is summarized in a standard<br />

deviation <strong>of</strong> 0.0528, which is a quite high value, as it will be seen later when standard<br />

deviations <strong>of</strong> impacts caused by other variables are analysed. At first sight, it may seem<br />

strange that this impact graph, like many others, is not centred. The vertical position <strong>of</strong><br />

an impact graph depends on the average value <strong>of</strong> the free <strong>diagnosis</strong> variable and its


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

value for the reference state. If they are the same, the graph is centred, if not, it is<br />

displaced upwards or downwards.<br />

0.10<br />

0.05<br />

-0.05<br />

-0.10<br />

-0.15<br />

-0.20<br />

Cycle heat rate impact due to ambient temperature<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.5. Cycle heat rate impact due to ambient temperature.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

The influence on boiler efficiency is also important as it can be seen in figure 6.6.<br />

Standard deviation <strong>of</strong> this impact is 0.402 (%).<br />

Finally, the impact on unit heat rate also shows the same cyclic behaviour but with<br />

less amplitude than the cycle (figure 6.7), because effects on boiler partially compensate<br />

effects <strong>of</strong> condenser pressure. Besides, the use <strong>of</strong> steam for pre-heating the air entering<br />

the preheaters implies an additional increment in unit heat rate at low temperatures.<br />

Standard deviation <strong>of</strong> this impact is 0.0448, which is lower than that <strong>of</strong> the cycle heat<br />

rate.<br />

141


Chapter 6<br />

%<br />

142<br />

0.10<br />

0.05<br />

-0.10<br />

-0.15<br />

1.00<br />

0.50<br />

-0.50<br />

-1.00<br />

-1.50<br />

Boiler efficiency impact due to ambient temperature<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.6. Impact <strong>of</strong> ambient temperature on boiler efficiency.<br />

Unit heat rate impact due to ambient temperature<br />

0.00<br />

24/07/1998 0:00<br />

-0.05<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00 Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.7. Unit heat rate impact due to ambient temperature.<br />

Unit 1<br />

Unit 2<br />

Unit 3


Impact<br />

Ambient temperature and its impact on cycle heat rate<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0<br />

-25 -20 -15 -10 -5 0 5 10 15<br />

Figure 6.8. Ambient temperature variation and its impact on cycle heat rate.<br />

Figure 6.8 shows the relation betwen the variation <strong>of</strong> ambient temperature and the<br />

impact produced in cycle heat rate. The slope <strong>of</strong> the graph (ed av ) is 0.00723 ºC -1 , and the<br />

regression coefficient (R 2 ) is 0.9998. This very high value <strong>of</strong> R 2 corresponds to a graph<br />

where the points are aligned, such as the one seen here. In the graph <strong>of</strong> impact on boiler<br />

efficiency, the slope is 0.0552 %/ºC and R 2 = 0.9957, which means a bit more<br />

dispersion. Finally, the impact on unit heat rate has a slope <strong>of</strong> 0.00614 ºC -1 and also a<br />

very high value <strong>of</strong> R 2 (0.999). These high values <strong>of</strong> R 2 indicate that constant values <strong>of</strong><br />

the impact factor might be used with low error, which is not true in general, as it will be<br />

seen later.<br />

Another weather-related free <strong>diagnosis</strong> variable is relative humidity. This variable<br />

has an average <strong>of</strong> 63.4 % and a standard deviation <strong>of</strong> 7.08 (%). It has also a stationary<br />

oscillation (more humidity in summer), but with more short term dispersion than<br />

temperature. This behaviour can be seen in the evolution <strong>of</strong> the impact on cycle heat<br />

rate, plotted in figure 6.9. It should be noted that dispersion is lower in the first two<br />

years because wet bulb temperature measurement was not reliable and a validation in<br />

relation to ambient temperature was used. The standard deviation <strong>of</strong> this impact is<br />

0.00784, which is seven times lower than that <strong>of</strong> the ambient temperature. Impact <strong>of</strong> this<br />

0.1<br />

0.05<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

Ambient temperature (ºC)<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

143


Chapter 6<br />

variable on boiler efficiency is reduced, and has a standard deviation <strong>of</strong> 0.0108 %. For<br />

this reason, impact on unit heat rate is almost parallel to that on cycle heat rate; it has a<br />

standard deviation <strong>of</strong> 0.00931.<br />

144<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

-0.03<br />

Impact on cycle heat rate due to relative humidity<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Impact<br />

Figure 6.9. Cycle heat rate impact due to relative humidity.<br />

Relative humidity and its impact on cycle heat rate<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

-20 -15 -10 -5 0 5 10 15 20 25 30<br />

-0.01<br />

-0.02<br />

-0.03<br />

Relative humidity (%)<br />

Figure 6.10. Relative humidity variation and its impact on cycle heat rate.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Figure 6.10 shows the relation between relative humidity variation and its impact on<br />

cycle heat rate, where it can be seen how the graph curves slightly. It should be noted<br />

that this non linear shape could not be obtained if all the derivatives were calculated<br />

only in the reference state. Slope <strong>of</strong> the curve is 0.00107 (1/%) and R 2 = 0.992.<br />

The graph relating relative humidity to its impact on unit heat rate is quite similar.<br />

Slope is 0.00127 (1/%) and R 2 = 0.992. Relation to impact on boiler efficiency is more<br />

disperse, with a slope <strong>of</strong> -0.00113 (%/%) and R 2 = 0.900 (Figure 6.11). Physically, this<br />

dispersion means that the variation in humidity has different impact depending on other<br />

factors, mainly the ambient temperature. As it will be seen later, this dispersion appears<br />

in other variables that have low influence.<br />

Impact (%)<br />

Relative humidity and its impact on boiler efficiency<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-20 -15 -10 -5 0 5 10 15 20 25 30<br />

-0.02<br />

-0.04<br />

-0.06<br />

Relative humidity variation (%)<br />

Figure 6.11. Relative humidity variation and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

The last ambient condition considered is wind speed, which affects the losses to the<br />

ambient in the boiler. The average value <strong>of</strong> this variable is 4.55 m/s and its standard<br />

deviation is 2.82 m/s. In figure 6.12, the evolution <strong>of</strong> this variable is shown.<br />

145


Chapter 6<br />

Wind speed (m/s)<br />

146<br />

25<br />

20<br />

15<br />

10<br />

5<br />

-5<br />

Wind speed evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.12. Wind speed evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Impact <strong>of</strong> wind speed on boiler efficiency has a standard deviation <strong>of</strong> 0.115%, and<br />

the graph relating variable and impact has a slope <strong>of</strong> -0.0406 (%·s/m), with R 2 = 0.985.<br />

For the impact on unit heat rate, standard deviation is 0.00412, and the graph has a slope<br />

<strong>of</strong> 0.00145 s/m with R 2 = 0.984. Influence on cycle heat rate is practically zero, with an<br />

impact standard deviation <strong>of</strong> 8.04·10 -5 .<br />

6.3.2 Fuel quality<br />

Free <strong>diagnosis</strong> variables related to fuel quality are coal high heating value (HHV),<br />

coal composition (carbon, hydrogen, nitrogen, sulphur, moisture and ashes), and amount<br />

<strong>of</strong> natural gas introduced. Oxygen in coal has not been considered because one <strong>of</strong> all the<br />

coal components is not free in order to accomplish that the summation <strong>of</strong> all the mass<br />

fractions should be 100%. The amount <strong>of</strong> natural gas is expressed in terms <strong>of</strong><br />

percentage <strong>of</strong> power introduced by the fuel. Gas properties variation has not been<br />

considered.<br />

Figure 6.13 shows the evolution <strong>of</strong> coal HHV. This variable has an average <strong>of</strong><br />

16975 kJ/kg and a standard deviation <strong>of</strong> 639 kJ/kg, which comes from both long term<br />

variability and short-term dispersion. It should be noted that fuel properties are<br />

determined by daily <strong>analysis</strong> performed in the plant laboratory. Although these analyses


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

are performed according to standard procedures, the sampling method induces a non<br />

negligible error which causes this dispersion.<br />

Coal HHV (kJ/kg)<br />

%<br />

19000<br />

18000<br />

17000<br />

16000<br />

15000<br />

Coal HHV evolution<br />

14000<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

1.00<br />

0.50<br />

-0.50<br />

-1.00<br />

-1.50<br />

-2.00<br />

-2.50<br />

-3.00<br />

-3.50<br />

Figure 6.13. Coal high heating value evolution.<br />

Impact on boiler efficiency due to coal HHV<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.14. Impact on boiler efficiency due to coal HHV.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

147


Chapter 6<br />

Impact (%)<br />

148<br />

Coal HHV and its impact on boiler efficiency<br />

0<br />

-3500 -3000 -2500 -2000 -1500 -1000 -500 0 500 1000 1500<br />

-0.5<br />

HHV variation (kJ/kg)<br />

Figure 6.15. Coal HHV variation and its impact on boiler efficiency.<br />

Variations in coal HHV induce strong variations in boiler efficiency, which can be<br />

seen in figure 6.14. Standard deviation <strong>of</strong> this impact is 0.574 %, the curve relating<br />

variable and impact has a slope <strong>of</strong> 0.000889 (%·kg/kJ) and it is very linear (R 2 = 0.999),<br />

as it can be seen in figure 6.15. This influence on boiler efficiency is obviously traduced<br />

in impacts on unit heat rate. Standard deviation <strong>of</strong> these impacts is 0.0237, and relation<br />

between variable and impact is -3.64·10 -5 kg/kJ (R 2 = 0.998). What is more surprising is<br />

that this magnitude affects also the cycle: standard deviation <strong>of</strong> cycle heat rate impact is<br />

0.00262 and the slope <strong>of</strong> the graph which relates variable variation and impact is -<br />

4.02·10 -6 kg/kJ (R 2 = 0.996). This influence <strong>of</strong> fuel-related variables on cycle heat rate<br />

can be explained by considering that fuel properties affect strongly the amount <strong>of</strong> air<br />

and gases flowing through the boiler, which, in turn, affects both tempering flows and<br />

the amount <strong>of</strong> steam consumed at coil heaters. So that, although steam cycle works in<br />

the same way, flows entering and exiting this part <strong>of</strong> the plant are affected, and heat rate<br />

varies. This example demonstrated that an <strong>analysis</strong> based on the decomposition into<br />

blocks is not suitable if a precision <strong>of</strong> 5% or lower is required.<br />

1<br />

0.5<br />

-1<br />

-1.5<br />

-2<br />

-2.5<br />

-3<br />

-3.5<br />

-4<br />

Unit 1<br />

Unit 2<br />

Unit 3


%<br />

1.20<br />

1.00<br />

0.80<br />

0.60<br />

0.40<br />

0.20<br />

-0.20<br />

-0.40<br />

Impact on boiler efficiency due to carbon in coal<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.16. Impact on boiler efficiency due to carbon in coal.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

One <strong>of</strong> the coal components that more affects plant behaviour is carbon. Its average<br />

value is 42.5 % and its standard deviation is 1.61 (%). In figure 6.16, impact <strong>of</strong> this<br />

variable on boiler efficiency is shown, where it can be observed both long term and<br />

short time variation. Standard deviation <strong>of</strong> this impact is 0.190 (%), and relation<br />

between variable and impact is -0.116 %/%, with R 2 = 0.997 (Figure 6.17). This<br />

negative value indicates that when carbon increases boiler efficiency decreases. It<br />

occurs because fuel HHV has been considered as an independent variable, which<br />

implies that carbon varies without modifying HHV. In this situation, an increment in<br />

carbon originates an increment <strong>of</strong> air to burn it and, subsequently, more losses. A<br />

similar situation will appear in other carbon components like hydrogen. This result<br />

appears because HHV is not actually an independent variable, because it is linked to<br />

fuel composition. It is further analyzed in Section 6.5.<br />

This impact on boiler efficiency originates an impact on unit heat rate. The standard<br />

deviation <strong>of</strong> this impact is 0.00978 and the slope <strong>of</strong> the graph impact-variable is<br />

0.00596 (1/%) (R 2 = 0.997). The standard deviation <strong>of</strong> the impact on cycle heat rate is<br />

149


Chapter 6<br />

0.00224 which comes from <strong>of</strong> a relation between variable and impact <strong>of</strong> 0.00137 (1/%),<br />

(R 2 = 0.997).<br />

Impact (%)<br />

%<br />

150<br />

Carbon in coal and its impact on boiler efficiency<br />

0<br />

-10 -8 -6 -4 -2 0 2 4<br />

1.5<br />

1.0<br />

0.5<br />

-0.5<br />

-1.0<br />

-1.5<br />

Carbon in coal variation (%)<br />

Figure 6.17. Carbon in coal variation and its impact on boiler efficiency.<br />

Figure 6.18. Impact on boiler efficiency due to hydrogen in coal.<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-0.2<br />

-0.4<br />

Impact on boiler efficiency due to hydrogen in coal<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Another important fuel component is hydrogen. This variable has an average value<br />

<strong>of</strong> 2.74 % and a standard deviation <strong>of</strong> 0.173 (%). Impact on boiler efficiency has a<br />

standard deviation <strong>of</strong> 0.310 and its evolution is plotted on figure 6.18, where it can be<br />

appreciated a long term evolution parallel to those seen in HHV and carbon.<br />

Impact (%)<br />

Hydrogen in coal and its impact on boiler efficiency<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1<br />

-0.5<br />

-1<br />

-1.5<br />

-2<br />

Hydrogen in coal variation (%)<br />

Figure 6.19. Hydrogen in coal and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.19 shows the relation <strong>of</strong> hydrogen in coal variation and its impact on boiler<br />

efficiency. Its slope is -1.80 (%/%) and its R 2 is 0.999. This variable shows the same<br />

effect as carbon: it has a negative slope because HHV has been defined as an<br />

independent variable. Hydrogen in fuel has also an important impact on unit heat rate,<br />

with a standard deviation <strong>of</strong> 0.01189 and a sensitivity coefficient <strong>of</strong> 0.0688 (1/%), (R 2 =<br />

0.999). Finally, the impact on cycle heat rate has a standard deviation <strong>of</strong> 0.000864 due<br />

to a sensitivity coefficient <strong>of</strong> 0.00499 (1/%), (R 2 = 0.997).<br />

151


Chapter 6<br />

Coal moisture has an average value <strong>of</strong> 18.9 % and a standard deviation <strong>of</strong> 1.62 %.<br />

Its impact on boiler efficiency has a standard deviation <strong>of</strong> 0.325 % and its evolution is<br />

plotted on figure 6.20. Although there is important point dispersion, it can be<br />

appreciated how moisture has been higher during the first years. It can be seen also<br />

peaks corresponding to summer months, when, because <strong>of</strong> the high ambient<br />

temperature, coal becomes dryer.<br />

%<br />

152<br />

1.0<br />

0.5<br />

-0.5<br />

-1.0<br />

-1.5<br />

Impact on boiler efficiency due to coal humidity<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.20. Impact on boiler efficiency due to coal moisture.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.21 shows the relation between the variation <strong>of</strong> moisture and its impact on<br />

boiler efficiency. Its slope is -0.201 (%/%), and R 2 is 0.999. Impact on unit heat rate has<br />

a standard deviation <strong>of</strong> 0.0126 and an impact factor <strong>of</strong> 0.00776 (1/%) with a very linear<br />

behaviour (R 2 = 0.999). Finally, moisture has a small impact on cycle heat rate, with a<br />

standard deviation <strong>of</strong> 0.000918 and a relation between impact and variable variation <strong>of</strong><br />

0.000564 (1/%), with R 2 = 0.996.


Impact (%)<br />

%<br />

Coal humidity and its impact on boiler efficiency<br />

1<br />

0.5<br />

-0.5<br />

-1<br />

-1.5<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0<br />

-6 -4 -2 0 2 4 6 8<br />

0.6<br />

0.4<br />

0.2<br />

-0.2<br />

-0.4<br />

-0.6<br />

Coal humidity variation (%)<br />

Figure 6.21. Coal moisture variation and its impact on boiler efficiency.<br />

Impact on boiler efficiency due to ash in coal<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

d<br />

Figure 6.22. Impact on boiler efficiency due to ash in coal.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

153


Chapter 6<br />

Figure 6.22 shows the impact on boiler efficiency due to ashes variation, which has<br />

a standard deviation <strong>of</strong> 0.134. It can be seen how the ash content <strong>of</strong> coal was lower<br />

during the first two years, which implied higher efficiency. The average ash content in<br />

coal is 23.3 % and the standard deviation is 1.84 %. The relation <strong>of</strong> ash variation and its<br />

impact on boiler efficiency is -0.0726 (%/%) with R 2 = 0.996 (Figure 6.23). It may seem<br />

that this impact due to ashes is quite high, considering that this component is an inert<br />

that only affects losses related to heat <strong>of</strong> ashes and slag. However, it affects also to<br />

losses due to carbon in ashes (the independent variable is a percentage) and mainly it<br />

makes to vary the oxygen component in coal (which is not an independent variable)<br />

and, subsequently, the amount <strong>of</strong> air needed for combustion. This effect <strong>of</strong> variation <strong>of</strong><br />

oxygen in coal <strong>of</strong> course appears in all the composition variables, but it is stronger in<br />

those which vary more, such as ashes and moisture.<br />

Impact (%)<br />

154<br />

Ash in coal and its impact on boiler efficiency<br />

0<br />

-8 -6 -4 -2 0 2 4 6<br />

0.6<br />

0.4<br />

0.2<br />

-0.2<br />

-0.4<br />

-0.6<br />

Ash in coal variation (%)<br />

Figure 6.23. Ash in coal variation and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Impact on unit heat rate has a standard deviation <strong>of</strong> 0.00566 and a relation between<br />

impact and variable increment <strong>of</strong> 0.00307 (1/%) with R 2 = 0.996. In the case <strong>of</strong> the<br />

impact on cycle heat rate, standard deviation is 0.000723, with an impact factor <strong>of</strong><br />

0.000392 (1/%) and R 2 = 0.996.


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Another significant coal property is sulphur content, which has also a strong<br />

importance from the environmental point <strong>of</strong> view. However, the operation <strong>of</strong> the plant<br />

for removing SO2 from flue gases is not considered in this work, apart from the amount<br />

<strong>of</strong> preheated gases extracted. Average value <strong>of</strong> sulphur in coal is 4.84 % and standard<br />

deviation is 0.4118 %. The impact <strong>of</strong> sulphur on boiler efficiency is shown in figure<br />

6.24, and its standard deviation is 0.0217 %. During the first year, coal had low sulphur<br />

content, so that boiler efficiency was higher. However, this variable has an important<br />

dispersion.<br />

%<br />

0.06<br />

0.04<br />

0.02<br />

-0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

-0.10<br />

Boiler efficiency impact due to sulphur in coal<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.24. Impact on boiler efficiency due to sulphur in coal.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Relation <strong>of</strong> sulphur variation and its impact on boiler efficiency is -0.0519 (%/%)<br />

with R 2 = 0.996 (Figure 6.25). Impact <strong>of</strong> sulphur on unit heat rate has a standard<br />

deviation <strong>of</strong> 0.00113, and an impact factor <strong>of</strong> 0.00272 (1/%) with R 2 = 0.996. Finally,<br />

sulphur has a negligible impact on cycle heat rate, with a standard deviation <strong>of</strong><br />

0.000279 and a relation between impact and variable <strong>of</strong> 0.000672 (1/%), R 2 = 0.997.<br />

The last coal component is nitrogen, and has a negligible effect. For example, the<br />

standard deviation <strong>of</strong> its impact on boiler efficiency is 0.00211, which is ten times lower<br />

than that <strong>of</strong> sulphur.<br />

155


Chapter 6<br />

Impact (%)<br />

%<br />

156<br />

Sulphur in coal and its impact on boiler efficiency<br />

0<br />

-1 -0.5 0 0.5 1 1.5 2<br />

10<br />

8<br />

6<br />

4<br />

2<br />

-2<br />

0.06<br />

0.04<br />

0.02<br />

-0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

-0.1<br />

Sulphur in coal variation (%)<br />

Figure 6.25. Sulphur in coal variation and its impact on boiler efficiency.<br />

Natural gas fraction evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.26. Natural gas fraction evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

The last free <strong>diagnosis</strong> variable related to fuel is the fraction <strong>of</strong> energy due to natural<br />

gas. This variable has an average value <strong>of</strong> 0.576 % and a standard deviation <strong>of</strong> 1.10 %.<br />

This high variability appears because this variable has a discontinuous evolution<br />

varying from 0 to almost 10% (Figure 6.26). The impact on boiler efficiency due to gas<br />

contribution to energy in fuel is reduced with a standard deviation <strong>of</strong> 0.00683 % and an<br />

averaged impact factor <strong>of</strong> -0.00542 (%/%). However, the main feature <strong>of</strong> this<br />

dependence is its variability with R 2 = 0.766, which can be seen in figure 6.27. This<br />

means that the introduction <strong>of</strong> gas natural can have different effect depending on the<br />

quality <strong>of</strong> the coal. In most cases, the use <strong>of</strong> natural gas makes the boiler efficiency to<br />

decrease, mainly due to the increment <strong>of</strong> losses related to hydrogen in fuel. It should be<br />

noted that effects such as influence <strong>of</strong> natural gas in boiler heat transfer are not<br />

considered, because these aspects are included in other free <strong>diagnosis</strong> variables.<br />

Impact (%)<br />

0.04<br />

0.02<br />

-0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

-0.1<br />

Natural gas fraction and its impact on boiler efficiency<br />

0<br />

-2 0 2 4 6 8 10<br />

Natural gas fraction variation (%)<br />

Figure 6.27. Natural gas fraction variation and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Impact <strong>of</strong> natural gas fraction on unit heat rate has a standard deviation <strong>of</strong> 0.000163<br />

and a relation between impact and variable <strong>of</strong> 0.0000979 (1/%), with high dispersion<br />

(R 2 = 0.437). Impact on cycle heat rate has a standard deviation <strong>of</strong> 0.0000852 and an<br />

impact factor <strong>of</strong> -0.0000767 (1/%). Surprisingly, it has low dispersion (R 2 = 0.985),<br />

which can be explained by considering that the use <strong>of</strong> natural gas always reduces the<br />

157


Chapter 6<br />

amount <strong>of</strong> air needed for combustion and, subsequently, the steam needed by coil<br />

preheaters.<br />

6.3.3 Steam cycle set points.<br />

These set points are four: temperature and pressure <strong>of</strong> live steam entering the high<br />

pressure turbine, temperature <strong>of</strong> reheated steam entering the first medium pressure<br />

turbine and gross power produced. Possibility <strong>of</strong> regenerative pre-heaters by-pass could<br />

also be considered as operator-controlled degrees <strong>of</strong> freedom, but these situations are<br />

considered in the temperature difference <strong>of</strong> preheaters (indicators <strong>of</strong> equipment<br />

efficiency).<br />

Figure 6.28 shows the evolution <strong>of</strong> live steam temperature. It can be seen how this<br />

set point is fixed around 539 ºC, although it is sometimes modified for example during<br />

the first two years in unit 1. Besides, several points appear below these fixed points, due<br />

to the control based on tempering. Average temperature is 537.9 ºC and standard<br />

deviation is 1.28 ºC.<br />

ºC<br />

158<br />

544<br />

542<br />

540<br />

538<br />

536<br />

534<br />

532<br />

530<br />

Live steam temperature evolution<br />

528<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.28. Live steam temperature evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Variation <strong>of</strong> live steam temperature has an impact on cycle heat rate. Standard<br />

deviation <strong>of</strong> this impact is 0.00101 and the relation between variation and impact is -<br />

0.00787 1/ºC with very linear behaviour (R 2 = 0.999), as it can be seen in figure 6.29.<br />

This impact on cycle heat rate entails an impact on unit heat rate, with a standard<br />

deviation <strong>of</strong> 0.01090 and an impact factor <strong>of</strong> -0.000850 1/ºC (R 2 = 0.999). Impact on<br />

boiler efficiency is negligible, with a standard deviation <strong>of</strong> 0.000357%.<br />

Impact<br />

Live steam temperature and its impact on cycle heat rate<br />

Figure 6.29. Live steam temperature variation and its impact on cycle heat rate.<br />

Reheated steam temperature has an average value <strong>of</strong> 538.2 ºC and a standard<br />

deviation <strong>of</strong> 2.10 ºC, and its evolution is plotted in figure 6.30. It can be seen how this<br />

set point is almost constant in unit 3 but has been set in different levels in unit 1. This<br />

set-point variation entails an impact on steam cycle heat rate which has a standard<br />

deviation <strong>of</strong> 0.00129, and an impact factor <strong>of</strong> -0.000623 (1/ºC) with R 2 = 0.999, as it can<br />

be seen in figure 6.31. This steam cycle heat rate variation implies a variation <strong>of</strong> unit<br />

heat rate <strong>of</strong> a standard deviation <strong>of</strong> 0.00132 and a relation between impact and variable<br />

<strong>of</strong> -0.000625 (1/ºC) with R 2 = 0.999. Impact on boiler efficiency is again negligible,<br />

with a standard deviation <strong>of</strong> 0.000475 %.<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0<br />

-10 -8 -6 -4 -2 0 2 4 6<br />

-0.002<br />

-0.004<br />

Live steam temperature variation (ºC)<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

159


Chapter 6<br />

160<br />

ºC<br />

Impact<br />

550<br />

545<br />

540<br />

535<br />

530<br />

525<br />

Reheated steam temperature evolution<br />

520<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.30. Reheated steam temperature evolution.<br />

Reheated steam temperature and its impact on cycle heat rate<br />

0.015<br />

0<br />

-20 -15 -10 -5 0 5 10 15<br />

0.01<br />

0.005<br />

-0.005<br />

-0.01<br />

Reheated steam temperature variation (ºC)<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.31. Reheated steam temperature variation and its impact on cycle heat rate.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Another set-point related to cycle operation is live steam pressure, which has an<br />

average value <strong>of</strong> 157.5 bar and a standard deviation <strong>of</strong> 1.35 bar. Its evolution is shown<br />

in figure 3.32, where it can be appreciated how this parameter is usually high in unit 2<br />

and low in unit 3.


ar<br />

Impact<br />

164<br />

162<br />

160<br />

158<br />

156<br />

154<br />

152<br />

Live steam pressure evolution<br />

Figure 6.32. Live steam pressure evolution.<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

150<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Live steam pressure and its impact on cycle heat rate<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0<br />

-6 -4 -2 0 2 4 6<br />

-0.002<br />

-0.004<br />

-0.006<br />

-0.008<br />

Live steam pressure variation (bar)<br />

Figure 6.33. Live steam pressure and its impact on cycle heat rate.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Relation between steam pressure and its impact on cycle heat rate is -0.00123 1/bar<br />

and is very linear (R 2 = 0.998) as it can be seen in figure 6.33. Standard deviation <strong>of</strong> this<br />

161


Chapter 6<br />

impact is 0.00166. Impact on unit heat rate has a standard deviation <strong>of</strong> 0.00183 and an<br />

impact factor <strong>of</strong> -0.00136 1/bar (R 2 = 0.999). Live steam pressure has a negligible<br />

impact on boiler efficiency, with a standard deviation <strong>of</strong> 0.000502 %.<br />

The last free <strong>diagnosis</strong> variable considered in this paragraph is gross power<br />

produced by the alternator, although it obviously affects not only the cycle but also the<br />

boiler. For the operation points studied (full load, or above 320 MW), this variable has<br />

an average value <strong>of</strong> 340.007 MW and a standard deviation <strong>of</strong> 7.00 MW (Figure 6.34).<br />

This high deviation appears because the units are not operated in fixed values but with<br />

load regulation.<br />

MW<br />

162<br />

380<br />

370<br />

360<br />

350<br />

340<br />

330<br />

320<br />

Gross electric power evolution<br />

310<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.34. Gross electric power evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Relation <strong>of</strong> electric power variation and its impact on cycle heat rate is shown in<br />

figure 6.35. Its slope has a value <strong>of</strong> 0.000542 1/MW (R 2 = 0.992). It may be surprising<br />

this positive value, which indicates that when load increases heat rate also does.<br />

However, it should be noted that condenser pressure is not a free <strong>diagnosis</strong> variable, and<br />

when load increases this pressure should also increase because the same cooling system<br />

has more heat to reject. Impact on boiler efficiency has a standard deviation <strong>of</strong> 0.0184<br />

(%) and a relation between variable and impact <strong>of</strong> 0.00197 (%/MW). The main<br />

characteristic <strong>of</strong> this relation is its high dispersion, which can be appreciated in Figure


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

6.36 (R 2 = 0.666). This dispersion also appears in the case <strong>of</strong> unit heat rate (R 2 = 0.613),<br />

which has an impact factor <strong>of</strong> 0.0000724 1/MW and a standard deviation <strong>of</strong> 0.000733.<br />

This low influence is due to the compensation <strong>of</strong> the influence negative in cycle and<br />

positive in boiler that a load increment has.<br />

Impact<br />

Impact (%)<br />

Gross electric power and its impact on cycle heat rate<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-20 -10 0 10 20 30 40<br />

-0.005<br />

-0.01<br />

-0.015<br />

Figure 6.35. Gross electric power variation and its impact on cycle heat rate.<br />

Gross electric power and its impact on boiler efficiency<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-20 -10 0 10 20 30 40<br />

-0.05<br />

-0.1<br />

Gross electric power variation (MW)<br />

Gross electric power variation (MW)<br />

Figure 6.36. Gross electric power variation and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

163


Chapter 6<br />

6.3.4 Boiler set points.<br />

In this section, operation set-points associated with the boiler are studied. The main<br />

free <strong>diagnosis</strong> variables belonging to this group are: oxygen content in flue gases<br />

leaving the boiler, cold-side average temperature <strong>of</strong> air preheaters and steam for<br />

blowing. There are other variables related to the distribution <strong>of</strong> air and flue gases in the<br />

pre-heaters area: (i) relation <strong>of</strong> tempering and primary air, (ii) relation <strong>of</strong> primary air<br />

and coal mass flows, (iii) amount <strong>of</strong> air extracted for desulfuration, (iv) temperature<br />

difference <strong>of</strong> flue gases entering the preheaters and (v) flue gases distribution among<br />

primary and secondary air preheaters.<br />

Figure 6.37 shows the evolution <strong>of</strong> the set-point <strong>of</strong> the oxygen in combustion gases<br />

at the exit <strong>of</strong> the boiler (before air preheaters). It can be seen how this variable has been<br />

modified throughout time following diverse operation strategies. First, it was quite high,<br />

then it decreased and finally it reached a stable value for the last three years. Differences<br />

among units can also be observed; for example, oxygen in unit 2 is usually higher than<br />

in unit 3. Average value <strong>of</strong> this variable is 2.05 % and it has a standard deviation <strong>of</strong><br />

0.259 %.<br />

%<br />

164<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

Oxygen set-point evolution<br />

1<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.37. Oxygen set-point evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Each increment <strong>of</strong> 1 point in oxygen in gases implies a reduction <strong>of</strong> 0.265 points in<br />

boiler efficiency. This relation is very linear (R 2 = 0.997) as it can be seen in figure<br />

6.38. Standard deviation <strong>of</strong> the impacts on boiler efficiency is 0.0683%.<br />

Impact (%)<br />

Oxygen set-point and its impact on boiler efficiency<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

-0.35<br />

Oxygen set-point variation (%)<br />

Figure 6.38. Oxygen set point variation and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

This impact on boiler efficiency originates an impact on unit heat rate, which has a<br />

standard deviation <strong>of</strong> 0.00339. Relation between this impact and oxygen variation is<br />

0.0131 (1/%) with R 2 = 0.997. Impact on cycle efficiency is reduced, with a standard<br />

deviation <strong>of</strong> 0.000747 and an impact factor <strong>of</strong> 0.00289 (1/%), R 2 = 0.995.<br />

The aim <strong>of</strong> average cold-side temperature on primary and secondary air preheaters is<br />

to avoid low temperatures in these heat exchangers, than can lead to acid condensation<br />

and corrosion. Each one <strong>of</strong> these parameters is maintained by preheating the air entering<br />

each preheater, by using a coil heater that consumes steam from the drum. Due to this<br />

fact, this set point should be high enough to avoid corrosion but not too high in order to<br />

reduce the steam consumption. It should be noted that these set points are minimum<br />

values, so that, if outside temperature is high enough it may happen that the amount <strong>of</strong><br />

steam needed reaches the value <strong>of</strong> zero, and the average cold-side temperature increases<br />

above the set point. In these cases, to be strict, impact corresponding to this last part<br />

should be associated to ambient temperature.<br />

165


Chapter 6<br />

Average cold-side temperature on secondary air preheaters has a mean value <strong>of</strong><br />

109.3 ºC and a standard deviation <strong>of</strong> 6.67 ºC. Figure 6.39 shows how this variable has<br />

been modified along time. During some months <strong>of</strong> the first years, it has reached low<br />

values because part <strong>of</strong> the heat exchangers had been replaced by using corrosion<br />

resistant alloy. However, then it was decided to maintain higher values. Besides, some<br />

peaks in summer can be seen, corresponding to situations in which steam extracted for<br />

heating the air has become zero.<br />

ºC<br />

166<br />

140<br />

130<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

Average cold-side temperature in secondary air preheaters<br />

60<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.39. Evolution <strong>of</strong> average cold-side temperature in secondary air preheaters.<br />

Due to the consumption <strong>of</strong> steam, this variable affects directly to unit heat rate, with<br />

an impact standard deviation <strong>of</strong> 0.00683. As it can be seen in figure 6.40, each<br />

additional degree implies an increment <strong>of</strong> 0.00102 in unit heat rate (R 2 = 0.999). Impact<br />

on boiler efficiency has a standard deviation <strong>of</strong> 0.174%, with an impact factor <strong>of</strong> -<br />

0.0262 %/ºC and R 2 = 0.999. This important impact appears because when the average<br />

cold-side temperature increases, air entering the preheater is hotter and this heat<br />

exchanger works worse. Impact on cycle heat rate is negligible, with a standard<br />

deviation <strong>of</strong> 0.000276.


Impact<br />

ºC<br />

Average cold-side temperatue in secondary air preheater<br />

and its impact on unit heat rate<br />

-0.01<br />

-0.02<br />

-0.03<br />

-0.04<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0<br />

-40 -30 -20 -10 0 10 20 30<br />

0.03<br />

0.02<br />

0.01<br />

Average temperature variation (ºC)<br />

Figure 6.40. Average cold-side temperature in secondary air preheaters variation and its<br />

impact on unit heat rate.<br />

150<br />

140<br />

130<br />

120<br />

110<br />

100<br />

90<br />

Average cold-side temperature in primary air preheaters<br />

80<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.41. Average cold-side temperature in primary air preheaters evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

167


Chapter 6<br />

Figure 6.41 shows the evolution <strong>of</strong> the average cold-side temperature in primary air<br />

preheaters. It can be seen how this variable was very high during the first months and<br />

then decreased. It decreased again at the end <strong>of</strong> 2000. The effect commented for<br />

secondary air preheaters and consisting <strong>of</strong> quite high values <strong>of</strong> this average temperature<br />

appearing when high ambient temperature or operation conditions make the steam<br />

required to be equal to zero, appears now in more points. Mean value <strong>of</strong> this variable is<br />

111.6 ºC and its standard deviation is 7.25 ºC.<br />

Impact <strong>of</strong> this variable on unit heat rate has a standard deviation <strong>of</strong> 0.00208, and the<br />

impact factor is 0.000281 1/ºC (R 2 = 0.990). As it can be seen in figure 6.42, relation <strong>of</strong><br />

variable increment and its impact is not homogeneous but it present two tendencies: a<br />

main one and another with fewer points. This secondary tendency corresponds to the<br />

first months when the old preheaters were working. As it will be seen later, these heat<br />

exchangers presented important air infiltration; this fact, in turn, originated an additional<br />

amount <strong>of</strong> air to be heated and subsequently more steam consumption and more impact.<br />

Impact<br />

168<br />

Average cold-side temperature in primary air preheaters<br />

and its impact on unit heat rate<br />

0.01<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0<br />

-30 -20 -10 0 10 20 30 40<br />

-0.002<br />

-0.004<br />

-0.006<br />

-0.008<br />

Average temperature variation (ºC)<br />

Figure 6.42. Average cold-side temperature in primary air preheaters variation and its<br />

impact on unit heat rate.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Impact on boiler efficiency has a standard deviation <strong>of</strong> 0.0534 % and an impact<br />

factor <strong>of</strong> -0.00724 %/ºC (R 2 = 0.994). In the graph relating variable and this impact, a


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

secondary tendency similar to that on figure 6.42 also appears. It should be noted that<br />

impact factors corresponding to average temperature at primary air preheaters are<br />

around three times and a half smaller than those <strong>of</strong> secondary preheaters, which<br />

approximately correspond to the relation <strong>of</strong> air mass flows. Finally, impact on cycle<br />

heat rate is negligible, with a standard deviation <strong>of</strong> 0.0000877.<br />

Blowing steam has an average value <strong>of</strong> 1.40 kg/s and a standard deviation <strong>of</strong> 0.564<br />

kg/s. Its evolution is plotted in figure 4.63, where it can be seen a tendency to increase<br />

throughout time and a quite high dispersion. It should be noted that figures considered<br />

are values averaged along every day from a continuous variable that usually is equal to<br />

zero and in some moments reaches values higher than those shown here.<br />

kg/s<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

-1<br />

Blowing steam evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.43. Sootblowing steam evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Steam consumption for sootblowing originates an important impact on unit heat rate,<br />

whose standard deviation is 0.00455. Relation <strong>of</strong> steam flow variation and its impact is<br />

0.00804 s/kg with R 2 = 0.999 (Figure 6.44). Impacts on cycle heat rate and boiler<br />

efficiency are lower, with standard deviations <strong>of</strong> 0.000941 and 0.00123%, respectively.<br />

It should be noted that, in the case <strong>of</strong> boiler efficiency, effects on flow pattern and heat<br />

transfer while blowing are not considered, which provides an impact lower than the real<br />

169


Chapter 6<br />

one. However, the main effect <strong>of</strong> blowing, which is steam consumption, has been<br />

considered.<br />

Impact<br />

170<br />

0.04<br />

0.035<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

-0.01<br />

-0.015<br />

Blowing steam and its impact on unit heat rate<br />

0<br />

-2 -1 0 1 2 3 4 5<br />

-0.005<br />

Blowing steam variation (kg/s)<br />

Figure 6.44. Sootblowing steam variation and its impact on unit heat rate.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

The main set point related to air and flue gases distribution in the preheaters area is<br />

the amount <strong>of</strong> hot air that is extracted from the boiler to heat the air leaving the flue gas<br />

desulfuration unit before it enters the stack. This variable has an average value <strong>of</strong> 4.86<br />

kg/s and a standard deviation <strong>of</strong> 8.79 kg/s. This high dispersion appears because this<br />

variable has usually low values that sometimes increase dramatically, as it can be seen<br />

in Figure 6.45.<br />

Extraction <strong>of</strong> this hot air originates an important impact on boiler efficiency, with a<br />

standard deviation <strong>of</strong> 0.140 % and an impact factor <strong>of</strong> -0.0160 %·s/kg with R 2 = 0.998<br />

(Figure 6.46). This boiler efficiency variation originates an impact on unit heat rate with<br />

a standard deviation 0.00540 and a relation between impact and variable <strong>of</strong> 0.000613<br />

s/kg (R 2 = 0.996). Standard deviation <strong>of</strong> impact on cycle heat rate is negligible<br />

(0.000214).


kg/s<br />

Impact (%)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

-10<br />

Air for SO 2 removing unit evolution<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

0.2<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1<br />

-1.2<br />

Figure 6.45. Air for flue gas desulfuration unit evolution.<br />

Air for SO 2 removing unit and its impact on boiler efficiency<br />

0<br />

-10 0 10 20 30 40 50 60 70<br />

Air mass flow variation (kg/s)<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.46. Air for flue gas desulfuration unit variation and its impact on boiler efficiency.<br />

171


Chapter 6<br />

Figure 6.47 shows the evolution <strong>of</strong> the relation <strong>of</strong> tempering and primary air (%).<br />

Actually, the most suitable <strong>diagnosis</strong> variable should have been the set point <strong>of</strong><br />

temperature <strong>of</strong> primary air leaving the mills; however, this information was not<br />

available, so that, this mass flow relation was chosen. In the graph it can be seen two<br />

different zones. The second one presents more dispersion and corresponds to the actual<br />

measured flows, while in the first one (the older periods) information about flows was<br />

not available and a correlation with ambient temperature has been used. Evolution <strong>of</strong><br />

this variable shows a cyclic tendency: in summer, the amount <strong>of</strong> tempering air should be<br />

increased in order to keep the temperature set point. Average value <strong>of</strong> this variable is<br />

33.8% and its standard deviation is 7.06 %.<br />

%<br />

172<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Tempering and primary air relation<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.47. Relation <strong>of</strong> tempering and primary air evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

This variable originates an impact on boiler efficiency with a standard deviation <strong>of</strong><br />

0.103 %. Relation between variable increment and its impact is -0.0148 (%/%) with R 2<br />

= 0.991 (Figure 6.48). The negative sign appears because when more primary air is bypassed,<br />

primary air preheaters work worse. Impact on unit heat rate has a standard<br />

deviation <strong>of</strong> 0.00345 and an impact factor <strong>of</strong> 0.000497 (1/%), with R 2 = 0.991. Effect on<br />

cycle heat rate is negligible.


Impact (%)<br />

Impact (%)<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Relation <strong>of</strong> tempering and primary air and its impact on boiler efficiency<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-25 -20 -15 -10 -5 0 5 10 15 20 25<br />

-0.1<br />

-0.2<br />

-0.3<br />

Flows relation variation (%)<br />

Figure 6.48. Variation <strong>of</strong> the relation <strong>of</strong> tempering and primary air and its impact on boiler<br />

efficiency.<br />

Primary air-coal ratio and its impact on boiler efficiency<br />

0.1<br />

0.05<br />

0<br />

-0.4 -0.2 0 0.2 0.4 0.6 0.8 1<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

-0.35<br />

-0.4<br />

Primary air-coal ratio variation (kg/kg)<br />

Figure 6.49. Primary air-coal ratio and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

The variable that controls air distribution among primary and secondary preheaters<br />

is the primary air to coal ratio (kg/kg). This variable has an average value <strong>of</strong> 1.97 kg/kg<br />

173


Chapter 6<br />

and a standard deviation <strong>of</strong> 0.143 kg/kg. Relation between impact on boiler efficiency<br />

and variable increment is on average -0.347 % but it has a little dispersion, with R 2 =<br />

0.970 (Figure 6.49). When the amount <strong>of</strong> primary air increases, also the amount <strong>of</strong><br />

tempering air does (relation <strong>of</strong> tempering and primary air is a free <strong>diagnosis</strong> variable),<br />

so that efficiency decreases; however, dispersion appears due to the crossed influence <strong>of</strong><br />

preheaters efficiency variation. Impact on boiler efficiency has a standard deviation <strong>of</strong><br />

0.0530 % and is plotted in figure 6.50. During the first three years, point dispersion is<br />

lower because information on flow measurement was not available, and relation <strong>of</strong><br />

primary and secondary air was fixed. In the last three years, it can be seen how this<br />

impact affects more negatively to unit 1 than unit 3.<br />

Impact on unit heat rate has a standard deviation <strong>of</strong> 0.00176 and an impact factor <strong>of</strong><br />

0.0114 kg/kg (R 2 = 0.964), while effect on cycle heat rate is negligible.<br />

%<br />

174<br />

0.10<br />

0.05<br />

-0.10<br />

-0.15<br />

-0.20<br />

-0.25<br />

-0.30<br />

-0.35<br />

-0.40<br />

Impact on boiler efficiency due to primary air-coal ratio<br />

0.00<br />

24/07/1998 0:00<br />

-0.05<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.50 Impact on boiler efficiency due to primary air-coal ratio.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Flue gases distribution among primary and secondary air preheaters is controlled by<br />

the percentage <strong>of</strong> flue gases entering primary air preheaters. This variable has an<br />

average value <strong>of</strong> 24.5 % and a standard deviation <strong>of</strong> 3.22 %. It has a very low influence<br />

on both boiler and unit, with impact standard deviations <strong>of</strong> 0.0157 % and 0.000646.<br />

Average values <strong>of</strong> impact factors are -0.00102 %/% and 0.0000523 1/%, but these


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

relations are characterized with high dispersion (R 2 = 0.080 and 0.123). Effect on cycle<br />

heat rate is negligible.<br />

The last variable in this section is temperature difference between flue gases<br />

entering primary air preheaters and flue gases entering secondary air preheaters. This<br />

difference is due to the effect <strong>of</strong> the by-pass <strong>of</strong> the first economizer. Average value <strong>of</strong><br />

this variable is 7.50 ºC and it has a standard deviation <strong>of</strong> 14.8 ºC. Impact <strong>of</strong> this variable<br />

is negligible, with a standard deviation <strong>of</strong> 0.00855 % in the boiler and 0.000526 in the<br />

unit. However, its evolution is shown in figure 6.51 because it shows a difference in<br />

operation strategy that induces a difference in aggregated boiler effectiveness, as it will<br />

be seen later. In this graph, it can be seen how during the first two years temperature<br />

difference was higher in winter, probably due to the opening <strong>of</strong> the by-pass in order to<br />

reduce the amount <strong>of</strong> steam consumed by the coil heaters.<br />

ºC<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

-20<br />

-30<br />

Temperature difference <strong>of</strong> flue gases entering preheaters<br />

0<br />

24/07/1998 0:00<br />

-10<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.51. Evolution <strong>of</strong> temperature difference <strong>of</strong> flue gases entering air preheaters.<br />

6.3.5 Steam cycle component parameters.<br />

Free <strong>diagnosis</strong> variables corresponding to this group are turbine isoentropic<br />

efficiencies, turbine flow coefficients (except in the high pressure one) and temperature<br />

differences in regenerative water heaters. Pressure drop in reheater and pressure<br />

increment provided by turbo-pump are also included because, although its physical<br />

175


Chapter 6<br />

causes are mainly located in the boiler, they affect above all the steam cycle. Finally,<br />

water losses are also considered.<br />

Figure 6.52 shows the evolution <strong>of</strong> high pressure steam turbine isoentropic<br />

efficiency. It can be seen how this variable presents an oscillating tendency: it has<br />

higher efficiency in summer than in winter. This effect appears because <strong>of</strong> the variations<br />

<strong>of</strong> pressure ratio and the effect <strong>of</strong> the regulation system: since the load is almost<br />

constant, in winter, with lower condenser pressure, governing valves should be more<br />

closed. Units 1 and 3 present quite constant values, but the former has higher efficiency<br />

than the latter. Efficiency <strong>of</strong> unit 2 tends to decrease but it is recovered again after being<br />

repaired in april 2004. Considered all units together, this variable has an average value<br />

<strong>of</strong> 79.5% with a standard deviation <strong>of</strong> 1.81 %.<br />

%<br />

176<br />

85<br />

83<br />

81<br />

79<br />

77<br />

75<br />

High pressure steam turbine isoentropic efficiency evolution<br />

73<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.52. Evolution <strong>of</strong> high pressure steam turbine isoentropic efficiency.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

High pressure steam turbine variation implies an important impact on cycle heat<br />

rate. Standard deviation <strong>of</strong> this impact is 0.00905 and the relation between variable<br />

increment and impact is -0.00499 1/%, with a very linear behaviour (R 2 = 0.999), as it<br />

can be seen in figure 6.53. This impact on cycle heat rate causes an effect on unit heat<br />

rate, with an impact standard deviation <strong>of</strong> 0.00993 and an impact factor <strong>of</strong> -0.00548 1/%<br />

(R 2 = 0.999). Impact on boiler efficiency is very reduced; it has a standard deviation <strong>of</strong><br />

0.00357 %.


Impact<br />

%<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Figure 6.53. High pressure steam turbine isoentropic efficiency variation and its impact on<br />

cycle heat rate.<br />

90<br />

88<br />

86<br />

84<br />

82<br />

80<br />

78<br />

76<br />

74<br />

High pressure steam turbine isoentropic efficiency<br />

and its impact on cycle heat rate<br />

0<br />

-8 -6 -4 -2 0 2 4<br />

Efficiency variation (%)<br />

Medium pressure 1 steam turbine isoentropic efficiency evolution<br />

72<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.54. Medium pressure 1 steam turbine isoentropic efficiency evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Evolution <strong>of</strong> isoentropic efficiency <strong>of</strong> first medium pressure steam turbine is plotted<br />

in figure 6.54. It can be seen how units 1 and 3 remain quite stable for a long period and<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

-0.03<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

177


Chapter 6<br />

increase their efficiency suddenly in may 2004 and may 2003 respectively, probably<br />

due to the maintenance works performed in these components. Efficiency <strong>of</strong> turbine in<br />

unit 2 does not present clear efficiency variations along the period considered, although<br />

it presents some dispersion during the second and third year. It should be noted that a<br />

quite important dispersion appears in all units during 1999. Average value <strong>of</strong> this<br />

variable is 80.9 % and it has a standard deviation <strong>of</strong> 2.10 %.<br />

Relation between this efficiency and its impact on cycle heat rate is -0.00244 1/%,<br />

and originates an impact standard deviation <strong>of</strong> 0.00513 (both values approximately half<br />

<strong>of</strong> those <strong>of</strong> the high pressure turbine). This relation is also quite linear (R 2 = 0.998), as it<br />

can be seen in figure 6.55. Impact on unit heat rate has a standard deviation <strong>of</strong> 0.00579<br />

and an impact factor <strong>of</strong> -0.00276 1/% (R 2 = 0.998). Standard deviation <strong>of</strong> impact on<br />

boiler efficiency is only 0.00222 %.<br />

Impact<br />

178<br />

Medium pressure 1 steam turbine isoentropic efficiency<br />

and its impact on cycle heat rate<br />

0<br />

-12 -10 -8 -6 -4 -2 0 2 4 6 8<br />

-0.005<br />

-0.015<br />

Figure 6.55. Medium pressure 1 steam turbine isoentropic efficiency variation and its impact<br />

on cycle heat rate.<br />

Isoentropic efficiency <strong>of</strong> the second medium pressure steam turbine has an average<br />

value <strong>of</strong> 88.2 % and a standard deviation <strong>of</strong> 1.26 %. Evolution <strong>of</strong> this variable is plotted<br />

in figure 6.56, where it can be appreciated a steadily degradation <strong>of</strong> the turbine on unit<br />

2. It can also be seen a strong degradation in unit 3 in may 2003, so that part <strong>of</strong> the<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

-0.01<br />

-0.02<br />

Efficiency variation (%)<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

improvement observed in this month in the first medium pressure turbine may be<br />

induced by a deviation in instrumentation.<br />

%<br />

Impact<br />

92<br />

91<br />

90<br />

89<br />

88<br />

87<br />

86<br />

85<br />

Medium pressure 2 steam turbine isoentropic efficiency evolution<br />

84<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.56. Medium pressure 2 steam turbine isoentropic efficiency evolution.<br />

Medium pressure 2 steam turbine isoentropic efficiency<br />

and its impact on cycle heat rate<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-4 -2 0 2 4 6 8<br />

-0.005<br />

-0.01<br />

-0.015<br />

-0.02<br />

-0.025<br />

-0.03<br />

Efficiency variation(%)<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.57. Medium pressure 2 steam turbine isoentropic efficiency variation and its impact<br />

on cycle heat rate.<br />

179


Chapter 6<br />

Impact on cycle heat rate due to variation <strong>of</strong> this variable has a standard deviation <strong>of</strong><br />

0.0383 (lower that the medium pressure 1 turbine) and an impact factor <strong>of</strong> -0.00303 1/%<br />

(a little bit more than the other medium pressure turbine). Relation between variable and<br />

impact is again linear (R 2 = 0.998), as figure 6.57 shows. Impact on unit heat rate has a<br />

standard deviation <strong>of</strong> 0.00433 and an impact factor <strong>of</strong> -0.00343 1/% (R 2 = 0.999). Effect<br />

on boiler efficiency is again negligible.<br />

The last isoentropic efficiency corresponds to the low pressure turbine, and has the<br />

important characteristic that it is not directly calculated from measurements, but it is has<br />

the value suitable to obtain the cycle heat rate calculated, for a given value <strong>of</strong> the other<br />

boundary conditions. For this reasons, this value is affected not by measurement errors<br />

<strong>of</strong> a few variables but by errors <strong>of</strong> all the other variables affecting the cycle and errors<br />

induced by all hypothesis made (efficiency <strong>of</strong> the turbo-pump, behaviour <strong>of</strong> the low<br />

pressure preheaters…). As a result, this variable has a big dispersion: it has a standard<br />

deviation <strong>of</strong> 3.60 % and an average value <strong>of</strong> 80.9%.<br />

Impact<br />

180<br />

0.15<br />

0.10<br />

0.05<br />

-0.05<br />

-0.10<br />

-0.15<br />

Impact on cycle heat rate due to low pressure steam turbine<br />

isoentropic efficiency<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.58. Impact on cycle heat rate due to low pressure steam turbine isoentropic<br />

efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Since, as explained above, the value <strong>of</strong> this variable includes effects not only due to<br />

the low pressure turbine itself but also to the hypothesis made in its calculation, it is


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

perhaps better to analyze its impact on cycle heat rate than the variable. This impact has<br />

a standard deviation <strong>of</strong> 0.0435 and its evolution is plotted on figure 6.58. In this figure it<br />

can be seen how impacts in units 1 and 2 increase steadily and then decreases suddenly<br />

in May 2004 and in April 2003 respectively. Impact on unit 3 has a quite high<br />

dispersion and it is the lowest <strong>of</strong> the three units.<br />

Relation <strong>of</strong> variable increment and its impact on cycle heat rate is -0.0120 1/% with<br />

R 2 = 0.999. Impact on unit heat rate has a standard deviation <strong>of</strong> 0.0493 and an impact<br />

factor <strong>of</strong> -0.0136 1/% with R 2 = 0.999 (Figure 6.59). Finally, effect on boiler efficiency<br />

is very low, with an impact standard deviation <strong>of</strong> 0.0158 %.<br />

Impact<br />

Low pressure steam turbine isoentropic efficiency and its impact<br />

on unit heat rate<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-15 -10 -5 0 5 10 15<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

Efficiency variation (%)<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.59. Low pressure steam turbine isoentropic efficiency variation and its impact on<br />

unit heat rate.<br />

Apart from the isoentropic efficiency, the other parameter used to characterize steam<br />

turbines behaviour is flow coefficient, except in the cases <strong>of</strong> the high pressure one,<br />

which works in controlled inlet pressure. They determine the value <strong>of</strong> steam pressure in<br />

turbine inlets, which affects not only the distribution <strong>of</strong> enthalpy drops between turbines<br />

but the properties <strong>of</strong> steam extracted to water heaters. However, it should be noted that<br />

influence <strong>of</strong> flow coefficients on cycle heat rate is about one order <strong>of</strong> magnitude lower<br />

than effects <strong>of</strong> isoentropic efficiency.<br />

181


Chapter 6<br />

Impact<br />

Impact<br />

182<br />

0.002<br />

0.001<br />

-0.001<br />

-0.002<br />

-0.003<br />

-0.004<br />

Impact on unit heat rate due to flow coefficient in medium<br />

pressure steam turbine 1<br />

0.000<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.60. Impact on unit heat rate due to flow coefficient <strong>of</strong> medium pressure steam<br />

turbine 1.<br />

Flow coefficient <strong>of</strong> medium pressure steam turbine 1 and its<br />

impact on cycle heat rate<br />

0<br />

-1.5 -1 -0.5 0 0.5<br />

-0.001<br />

-0.002<br />

-0.003<br />

-0.004<br />

Flow coefficient variation (kg 1/2 m 3/2 s -1 bar -1/2 )<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.61. Flow coefficient <strong>of</strong> medium pressure steam turbine 1 variation and its impact<br />

on cycle heat rate.<br />

0.002<br />

0.001<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Flow coefficient <strong>of</strong> the first medium pressure steam turbine has an average value <strong>of</strong><br />

13.1 kg 1/2 ·m 3/2 ·s -1 ·bar -1/2 and a standard deviation <strong>of</strong> 0.234 kg 1/2 ·m 3/2 ·s -1 ·bar -1/2 . Its impact<br />

on cycle heat rate has a standard deviation <strong>of</strong> 0.000743 and its evolution is plotted in<br />

figure 6.60. It can be seen how this parameter remains quite constant; the only<br />

remarkable variation is an increment in June 2003 in unit 3.<br />

Relation between this variable and its impact on cycle heat rate is shown in figure<br />

6.61. Average slope <strong>of</strong> the curve is 0.00331 kg -1/2 ·m -3/2 ·s 1 ·bar 1/2 and it presents a little<br />

dispersion (R 2 = 0.978). Impact on unit heat rate has a standard deviation <strong>of</strong> 0.00123<br />

and an impact factor <strong>of</strong> 0.00546 kg -1/2 ·m -3/2 ·s 1 ·bar 1/2 (R 2 = 0.990).<br />

Flow coefficients <strong>of</strong> the second medium pressure and low pressure steam turbines<br />

have lower influence on cycle heat rate, with standard deviations <strong>of</strong> 0.000394 and<br />

0.000370 respectively. The first has an impact factor <strong>of</strong> 0.000796 although dispersion is<br />

quite high (R 2 = 0.898). The value <strong>of</strong> the impact factor <strong>of</strong> the second is not significant,<br />

because it has a very high dispersion (R 2 = 0.00403).<br />

ºC<br />

15<br />

10<br />

5<br />

-5<br />

-10<br />

Hot side temperature difference <strong>of</strong> 6 th water preheater evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.62. Hot side temperature difference (TTD) <strong>of</strong> 6 th preheater evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

183


Chapter 6<br />

Impact<br />

184<br />

Hot side temperature difference <strong>of</strong> 6 th water preheater and its<br />

impact on cycle heat rate<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-10 -5 0 5 10 15<br />

-0.005<br />

-0.01<br />

Temperature difference variation (ºC)<br />

Figure 6.63. TTD <strong>of</strong> 6 th preheater variation and its impact on cycle heat rate.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

After analysing variables related to turbines, parameters characterizing regenerative<br />

water heaters are going to be studied. These parameters are temperature differences in<br />

hot and cold sides: terminal temperature difference (TTD) and drain cooling approach<br />

temperature (TDCA). The first ones have an important influence on cycle and unit heat<br />

rate, while the others have low impact.<br />

Figure 6.62 shows the evolution <strong>of</strong> TTD <strong>of</strong> 6 th water preheater. This difference is<br />

calculated by substracting the saturation temperature at the pressure <strong>of</strong> steam exiting the<br />

turbine minus temperature <strong>of</strong> hot water leaving the preheater. Since steam is<br />

superheated, this difference can reach also negative values. The fact that pressure is<br />

considered at the exit <strong>of</strong> turbine is important because if a pressure drop occurs in the<br />

duct or in the preheater entry, it is diagnosed as an increment in temperature difference<br />

in the preheater. This phenomenon occurs during the first year in unit 3 and during other<br />

periods in unit 2, when temperature difference can reach values <strong>of</strong> 10 and 5 ºC<br />

respectively. A small oscillating tendency with temperatures slightly higher in summer<br />

than in winter can also be appreciated. Average value <strong>of</strong> this variable is -0.76 ºC and it<br />

has a standard deviation <strong>of</strong> 2.27 ºC.


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Relation <strong>of</strong> temperature difference and its impact on cycle heat rate is 0.000710<br />

1/ºC, with R 2 = 0.998 (Figure 6.63). This impact has a standard deviation <strong>of</strong> 0.00162.<br />

Influence on cycle heat rate originates an impact on unit heat rate, characterized by an<br />

impact factor <strong>of</strong> 0.00103 1/ºC (R 2 = 0.999) and a standard deviation <strong>of</strong> 0.00236.<br />

Evolution <strong>of</strong> TTD <strong>of</strong> 5 th water preheater is plotted in figure 6.64. It can be<br />

appreciated how this variable reaches values near 10 ºC during the first three years in<br />

unit 2, probably due to pressure drop in the steam duct. There are other periods in this<br />

unit in which temperature difference reaches almost around 40 ºC. In these situations,<br />

water leaves the preheater at the same temperature at which it enters, which means that<br />

the heat exchanger is by-passed. This situation demonstrated how the <strong>diagnosis</strong> method<br />

is able to deal with discrete changes by using continuous variables chosen in a suitable<br />

way. Seasonal oscillation is also present but is difficult to see due to the high range <strong>of</strong><br />

the y axis. Average value <strong>of</strong> this variable is 1.78 ºC and it has a standard deviation <strong>of</strong><br />

6.70 ºC.<br />

ºC<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

-10<br />

Hot side temperature difference <strong>of</strong> 5 th water preheater evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.64. Hot side temperature difference <strong>of</strong> 5 th water preheater evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Relation between variable increment and its impact on cycle heat rate is 0.000397<br />

1/ºC, with R 2 = 0.996, as it can be seen in figure 6.64. Standard deviation <strong>of</strong> this impact<br />

is 0.00265. Effect on unit heat rate is characterized by an impact factor <strong>of</strong> 0.000455 1/ºC<br />

(R 2 = 0.9962) and a standard deviation <strong>of</strong> 0.00304.<br />

185


Chapter 6<br />

Impact<br />

ºC<br />

186<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

-0.005<br />

Hot side temperature difference <strong>of</strong> 5 th water preheater and its impact<br />

on cycle heat rate<br />

0<br />

-10 0 10 20 30 40 50 60<br />

Temperature difference variation (%)<br />

Figure 6.65. Hot side temperature difference (TTD) <strong>of</strong> 5 th water preheater variation and its<br />

impact on cycle heat rate.<br />

10<br />

8<br />

6<br />

4<br />

2<br />

-4<br />

-6<br />

-8<br />

-10<br />

Hot side temperature difference <strong>of</strong> 3 rd water preheater evolution<br />

0<br />

24/07/1998 0:00<br />

-2<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.66. Hot side temperature difference (TTD) <strong>of</strong> 3 rd water preheater evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

The last hot side temperature considered is that corresponding to third preheater,<br />

because <strong>of</strong> the lack <strong>of</strong> instrumentation <strong>of</strong> low pressure area. Its evolution is plotted in<br />

figure 6.66. In this graph, it can be seen the cyclic evolution commented in the other<br />

preheaters, as well as some differences among units (variable value is bigger in unit 3).<br />

However, no important differences due to by-passes or pressure drops in steam ducts<br />

are observed. Average value <strong>of</strong> this temperature difference is 0.11 ºC and it has a<br />

standard deviation <strong>of</strong> 2.02 ºC. Due to this low variability, this variable has low influence<br />

on cycle heat rate. Its impact has a standard deviation <strong>of</strong> 0.00223 and an impact factor<br />

<strong>of</strong> 0.00111 1/ºC, with very linear behaviour (R 2 = 0.999). This impact on cycle<br />

originates an impact on unit heat rate, with an impact factor <strong>of</strong> 0.00125 1/ºC (R 2 =<br />

0.999) and a standard deviation <strong>of</strong> 0.00252.<br />

The other parameters used to characterize water preheaters behaviour are<br />

temperature differences at cold side (drain cooling approach temperature, TDCA).<br />

These differences are calculated by substracting drain water temperature minus entering<br />

water temperature. However, influence <strong>of</strong> these parameters is lower.<br />

Cold side temperature difference <strong>of</strong> water <strong>of</strong> preheater 6 has an average value <strong>of</strong><br />

10.9 ºC and a standard deviation <strong>of</strong> 5.40 ºC. Its impacts on cycle and unit heat rates<br />

have standard deviations <strong>of</strong> 0.000256 and 0.000293 respectively, which are about eight<br />

times smaller than those <strong>of</strong> hot side temperature difference in the same preheaters.<br />

Impact factors are 0.0000470 1/ºC (R 2 = 0.985) and 0.0000537 1/ºC (R 2 = 0.985).<br />

In the case <strong>of</strong> 5 th preheater, cold side temperature difference has an average value <strong>of</strong><br />

8.06 ºC and a standard deviation <strong>of</strong> 17.5 ºC. This variability appears mainly because <strong>of</strong><br />

difference among units: values in unit 2 are around 25 or 30 ºC, in unit 3 around 10<br />

degrees and in unit 1 in the middle. Impact factors are around twice as high as those <strong>of</strong><br />

6 th preheater: 0.000106 1/ºC (R 2 = 0.998) and 0.000120 1/ºC (R 2 = 0.998). The<br />

combination <strong>of</strong> high variable deviation and impact factors is responsible for impact<br />

standard deviations in cycle and unit heat rates more than three times higher than in 6 th<br />

preheater (0.000850 and 0.000961).<br />

Pressure loses in the water-steam circuit are summarized in two variables: pressure<br />

drop in reheater and pressure increment provided by the turbo-pump. Although these<br />

parameters vary also due to degradation <strong>of</strong> components located in the boiler, they affect<br />

mainly to the cycle heat rate, so they are presented here.<br />

Figure 6.67 shows the evolution <strong>of</strong> pressure drop in reheater. It can be seen how unit<br />

1 presents an oscillating tendency, and in unit 3 a jump is produced in June 2003. In unit<br />

187


Chapter 6<br />

2, some points indicate a very low pressure drop during the first months <strong>of</strong> 2004;<br />

however, in this period 6 th water preheater was by-passed, so that pressure indication<br />

was not very reliable. Average value <strong>of</strong> this variable is 1.52 bar and its standard<br />

deviation is 0.669 bar. It should be noted that, for convenience, this variable is defined<br />

as the pressure difference <strong>of</strong> steam entering the 6 th preheater and reheated steam<br />

entering medium pressure turbine. So that, it comprises the pressure drop in the reheater<br />

and ducts minus the pressure drop in the duct connecting high pressure turbine<br />

extraction with 6 th water preheater. This fact justifies the low values <strong>of</strong> this variable.<br />

bar<br />

188<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

-1<br />

Pressure drop in reheater evolution<br />

0<br />

24/07/1998 0:00<br />

-0.5<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.67 Pressure drop in reheater evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

This pressure drop variation originates an important impact on cycle heat rate.<br />

Relation <strong>of</strong> variable increment and its impact has an average value <strong>of</strong> 0.00524 1/bar (R 2<br />

= 0.999) and is shown in figure 6.68, while impact factor for unit heat rate is 0.00535<br />

1/bar (R 2 = 0.999). Standard deviations are 0.00350 and 0.00358, respectively. Impact<br />

on boiler efficiency has a standard deviation <strong>of</strong> only 0.000972 %.<br />

Pressure increment provided by turbo-pump is plotted in figure 6.69. It can be seen<br />

how units 2 and 3 have quite constant values, while unit 1 presents two peaks during the<br />

first months <strong>of</strong> 2000 and 2004. Average value <strong>of</strong> this variable is 177.8 bar and it has a<br />

standard deviation <strong>of</strong> 2.76 bar.


Impact<br />

bar<br />

Pressure drop in preheater and its impact on cycle heat rate<br />

0.015<br />

0.01<br />

0.005<br />

-0.005<br />

-0.01<br />

-0.015<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0<br />

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5<br />

190<br />

185<br />

180<br />

175<br />

170<br />

165<br />

Pressure drop variation (bar)<br />

Figure 6.68. Pressure drop in preheater variation and its impact on cycle heat rate.<br />

Pressure increment in turbo-pump evolution<br />

160<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.69. Pressure increment in turbo-pump evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Influence <strong>of</strong> this variable on cycle heat rate is characterized by an impact factor <strong>of</strong><br />

0.000333 1/bar (R 2 = 0.990) and by its slightly curved shape, as it can be seen in figure<br />

189


Chapter 6<br />

6.70. The impact has a standard deviation <strong>of</strong> 0.00104. This variation <strong>of</strong> cycle heat rate<br />

originates an impact on unit heat rate with an impact factor <strong>of</strong> 0.000376 1/bar (R 2 =<br />

0.991) and a standard deviation <strong>of</strong> 0.00117.<br />

Impact<br />

190<br />

Pressure increment in turbo-pump and its effect on cycle heat rate<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0<br />

-10 -5 0 5 10 15 20<br />

-0.002<br />

-0.004<br />

Pressure increment variation (bar)<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.70. Pressure increment in turbo-pump variation and its impact on cycle heat rate.<br />

The last variable to be considered in this section is uncontrolled water losses. These<br />

losses are distributed along cycle and boiler, and conventionally are supposed to be<br />

located in the drum. However, they mainly affect the unit heat rate and also the cycle<br />

heat rate, so that they are presented here. Evolution <strong>of</strong> this variable is plotted in Figure<br />

6.71. It has an average value <strong>of</strong> 2.52 kg/s and a standard deviation <strong>of</strong> 1.69 kg/s. This<br />

variable is calculated by mass balance <strong>of</strong> the water-steam circuit, and it is affected by<br />

some error sources, mainly tank levels. So that it has a lot <strong>of</strong> dispersion and some<br />

outliers with high values appear. Besides, assumption on the location <strong>of</strong> these loses also<br />

introduces an important uncertainty in the calculation <strong>of</strong> its impact.<br />

Water losses have an important effect on unit heat rate, because they entail a loss <strong>of</strong><br />

heated water, apart from the cost <strong>of</strong> demineralised water that is not considered in this<br />

work. Relation between variable increment and its impact on unit heat rate is 0.00339<br />

s/kg (R 2 = 0.999) and it is plotted on figure 6.72. Standard deviation <strong>of</strong> this impact is<br />

0.00576. Impact on cycle heat rate is more reduced, with a standard deviation <strong>of</strong>


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0.00282 and an impact factor <strong>of</strong> -0.00167 s/kg (R 2 = 0.999). It may seem strange the<br />

negative value <strong>of</strong> this factor, but it should be noted that water loses are located at the<br />

drum, so that, when they increases, the amount <strong>of</strong> water entering the boiler (which has a<br />

negative sign in the calculation <strong>of</strong> the heat given to the cycle) also does.<br />

kg/s<br />

Impact<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

-2<br />

Water loses evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.71. Water losses evolution.<br />

Water loses and its impact on unit heat rate<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

-6 -4 -2 0 2 4 6 8 10 12<br />

-0.01<br />

-0.02<br />

Water loses variation (kg/s)<br />

Figure 6.72. Water losses variation and its impact on unit heat rate.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

191


Chapter 6<br />

6.3.6 Cooling system component parameters<br />

In this section, influence <strong>of</strong> condenser effectiveness, cooling water flow rate and<br />

cooling tower effectiveness is analyzed. These parameters characterize the behaviour <strong>of</strong><br />

the cooling system, whose degradation implies higher condenser pressure and higher<br />

cycle heat rate for the same ambient conditions. In their determination, temperatures <strong>of</strong><br />

cooling water entering and exiting the condenser are used. These values are very<br />

difficult to measure, because there are only two thermocouples in each section, so that<br />

big uncertainty appears in these three free <strong>diagnosis</strong> variables.<br />

192<br />

1<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

Condenser effectiveness evolution<br />

0.5<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.73. Condenser effectiveness evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Evolution <strong>of</strong> condenser effectiveness is plotted in figure 6.73. Although high<br />

dispersion is present, it can be seen how this effectiveness is higher in unit 1. This<br />

variable has an average value <strong>of</strong> 0.689 and a standard deviation <strong>of</strong> 0.0622. Relation<br />

between condenser effectiveness and its impact on cycle heat rate is shown in figure<br />

6.74; this curved graph demonstrates the capability <strong>of</strong> the method to deal with nonlinear<br />

behaviour. Impact factor is -0.177 (R 2 = 0.983) and impact standard deviation is<br />

0.0110. Relation between variable increment and its impact on unit heat rate presents


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

the same curved shape. Its average slope is -0.200 (R 2 = 0.982), and the impact has a<br />

standard deviation <strong>of</strong> 0.0125. Effect on boiler efficiency is negligible.<br />

Impact<br />

kg/s<br />

Condenser effectiveness and its impact on cycle heat rate<br />

0<br />

-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3<br />

13000<br />

12000<br />

11000<br />

10000<br />

9000<br />

8000<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

-0.03<br />

-0.04<br />

Effectiveness variation<br />

Figure 6.74. Condenser effectiveness variation and its impact on cycle heat rate.<br />

Cooling water mass flow evolution<br />

7000<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.75. Cooling water flow rate evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

193


Chapter 6<br />

Cooling water flow rate has an average value <strong>of</strong> 10699 kg/s and a standard deviation<br />

<strong>of</strong> 553 kg/s. Figure 6.75 shows how evolution <strong>of</strong> this variable is characterized by a big<br />

dispersion, above all in unit 3. Although not very clearly, it can be seen how unit 1 had<br />

a sudden flow reduction in April 2003 and how flows in units 2 and 3 have tended to<br />

increase along time. These variations in flow rate cause an important impact on cycle<br />

heat rate, characterized by an impact factor <strong>of</strong> -0.0000190 s/kg with R 2 = 0.997 (Figure<br />

6.76) and a standard deviation <strong>of</strong> 0.0105. In the case <strong>of</strong> unit heat rate, relation between<br />

variable increment and impact is -0.0000216 s/kg (R 2 = 0.997), which originates an<br />

impact standard deviation <strong>of</strong> 0.0119.<br />

Impact<br />

194<br />

Cooling water mass flow and its impact on cycle heat rate<br />

0<br />

-3000 -2500 -2000 -1500 -1000 -500 0 500 1000 1500<br />

Mass flow variation (kg/s)<br />

Figure 6.76. Cooling water flow rate variation and its impact on cycle heat rate.<br />

Finally, figure 6.77 shows the evolution <strong>of</strong> cooling tower effectiveness. The main<br />

characteristic <strong>of</strong> this graph is its cyclic evolution, which clearly reveals that this variable<br />

is not actually independent but it is strongly linked to ambient conditions, mainly<br />

ambient temperature. This problem is tackled in a following section. Besides this<br />

oscillating behaviour, it can be seen how effectiveness tends to decrease along time,<br />

which indicates a progressive degradation. Average value <strong>of</strong> this variable is 0.4580 and<br />

it has a standard deviation <strong>of</strong> 0.05860.<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

-0.03<br />

Unit 1<br />

Unit 2<br />

Unit 3


Cooling tower effectiveness evolution<br />

Figure 6.77. Cooling tower effectiveness evolution.<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

0.3<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Impact<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

0.4<br />

0.35<br />

Cooling tower effectiveness and its impact on cycle heat rate<br />

0<br />

-0.25 -0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.78. Cooling tower effectiveness variation and its impact on cycle heat rate.<br />

This variable has a strong influence on both cycle and unit heat rates, with impact<br />

factors <strong>of</strong> -0.329 (R 2 = 0.998) and -0.372 (R 2 = 0.998) respectively (Figure 6.78). Impact<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.02<br />

-0.04<br />

Tower effectiveness variation<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

195


Chapter 6<br />

standard deviations are 0.0195 and 0.0221. Impact on boiler efficiency has a standard<br />

deviation <strong>of</strong> only 0.0101 %.<br />

6.3.7 Boiler component parameters<br />

In this last section are included indicators related to heat transfer, such as preheaters<br />

effectiveness and aggregated boiler effectiveness, air infiltration in preheaters and<br />

residual carbon in ashes. This last variable is an indicator <strong>of</strong> the quality <strong>of</strong> the<br />

combustion, and depends mainly on coal quality, oxygen in flue gases set point and<br />

burners operation. However, since this dependence is difficult to be modelled precisely,<br />

it is considered as an uncontrollable variable.<br />

Evolution <strong>of</strong> secondary air preheater effectiveness is shown in figure 6.79. Point<br />

dispersion appears because effectiveness is not a pure independent variable but it<br />

depends on flow rates and temperatures <strong>of</strong> flows entering the heat exchanger. However,<br />

thanks to the use <strong>of</strong> a lot <strong>of</strong> points, tendencies can be appreciated. It can be seen how<br />

units 1 and 2 have an important improvement in May 2001, while unit 3 remains more<br />

constant. This variable has an average value <strong>of</strong> 0.838 and a standard deviation <strong>of</strong><br />

0.0324.<br />

Preheater effectiveness<br />

196<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

Secondary air preheater effectiveness evolution<br />

0.7<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.79. Secondary air preheater effectiveness evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Impact on boiler efficiency due to secondary air preheater effectiveness has a<br />

standard deviation <strong>of</strong> 0.249 % and relation between variable increment and impact is<br />

7.71 %, with R 2 = 0.997 (Figure 6.80). This variation <strong>of</strong> boiler efficiency entails an<br />

important effect on unit heat rate, characterized by an impact factor <strong>of</strong> -0.262 (R 2 =<br />

0.997) and a standard deviation <strong>of</strong> 0.00844.<br />

Impact (%)<br />

Secondary air preheater effectiveness and its impact on boiler efficiency<br />

0<br />

-0.15 -0.1 -0.05 0 0.05 0.1<br />

-0.2<br />

Figure 6.80. Secondary air preheater effectiveness variation and its impact on boiler<br />

efficiency.<br />

Primary air preheater effectiveness has an average value <strong>of</strong> 0.873 and a standard<br />

deviation <strong>of</strong> 0.0225. In figure 6.81, it can be seen how effectiveness has decreased<br />

steadily or remained constant except two improvements in December 1999 and May<br />

2001. Effectiveness <strong>of</strong> unit 1 is lower from the last months <strong>of</strong> 2002 to the repairing in<br />

May 2004.<br />

Impact <strong>of</strong> this variable is lower than that <strong>of</strong> secondary air preheater effectiveness,<br />

because the amount <strong>of</strong> primary air is also lower. Effect on boiler efficiency is<br />

characterized by an impact factor <strong>of</strong> 2.16 % with R 2 = 0.991 (Figure 6.82) and a<br />

standard deviation <strong>of</strong> 0.0490 %. Impact on unit heat rate has a standard deviation <strong>of</strong><br />

0.00167 and relation between variable and impact is -0.0733 (R 2 = 0.989). It should be<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1<br />

-1.2<br />

Effectiveness variation<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

197


Chapter 6<br />

noted that these correlation coefficients are not as high as in other variables due to<br />

crossed effects caused by air distribution variation.<br />

Preheater effectiveness<br />

Impact (%)<br />

198<br />

1<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

Primary air preheater effectiveness evolution<br />

0.75<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.81. Primary air preheater effectiveness evolution.<br />

Primary air preheater effectiveness and its impact on boiler efficiency<br />

0<br />

-0.15 -0.1 -0.05 0 0.05 0.1<br />

-0.05<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

Effectiveness variation<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.82. Primary air preheater effectiveness variation and its impact on boiler efficiency.


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Heat transfer effectiveness <strong>of</strong> all heat exchangers in the boiler different from air<br />

preheaters is summarized in only one variable called aggregated boiler effectiveness.<br />

Evolution <strong>of</strong> this variable is plotted in figure 6.83, where two parts can be appreciated.<br />

During the first years, an oscillating tendency appears, with higher values in summer<br />

than in winter. Afterwards, this behaviour disappears and effectiveness remains quite<br />

constant. This behaviour can be justified by considering the evolution <strong>of</strong> the<br />

temperature difference between gases entering primary and secondary air preheaters<br />

(Figure 6.51), which follows the same pattern. During the first years, when the gases<br />

by-pass was opened in winter, not only temperature difference but also average gas<br />

temperature increased, so that aggregated boiler effectiveness decreased. The effect in<br />

summer was the opposite. Aggregated boiler effectiveness has an average value <strong>of</strong><br />

0.861 and a standard deviation <strong>of</strong> 0.0182.<br />

Boiler effectiveness<br />

0.95<br />

0.93<br />

0.91<br />

0.89<br />

0.87<br />

0.85<br />

0.83<br />

0.81<br />

0.79<br />

0.77<br />

Aggregated boiler effectiveness evolution<br />

0.75<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.83. Aggregated boiler effectiveness evolution.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

This variable affects strongly the boiler efficiency and hence the unit heat rate.<br />

Impact standard deviations are 0.424 % and 0.0114 respectively. Relation between<br />

aggregated boiler effectiveness and its impact on boiler efficiency is 23.4 %, with R 2 =<br />

0.998 (Figure 6.84). In the case <strong>of</strong> unit heat rate, this impact factor is -0.627 (R 2 =<br />

0.998).<br />

199


Chapter 6<br />

Impact<br />

Air infiltration (%)<br />

200<br />

Aggregated boiler effectiveness and its impact on boiler efficiency<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.1 -0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08 0.1<br />

-0.5<br />

-1<br />

-1.5<br />

-2<br />

-2.5<br />

Effectiveness variation<br />

Figure 6.84. Aggregated boiler effectiveness variation and its impact on boiler efficiency.<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Air infiltration in preheaters evolution<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.85. Air infiltration in preheaters evolution.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.85 shows the evolution <strong>of</strong> air infiltration in preheaters, expressed in mass<br />

percentage <strong>of</strong> flue gases. It can be seen how during the first months, infiltration was<br />

very high. Then, preheaters were replaced, so that infiltration decreased dramatically.


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Then, infiltration tends to increase along time, except in some points when preheaters<br />

are repaired: May 2002 for units 2 and 3, May 2003 for unit 3 and May 2004 for units 1<br />

and 2. Average value <strong>of</strong> infiltration is 13.0 % and it has a standard deviation <strong>of</strong> 3.56 %.<br />

Each point in air infiltration entails a reduction <strong>of</strong> 0.0278 points in boiler efficiency,<br />

with R 2 = 0.998 (Figure 6.86), and an increment <strong>of</strong> 0.00107 in unit heat rate (R 2 =<br />

0.996). Impact standard deviations are 0.0997 % and 0.00387, respectively. Effect on<br />

cycle heat rate is negligible.<br />

Impact (%)<br />

Air infiltration in preheaters and its impact on boiler efficiency<br />

0.4<br />

0.2<br />

0<br />

-10 -5 0 5 10 15 20 25 30 35<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1<br />

Air infiltration variation (%)<br />

Figure 6.86. Air infiltration in preheaters and its impact on boiler efficiency.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

The last variable considered is carbon percentage in ashes. It should be noted that<br />

this variable is not actually independent, because it depends on fuel and boiler<br />

operation. This dependence may be included by using a suitable combustion model<br />

(which is a complex task which falls out <strong>of</strong> the scope <strong>of</strong> this work) and <strong>causal</strong>ity chains<br />

theory. Carbon in ashes has an average value <strong>of</strong> 0.955 % and a standard deviation <strong>of</strong><br />

0.556 %, and is plotted in figure 6.87. This variable has quite dispersion, but it can be<br />

seen how it was low during the first two years and then increased.<br />

Relation between carbon percentage and its impact on boiler efficiency is -0.452<br />

%/%, with R 2 = 0.997 (Figure 6.88). This impact has a standard deviation <strong>of</strong> 0.252 %.<br />

Effect on unit heat rate is characterized by an impact factor <strong>of</strong> 0.0162 1/% (R 2 = 0.997)<br />

201


Chapter 6<br />

and a standard deviation <strong>of</strong> 0.0901. Impact on cycle heat rate has a standard deviation <strong>of</strong><br />

only 0.000178.<br />

Carbon %<br />

Impact (%)<br />

202<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Carbon in ashes evolution<br />

0<br />

24/07/1998 0:00<br />

-0.5<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.87. Carbon in ashes evolution.<br />

Carbon in ashes and its impact on boiler efficiency<br />

1<br />

0.5<br />

0<br />

-1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4<br />

-0.5<br />

-1<br />

-1.5<br />

-2<br />

Carbon in ashes variation (%)<br />

Figure 6.88. Carbon in ashes variation and its impact on boiler efficiency.<br />

Reference<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

6.3.8 Summary <strong>of</strong> results.<br />

In the previous sections, a lot <strong>of</strong> results corresponding to almost all <strong>diagnosis</strong><br />

variables have been presented. Here, all <strong>of</strong> them are collected in order to summarize<br />

information and to facilitate comparison among them. To achieve this, indicators<br />

defined and applied previously are shown in Table 6.1: average and standard deviation<br />

<strong>of</strong> the free <strong>diagnosis</strong> variables, and impact factors (with their R 2 values) and standard<br />

deviation <strong>of</strong> their impacts on the three global efficiency indicators. These standard<br />

deviations quantify the importance <strong>of</strong> each free <strong>diagnosis</strong> variables. Units indicated in<br />

the third column correspond to the variable average and standard deviation. Units for<br />

impacts are % in boiler efficiency and dimensionless for cycle and unit heat rates. R 2<br />

values are dimensionless. Finally, units for impact factors are the inverse <strong>of</strong> those<br />

indicated in the unit column for impacts on cycle and unit heat rate, and these inverse<br />

values multiplied times % for impacts on boiler efficiency.<br />

In order to clearly distinguish the most important variables from those which have<br />

negligible impact, they have been classified according to the standard deviation <strong>of</strong> their<br />

impacts on boiler efficiency, cycle and unit heat rates. Variables whose impact standard<br />

deviation is above 20% <strong>of</strong> the maximum are labelled pink, those from 10 to 20% are<br />

labelled orange, those from 5 to 10% are yellow and variables from 1 to 5% are bluegreen.<br />

The other variables remain white.<br />

As it can be seen, most important variables regarding boiler efficiency are ambient<br />

temperature, wind speed, most variables related to coal, average cold side temperature<br />

in secondary air preheaters, air extracted to flue gas desulfuration, aggregated boiler<br />

effectiveness, preheaters effectiveness and carbon in ashes. On the other hand, influence<br />

<strong>of</strong> variables related to cycle is negligible.<br />

Cycle heat rate varies mainly due to ambient temperature and humidity, cooling<br />

system condition (condenser, tower and cooling water flow rate), and also turbines<br />

isoentropic efficiencies.<br />

Finally, unit heat rate is influenced by variables related to both the boiler and the<br />

steam cycle. Due to this fact, the number <strong>of</strong> variables whose influence is lower than 1%<br />

<strong>of</strong> the maximum is only 5, compared to 22 and 20 in boiler efficiency and cycle heat<br />

rate respectively.<br />

203


Chapter 6<br />

Impact on cycle heat rate Impact on unit heat rate<br />

Variable Impact on boiler<br />

Num. Description Units<br />

effficiency<br />

Stand.<br />

R 2<br />

Impact<br />

dev.<br />

factor<br />

R 2 Stand.<br />

dev.<br />

0.9991 4.484<br />

6.137<br />

0.9998 5.276<br />

Mean Stand. Impact R<br />

dev. factor<br />

2 Stand. Impact<br />

dev. factor<br />

1 Ambient temperature ºC 15.5 7.30 0.05524 0.9958 0.4017 7.234<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

0.9920 9.314<br />

1.273<br />

0.9923 7.844<br />

0.8996 0.01081 1.073<br />

2 Relative humidity % 63.4 7.08 -1.132<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

0.9838 4.117<br />

1.455<br />

0.9827 8.039<br />

0.9854 0.1150 2.837<br />

3 Wind speed m/s 4.55 2.82 -4.065<br />

·10 -3<br />

·10 -3<br />

·10 -5<br />

·10 -5<br />

·10 -2<br />

0.9977 2.367<br />

-3.644<br />

0.9960 2.618<br />

0.9989 0.5744 -4.020<br />

4 Coal high heating value kJ/kg 16975 639 8.889<br />

·10 -2<br />

·10 -5<br />

·10 -3<br />

·10 -6<br />

·10 -4<br />

0.9969 9.782<br />

5.959<br />

0.9971 2.238<br />

5 Carbon mass fraction in coal % 42.53 1.61 -0.1164 0.9969 0.1898 1.365<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

0.9988 1.189<br />

6.877<br />

0.9967 8.645<br />

6 Hydrogen mass fraction in coal % 2.74 0.173 -1.796 0.9993 0.3102 4.989<br />

·10 -2<br />

·10 -2<br />

·10 -4<br />

·10 -3<br />

0.9988 1.257<br />

7.757<br />

0.9964 9.181<br />

7 Moisture mass fraction in coal % 18.87 1.62 -0.2007 0.9994 0.3249 5.641<br />

·10 -2<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

0.9956 5.657<br />

3.066<br />

0.9963 7.228<br />

0.9958 0.1339 3.921<br />

8 Ash mass fraction in coal % 23.16 1.84 -7.263<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -2<br />

0.9962 1.134<br />

2.717<br />

0.9972 2.791<br />

6.721<br />

0.9962 2.169<br />

9 Sulphur mass fraction in coal % 4.84 0.412 -5.191<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -2<br />

·10 -2<br />

0.9969 1.040<br />

1.755<br />

0.9969 2.221<br />

3.735<br />

0.9973 2.109<br />

-3.558<br />

10 Nitrogen mass fraction in coal % 0.66 5.94<br />

·10 -4<br />

·10 -3<br />

·10 -5<br />

·10 -4<br />

·10 -3<br />

·10 -2<br />

·10 -2<br />

Table 6.1.A. Summary <strong>of</strong> <strong>diagnosis</strong> results.<br />

204


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Impact on cycle heat rate Impact on unit heat rate<br />

Variable Impact on boiler<br />

Num. Description Units<br />

effficiency<br />

Stand.<br />

R 2<br />

Impact<br />

R 2 Stand.<br />

dev.<br />

Impact<br />

R 2 Stand.<br />

dev.<br />

Impact<br />

dev.<br />

factor<br />

factor<br />

factor<br />

Mean Stand.<br />

dev.<br />

0.4373 1.633<br />

9.792<br />

0.9854 8.518<br />

-7.669<br />

0.7648 6.829<br />

11 Energy provided by natural gas % 0.576 1.10 -5.415<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -5<br />

·10 -3<br />

·10 -3<br />

0.9991 1.090<br />

-8.504<br />

0.9991 1.009<br />

-7.871<br />

0.5825 3.566<br />

12 Live steam temperature ºC 537.9 1.28 -2.038<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

0.9997 1.310<br />

-6.239<br />

0.9997 1.308<br />

-6.239<br />

0.6481 4.750<br />

13 Reheated steam temperature ºC 538.2 2.10 -1.163<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

0.9987 1.830<br />

-1.351<br />

0.9984 1.664<br />

-1.228<br />

0.5918 5.025<br />

14 Live steam pressure bar 157.5 1.35 -2.832<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

0.6125 7.331<br />

7.245<br />

0.9923 3.804<br />

5.415<br />

0.6664 1.843<br />

15 Gross electric power MW 340.01 7.003 1.971<br />

·10 -4<br />

·10 -5<br />

·10 -3<br />

·10 -4<br />

·10 -2<br />

·10 -3<br />

0.9967 3.388<br />

1.310<br />

0.9951 7.474<br />

2.887<br />

% 2.05 0.259 -0.2645 0.9967 6.832<br />

16 Oxygen in flue gases leaving the<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -3<br />

·10 -2<br />

boiler<br />

0.9992 6.830<br />

1.025<br />

0.9854 2.762<br />

0.9989 0.1741 4.140<br />

109.3 6.67 -2.617<br />

ºC<br />

17 Average cold-side temperature<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

in secondary air preheaters<br />

0.9897 2.082<br />

2.809<br />

0.9811 8.771<br />

1.174<br />

0.9942 5.3412<br />

111.6 7.24 -7.243<br />

ºC<br />

18 Average cold-side temperature<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -2<br />

·10 -3<br />

in primary air preheaters<br />

0.9992 4.550<br />

8.041<br />

0.9986 9.412<br />

-1.662<br />

0.5924 1.232<br />

19 Sootblowing steam flow rate kg/s 1.40 0.564 1.579<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

0.9962 5.396<br />

6.134<br />

0.9919 2.141<br />

0.9977 0.1403 2.425<br />

4.86 8.79 -1.595<br />

kg/s<br />

20 Air for flue gas desulfuration<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

unit<br />

Table 6.1.B. Summary <strong>of</strong> <strong>diagnosis</strong> results.<br />

205


Chapter 6<br />

Impact on cycle heat rate Impact on unit heat rate<br />

Variable Impact on boiler<br />

Num. Description Units<br />

effficiency<br />

Stand.<br />

R 2<br />

Impact<br />

dev.<br />

factor<br />

R 2 Stand.<br />

dev.<br />

0.9912 3.455<br />

4.964<br />

0.5181 2.314<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

0.9637 1.764<br />

1.142<br />

0.6548 2.216<br />

Mean Stand. Impact R<br />

dev. factor<br />

2 Stand. Impact<br />

dev. factor<br />

21 Tempering and primary air kg/kg 33.8 7.06 -1.477<br />

relation<br />

·10 -2<br />

0.9909 0.1028 2.276<br />

·10 -6<br />

22 Primary air-coal ratio kg/kg 1.97 0.143 -0.3467 0.9704 5.299 9.733<br />

·10 -3<br />

·10 -2<br />

·10 -5<br />

·10 -5<br />

·10 -2<br />

0.6739 5.261<br />

2.796<br />

0.9869 2.365<br />

1.594<br />

8.549<br />

8.093<br />

7.50 14.8 -1.460<br />

ºC<br />

23 Temperature difference <strong>of</strong> flue<br />

·10 -4<br />

·10 -5<br />

·10 -4<br />

·10 -5<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

gases entering preheaters<br />

0.1231 6.461<br />

5.233·1<br />

0.4478 4.786<br />

7.759<br />

1.568<br />

8.001<br />

24.5 3.22 -1.021<br />

-<br />

24 Fraction <strong>of</strong> flue gases through<br />

·10 -4<br />

0 -5<br />

·10 -5<br />

·10 -6<br />

·10 -2<br />

·10 -2<br />

·10 -3<br />

PAH<br />

0.9994 9.927<br />

-5.478<br />

0.9994 9.051<br />

-4.994<br />

0.7024 3.567<br />

79.5 1.81 -1.434<br />

%<br />

25 High pressure turbine<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

isoentropic efficiency<br />

0.9985 5.793<br />

-2.761<br />

0.9982 5.127<br />

-2.443<br />

0.6685 2.219<br />

80.9 2.10 -7.000<br />

%<br />

26 Intermediate pressure 1<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

isoentropic efficiency<br />

0.9986 4.327<br />

-3.428<br />

0.9982 3.830<br />

-3.032<br />

0.6378 1.719<br />

88.2 1.26 -8.515<br />

%<br />

27 Intermediate pressure 2<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

isoentropic efficiency<br />

0.9990 4.925<br />

-1.359<br />

0.9989 4.351<br />

-1.199<br />

0.5902 1.577<br />

80.9 3.60 -3.130<br />

%<br />

28 Low pressure isoentropic<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -3<br />

efficiency<br />

0.9898 1.233<br />

5.462<br />

0.9783 7.434<br />

3.314<br />

0.6561 5.467<br />

13.1 0.233 1.301<br />

kg 1/2 ·m 3/2<br />

29 Intermediate pressure 1 flow<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·s -1 ·bar -1/2<br />

coefficient<br />

Table 6.1.C. Summary <strong>of</strong> <strong>diagnosis</strong> results.<br />

206


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

Impact on cycle heat rate Impact on unit heat rate<br />

Variable Impact on boiler<br />

Num. Description Units<br />

effficiency<br />

Stand.<br />

R 2<br />

Impact<br />

R 2 Stand.<br />

dev.<br />

Impact<br />

dev.<br />

factor<br />

factor<br />

R 2 Stand.<br />

dev.<br />

0.9016 4.509<br />

9.126<br />

0.8985 3.943<br />

7.956<br />

0.5958 1.462<br />

Mean Stand. Impact<br />

dev. factor<br />

24.7 0.439 2.321<br />

kg 1/2 ·m 3/2<br />

30 Intermediate pressure 2 flow<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·s -1 ·bar -1/2<br />

coefficient<br />

4.230<br />

1.418<br />

2.323<br />

3.695<br />

4.032<br />

-1.091<br />

1.278<br />

9.476<br />

51.4 0.964 5.734<br />

31 Low pressure flow coefficient kg 1/2 ·m 3/2<br />

·10 -4<br />

·10 -2<br />

·10 -5<br />

·10 -4<br />

·10 -3<br />

·10 -5<br />

·10 -4<br />

·10 -3<br />

·10 -6<br />

·s -1 ·bar -1/2<br />

0.9985 2.356<br />

1.034<br />

0.9983 1.616<br />

7.093<br />

0.6635 8.663<br />

32 TTD 6 th water heater ºC -0.76 2.27 2.933<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

0.9962 3.040<br />

4.548<br />

0.9959 2.652<br />

3.966<br />

0.7828 1.057<br />

33 TTD 5 th water heater ºC 1.78 6.70 1.371<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

0.9996 2.523<br />

1.253<br />

·10 -3<br />

0.9997 2.226<br />

1.105<br />

0.6505 7.914<br />

34 TTD 3 rd water heater ºC 0.11 2.02 3.128<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

0.9854 2.932<br />

5.373<br />

0.9847 2.564<br />

4.697<br />

0.6576 8.936<br />

35 TDCA 6 th water heater ºC 10.9 5.44 1.326<br />

·10 -4<br />

·10 -5<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -5<br />

0.9978 9.612<br />

1.202<br />

0.9977 8.501<br />

1.063<br />

0.6466 3.026<br />

36 TDCA 5 th water heater ºC 17.5 8.06 2.819<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

0.9990 3.579<br />

5.352<br />

0.9989 3.503<br />

5.238<br />

0.6757 9.723<br />

37 Pressure drop in reheater bar 1.52 0.669 1.193<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

0.9907 1.165<br />

3.756<br />

0.9898 1.039<br />

3.333<br />

0.6467 5.117<br />

177.8 2.76 8.762<br />

bar<br />

38 Pressure increment in turbo-<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

pump<br />

Table 6.1.D. Summary <strong>of</strong> <strong>diagnosis</strong> results.<br />

207


Chapter 6<br />

Impact on cycle heat rate Impact on unit heat rate<br />

Variable Impact on boiler<br />

Num. Description Units<br />

effficiency<br />

Stand.<br />

R 2<br />

Impact<br />

R 2 Stand.<br />

dev.<br />

Impact<br />

R 2 Stand.<br />

dev.<br />

Impact<br />

dev.<br />

factor<br />

factor<br />

factor<br />

Mean Stand.<br />

dev.<br />

0.9989 5.756<br />

3.391<br />

0.9987 2.823<br />

-1.671<br />

0.6215 1.870<br />

39 Water losses kg/s 2.52 1.69 7.813<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

-0.2002 0.9825 1.246<br />

-0.1771 0.9826 1.102<br />

0.6620 3.505<br />

-4.579<br />

40 Condenser effectiveness - 0.689 6.22<br />

·10 -2<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

·10 -2<br />

0.9967 1.190<br />

-2.155<br />

0.9970 1.050<br />

-1.903<br />

0.64620 3.910<br />

41 Cooling water flow rate kg/s 10699 553 -5.757<br />

·10 -2<br />

·10 -5<br />

·10 -2<br />

·10 -5<br />

·10 -3<br />

·10 -6<br />

-0.3724 0.9982 2.206<br />

-0.3286 0.9983 1.946<br />

-0.1202 0.7997 1.009·<br />

42 Cooling tower effectiveness - 0.458 5.86<br />

·10 -2<br />

·10 -2<br />

10 -2<br />

·10 -2<br />

-0.2615 0.9973 8.441<br />

0.8510 6.703<br />

7.707 0.9972 0.2488 -1.880<br />

0.838 3.24<br />

-<br />

43 Secondary air preheater<br />

·10 -3<br />

·10 -5<br />

·10 -3<br />

·10 -2<br />

effectiveness<br />

0.9894 1.668<br />

-7.330<br />

0.5778 1.403<br />

-4.135<br />

2.163 0.9912 4.904<br />

0.873 2.25<br />

-<br />

44 Primary air preheater<br />

·10 -3<br />

·10 -2<br />

·10 -5<br />

·10 -4<br />

·10 -2<br />

·10 -2<br />

effectiveness<br />

-0.6269 0.9976 1.140<br />

23.38 0.9984 0.4243 0.1144 0.9891 2.070<br />

45 Aggregated boiler effectiveness - 0.861 1.82<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

0.9956 3.869<br />

1.070<br />

0.9889 1.593<br />

4.404<br />

0.9984 9.969<br />

46 Air infiltration in preheaters % 13.0 3.56 -2.776<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

·10 -2<br />

0.9965 9.010<br />

1.618<br />

0.9936 1.775<br />

47 Carbon in ashes % 0.955 0.556 -0.4522 0.9974 0.2518 3.182<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -4<br />

Table 6.1.E. Summary <strong>of</strong> <strong>diagnosis</strong> results.<br />

208


6.4 Residual term <strong>analysis</strong>.<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

The idea <strong>of</strong> the <strong>diagnosis</strong> procedure is to share the variation <strong>of</strong> an efficiency<br />

indicator into several terms, each one corresponding to an independent free <strong>diagnosis</strong><br />

variable. Graphs shown in the previous section relating the variations <strong>of</strong> the free<br />

variables and their impacts prove that the method provides coherent results. However,<br />

to demonstrate the accuracy <strong>of</strong> the methodology, it should be checked that the residual<br />

term is low enough. This residual term ε is the difference between the variation <strong>of</strong> the<br />

efficiency indicator and the summation <strong>of</strong> all impacts originated by the free <strong>diagnosis</strong><br />

variables considered (section 3.3.1):<br />

Δ = ed ⋅Δ x +<br />

6.8<br />

t<br />

e d ε<br />

On the other hand, the good correlation among variables and impacts apparently<br />

shows that it would be enough to consider constant impact factors (ed av ), without<br />

applying the whole procedure for each period.<br />

In this section, residual term ε is studied for both the <strong>diagnosis</strong> method proposed<br />

(variable impact factors) and for a hypothetical approximated method which would use<br />

constant impact factors ed av (those appearing in table 6.1). Histograms are used to<br />

represent the distribution <strong>of</strong> the residual terms, for cycle heat rate, boiler efficiency and<br />

unit heat rate. First, the values <strong>of</strong> residuals are used directly. Then, non-dimensional<br />

values (calculated by dividing the residual term into the maximum impact are used).<br />

Results show that the <strong>diagnosis</strong> method used provides very low residuals, above all<br />

compared to residuals appearing when constant impact factors are used. In other words,<br />

the quantitative <strong>causal</strong>ity <strong>analysis</strong> provides good relation between accuracy and effort<br />

required.<br />

6.4.1 Residual term distribution.<br />

Figure 6.89 shows a histogram with the distribution <strong>of</strong> the residual term <strong>of</strong> cycle<br />

heat rate. Blue bars show the distribution when the proposed methodology is used<br />

(variable impact factors), while red bars show the residuals corresponding to constant<br />

impact factors. As it can be seen, residuals distribution with variable impact factor is<br />

more concentrated around the zero than the distribution with constant impact factor. For<br />

example, in the first case, more than 63% <strong>of</strong> points have residuals between -0.0015 and<br />

209


Chapter 6<br />

0.0015, while in the second this figure is lower than 40%. To put in context these<br />

figures, it should be taken into account that cycle heat rate has a variation range <strong>of</strong><br />

almost 0.2 (Figure 6.1). Another interesting characteristic <strong>of</strong> these graphs is that they<br />

are bell-shaped butt not necessarily symmetrical. This lack <strong>of</strong> symmetry appears<br />

because the state selected as reference does not correspond to the average values <strong>of</strong> all<br />

free <strong>diagnosis</strong> variables.<br />

Fraction <strong>of</strong> points<br />

210<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

< -0.0065<br />

-0.0065 -0.0055<br />

-0.0055 -0.0045<br />

-0.0045 -0.0035<br />

-0.0035 -0.0025<br />

-0.0025 -0.0015<br />

-0.0015 -0.0005<br />

-0.0005 0.0005<br />

0.0005 0.0015<br />

0.0015 0.0025<br />

0.0025 0.0035<br />

0.0035 0.0045<br />

0.0045 0.0055<br />

0.0055 0.0065<br />

0.0065 0.0075<br />

0.0075 0.0085<br />

0.0085 0.0095<br />

0.0095 0.0105<br />

0.0105 0.0115<br />

0.0115 0.0125<br />

0.0125 0.0135<br />

0.0135 0.0145<br />

0.0145 0.0155<br />

0.0155 0.0165<br />

Residual <strong>of</strong> cycle heat rate<br />

0.0165 0.0175<br />

0.0175 0.0185<br />

> 0.0185<br />

Figure 6.89. Distribution <strong>of</strong> residuals <strong>of</strong> cycle heat rate.<br />

Variable impact factors<br />

Constant impact factors<br />

The distribution <strong>of</strong> residuals <strong>of</strong> boiler efficiency appears in figure 6.90. Again, this<br />

distribution is appreciably more concentrated for variable impact factors: more than<br />

67% between -0.03% and 0.03% for variable impact factors and less than 40% between<br />

-0.01% and 0.05% for constant impact factors. The range <strong>of</strong> variation <strong>of</strong> boiler<br />

efficiency is about 4% (see Figure 6.2). Finally, the centre <strong>of</strong> the distribution<br />

corresponding to constant factor is slightly displaced. This effect appears because the<br />

factors correspond to the average slopes considering all points, which can be slightly<br />

different from the slope around the reference state.<br />

Finally, figure 6.91 shows the residual distribution for unit heat rate. By using<br />

variable impact factors, 56% <strong>of</strong> points have residuals between -0.015 and 0.015. This


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

percentage is only 26% when constant impact factors are used. As it can be seen in<br />

Figure 6.3, the range <strong>of</strong> variation <strong>of</strong> unit heat rate is about 0.2.<br />

Fraction <strong>of</strong> points<br />

Fraction <strong>of</strong> points<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

< -0.27<br />

-0.27 -0.25<br />

-0.25 -0.23<br />

-0.23 -0.21<br />

-0.21 -0.19<br />

-0.19 -0.17<br />

-0.17 -0.15<br />

-0.15 -0.13<br />

-0.13 -0.11<br />

-0.11 -0.09<br />

-0.09 -0.07<br />

-0.07 -0.05<br />

-0.05 -0.03<br />

-0.03 -0.01<br />

-0.01 0.01<br />

0.01 0.03<br />

0.03 0.05<br />

0.05 0.07<br />

0.07 0.09<br />

0.09 0.11<br />

0.11 0.13<br />

0.13 0.15<br />

0.15 0.17<br />

0.17 0.19<br />

0.19 0.21<br />

0.21 0.23<br />

> 0.23<br />

Residual <strong>of</strong> boiler efficiency (%)<br />

Figure 6.90. Distribution <strong>of</strong> residuals <strong>of</strong> boiler efficiency.<br />

< 0.0115<br />

-0.0105 -0.0095<br />

-0.0085 -0.0075<br />

-0.0065 -0.0055<br />

-0.0045 -0.0035<br />

-0.0025 -0.0015<br />

-0.0005 0.0005<br />

0.0015 0.0025<br />

0.0035 0.0045<br />

0.0055 0.0065<br />

0.0075 0.0085<br />

0.0095 0.0105<br />

0.0115 0.0125<br />

0.0135 0.0145<br />

0.0155 0.0165<br />

0.0175 0.0185<br />

Residual <strong>of</strong> unit heat rate<br />

0.0195 0.0205<br />

0.0215 0.0225<br />

> 0.0235<br />

Figure 6.91. Distribution <strong>of</strong> residuals <strong>of</strong> unit heat rate.<br />

Variable impact factors<br />

Constant impact factors<br />

Variable impact factors<br />

Constant impact factors<br />

211


Chapter 6<br />

6.4.2 Non-dimensional residual term distribution.<br />

Graphs presented previously demonstrated that residual terms are low in absolute<br />

values. However, it should be noted that the residual admitted in a <strong>diagnosis</strong> may vary<br />

according to the magnitude <strong>of</strong> the impacts. So that, it is interesting to use nondimensional<br />

residual terms, obtained by dividing each one <strong>of</strong> these residuals into the<br />

biggest impact <strong>of</strong> its <strong>diagnosis</strong> case.<br />

212<br />

ε<br />

ε = 6.8<br />

nd , i<br />

i<br />

max j<br />

i<br />

j<br />

Where<br />

( ) I<br />

i<br />

I j is the impact originated by the j th free <strong>diagnosis</strong> variable in the i th<br />

<strong>diagnosis</strong> test.<br />

Figure 6.92 shows the non-dimensional impact distribution <strong>of</strong> cycle heat rate. By<br />

using variable impact factors, 73% <strong>of</strong> points have residuals between -3 and 3% <strong>of</strong> the<br />

maximum impact, which can be considered as a very narrow range. The amount <strong>of</strong><br />

points whose residuals are out <strong>of</strong> the ±10% range is negligible. In the case <strong>of</strong> constant<br />

impact factors, only 47% <strong>of</strong> points below to the ±3% range.<br />

Fraction <strong>of</strong> points<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0.19<br />

Non-dimensional residual <strong>of</strong> cycle heat rate<br />

Figure 6.92. Distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> cycle heat rate.<br />

Variable impact factors<br />

Constant impact factors<br />

The distribution <strong>of</strong> residuals in boiler efficiency is also concentrated but less than in<br />

the previous case, as it can be seen in figure 6.93. The fraction <strong>of</strong> points with residuals<br />

in the ±3% range is 69%, but almost all cases (94%) are in the ±9% range, which can be


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

considered as an error low enough. Regarding the constant value approach, 29% <strong>of</strong> the<br />

points belong to the -5 to 1% range, and 70% to the -13 to 5% range.<br />

Fraction <strong>of</strong> points<br />

Fraction <strong>of</strong> points<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

< 0.25<br />

-0.25 -0.23<br />

-0.23 -0.21<br />

-0.21 -0.19<br />

-0.19 -0.17<br />

-0.17 -0.15<br />

-0.15 -0.13<br />

-0.13 -0.11<br />

-0.11 -0.09<br />

-0.09 -0.07<br />

-0.07 -0.05<br />

-0.05 -0.03<br />

-0.03 -0.01<br />

-0.01 0.01<br />

0.01 0.03<br />

0.03 0.05<br />

0.05 0.07<br />

0.07 0.09<br />

0.09 0.11<br />

0.11 0.13<br />

0.13 0.15<br />

0.15 0.17<br />

0.17 0.19<br />

0.19 0.21<br />

0.21 0.23<br />

0.23 0.25<br />

0.25 0.27<br />

0.27 0.29<br />

0.29 0.31<br />

0.31 0.33<br />

> 0.33<br />

Non-dimensional residual <strong>of</strong> boiler efficiency<br />

Figure 6.93. Distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> boiler efficiency<br />

0.25<br />

Non-dimensional residual <strong>of</strong> unit heat rate<br />

Figure 6.94. Distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> unit heat rate<br />

Variable impact factors<br />

Constant impact factors<br />

Variable impact factors<br />

Constant impact factors<br />

213


Chapter 6<br />

Figure 6.94 represents the distribution <strong>of</strong> non-dimensional residuals <strong>of</strong> unit heat rate.<br />

When variable impact factors are used, 73% <strong>of</strong> the points are in the ±3% range, and<br />

more than 99% in the ±9% range. With the approximation <strong>of</strong> constant impact factors,<br />

these fractions decrease to 47 and 86% respectively.<br />

6.4.3 Conclusion<br />

Graphs previously presented show that, when the proposed <strong>diagnosis</strong> method is<br />

used, about 70% <strong>of</strong> points have a non-dimensional residual term very reduced (in the<br />

±3% range). This value decreases to less than 50% is constant impact factors is used.<br />

These results are very important because prove that the quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong> proposed presents a good compromise between accuracy and implementation<br />

cost. The use <strong>of</strong> a simulator would provide an unappreciable accuracy increment but it<br />

entails a substantial increment in computation and implementation cost. On the other<br />

hand, accuracy decreases with the use <strong>of</strong> constant values. It should be noted that this<br />

difference would be higher if data <strong>of</strong> all load range would have been used. Besides, the<br />

use <strong>of</strong> constant value requires the assessment <strong>of</strong> the validity <strong>of</strong> these coefficients along<br />

time, while the method proposed is self-adjustable.<br />

6.5 Impacts aggregation<br />

Results shown in previous sections demonstrate that the quantitative <strong>causal</strong>ity<br />

<strong>diagnosis</strong> method is able to decompose the variations <strong>of</strong> boiler efficiency and cycle and<br />

unit heat rate into a summation <strong>of</strong> terms each one corresponding to an independent free<br />

<strong>diagnosis</strong> variable with a negligible residual term. The method provides an exhaustive<br />

<strong>analysis</strong> taking into account a large amount <strong>of</strong> independent causes. This fact constitutes<br />

and advantage but, due to the large amount <strong>of</strong> information provided, it sometimes<br />

makes it difficult to analyse the results.<br />

In this context, it can be remembered that impacts are additive, so that it is possible<br />

to group them in families. This allows to follow a zooming approach in the<br />

interpretation <strong>of</strong> results: first consider the evolution <strong>of</strong> the global efficiency indicators,<br />

then analyse the evolution <strong>of</strong> impacts due to a family <strong>of</strong> causes, and finally to analyse<br />

the impacts <strong>of</strong> individual causes, if necessary. It should be noted that this approach is<br />

not obliged by the methodology: the method provides in one step the impacts provided<br />

214


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

by each one <strong>of</strong> the free <strong>diagnosis</strong> variables, and the impact <strong>of</strong> a group <strong>of</strong> variables can<br />

be calculated later by adding the individual impacts. Obviously, in order to facilitate the<br />

analyst interpretation, results can be shown in a zooming manner: first, impacts <strong>of</strong><br />

groups <strong>of</strong> variables, and then individual impacts.<br />

The choice <strong>of</strong> the groups <strong>of</strong> variables depends on the interest <strong>of</strong> the analyst. A<br />

reasonable option for the working example is: ambient variables, fuel properties, setpoints<br />

for boiler and cycle and components efficiency indicators for boiler, cycle and<br />

cooling system. The distribution <strong>of</strong> free <strong>diagnosis</strong> variables into categories is detailed in<br />

section 5.3. Further aggregation can be made: for example, all boiler related variables<br />

when cycle heat rate is considered, or all set points. This strategy has been applied to the<br />

three global efficiency indicators <strong>of</strong> the working example: cycle heat rate, boiler<br />

efficiency and unit heat rate.<br />

6.5.1 Cycle heat rate<br />

Figure 6.95 shows the evolution <strong>of</strong> impact on cycle heat rate due to ambient<br />

conditions. The graph has a cyclic evolution originated by the variation <strong>of</strong> ambient<br />

conditions throughout the year. This impact has the highest standard deviation <strong>of</strong> all the<br />

variables groups considered: 0.0470.<br />

Influence <strong>of</strong> cycle set-points is represented in Figure 6.96, and has a standard<br />

deviation <strong>of</strong> only 0.01616. Although there is an important short-time dispersion, it can<br />

be appreciated some evolution along time and some differences among units.<br />

Finally, impact <strong>of</strong> variables related to boiler (fuel, boiler set-points and boiler<br />

components) are represented in Figure 6.97. It can be appreciated the higher influence<br />

during the first two years, caused by the variation <strong>of</strong> aggregated boiler effectiveness<br />

induced by the different operation <strong>of</strong> the economizer by-pass (see Fig. 6.83). In any<br />

case, the he standard deviation <strong>of</strong> this impact is very reduced: 0.003410.<br />

215


Chapter 6<br />

216<br />

Impact<br />

Impact<br />

0.10<br />

0.05<br />

-0.05<br />

-0.10<br />

-0.15<br />

Cycle heat rate impact due to ambient conditions<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

0.025<br />

0.020<br />

0.015<br />

0.010<br />

0.005<br />

-0.005<br />

-0.010<br />

-0.015<br />

-0.020<br />

Figure 6.95. Impact on cycle heat rate due to ambient conditions.<br />

Cycle heat rate impact due to cycle set-points<br />

0.000<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.96. Impact on cycle heat rate due to cycle set-points.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3


Impact<br />

Impact<br />

0.020<br />

0.015<br />

0.010<br />

0.005<br />

-0.005<br />

-0.010<br />

Cycle heat rate impact due to boiler<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

0.000<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.02<br />

-0.04<br />

-0.06<br />

Figure 6.97. Impact on cycle heat rate due to boiler.<br />

Cycle heat rate impact due to cooling system<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 5.98. Impact on cycle heat rate due to cooling system.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

217


Chapter 6<br />

Impact<br />

218<br />

0.15<br />

0.10<br />

0.05<br />

-0.05<br />

-0.10<br />

-0.15<br />

Cycle heat rate impact due to cycle components<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.99. Impact on cycle heat rate due to cycle components.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

After having removed effects <strong>of</strong> ambient condition, set-points and boiler, only the<br />

components <strong>of</strong> cooling system and cycle itself remain. Cooling system components<br />

appear in Figure 6.98. The main characteristic <strong>of</strong> this graph is the oscillating evolution,<br />

originated by the dependence <strong>of</strong> cooling tower effectiveness with ambient conditions<br />

(see Figure 6.77). To filtrate this dependence, <strong>causal</strong>ity chains theory is used later.<br />

Besides this cyclic behaviour, a slight tendency to increase along time (due to<br />

components degradation) can be observed. Standard deviation <strong>of</strong> this impact is 0.01917.<br />

Figure 6.99 shows the evolution <strong>of</strong> impact caused by cycle components. Although<br />

there is a lot <strong>of</strong> dispersion, it can be appreciated how impact tends to increase along<br />

time due to components degradation except in the springs <strong>of</strong> 1999 and 2004. Besides,<br />

some differences among units can be observed. This impact has a standard deviation <strong>of</strong><br />

0.04315.<br />

6.5.2 Boiler efficiency<br />

Variables affecting boiler efficiency can be classified into: ambient conditions, fuel,<br />

boiler set-points, boiler components and cycle influence. Figure 6.100 shows the impact<br />

on boiler efficiency due to ambient conditions. It can be appreciated the cyclic evolution


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

corresponding to the seasonal weather variation. This impact has a standard deviation <strong>of</strong><br />

0.4579 %.<br />

Impact (%)<br />

Impact (%)<br />

1.0<br />

0.5<br />

-0.5<br />

-1.0<br />

-1.5<br />

-2.0<br />

Boiler efficiency impact due to ambient conditions<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

2<br />

1<br />

-1<br />

-2<br />

-3<br />

-4<br />

Figure 6.100. Impact on boiler efficiency due to ambient conditions.<br />

Boiler efficiency impact due to fuel<br />

0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.101. Impact on boiler efficiency caused by fuel variation.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

219


Chapter 6<br />

Figure 6.101 represents the evolution <strong>of</strong> the impact on boiler efficiency due to fuel<br />

variation. It can be appreciated a long term evolution along with a short term dispersion<br />

with an amplitude <strong>of</strong> about 1%. This impact has a quite high standard deviation (0.5515<br />

%). However this figure is slightly lower than the standard deviation <strong>of</strong> the impacts<br />

caused by the coal HHV (0.5744%). This is result can be explained because there is a<br />

dependence among coal properties that compensate their impacts; when carbon or<br />

hydrogen content increase (and boiler efficiency decrease) HHV also increase (and<br />

boiler efficiency increase). All units consume the same coal, so that differences among<br />

them are negligible (only caused by the possible use <strong>of</strong> natural gas).<br />

Effect on boiler efficiency caused by cycle and cooling system (including<br />

components and set-points) is negligible, with a standard deviation <strong>of</strong> only 0.0244%.<br />

Evolution <strong>of</strong> this impact is plotted in Figure 6.102, where some peaks caused by the<br />

dependence on cooling tower effectiveness with ambient conditions.<br />

Impact (%)<br />

220<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

-0.05<br />

-0.10<br />

-0.15<br />

Boiler efficiency impact due to cycle<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.102. Impact on boiler efficiency caused by cycle and cooling system.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.103 shows the impact caused by components and set-points <strong>of</strong> the boiler.<br />

This aggregated impact has a standard deviation <strong>of</strong> 0.4968 %, which is higher than the<br />

deviation <strong>of</strong> ambient conditions and lower than that <strong>of</strong> fuel. However, the graph is


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

characterized by point dispersion, which makes difficult to appreciate tendencies. Only<br />

the lower efficiency during the first months, before the substitution <strong>of</strong> air preheaters,<br />

and a seasonal oscillation during the first years are clearly visible. It should be noted<br />

that heat transfer reduction is mainly caused by fouling, which has short time<br />

characteristic times, compared to degradation <strong>of</strong> other components such as turbines.<br />

Impact (%)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

-0.5<br />

-1.0<br />

-1.5<br />

-2.0<br />

-2.5<br />

Boiler efficiency impact due to components and set-points<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.103. Impact on boiler efficiency caused by components and set-points <strong>of</strong> the boiler.<br />

Finally, in order to clarify the picture provided by Figure 6.103, it is advisable to<br />

consider separately set-points and boiler components. Impact <strong>of</strong> set points has a<br />

standard deviation <strong>of</strong> 0.3020 % and its evolution is plotted in Figure 6.104. In this<br />

graph, point dispersion is low and a cyclic evolution can be appreciated, with better<br />

efficiency in summer than in winter. This phenomenon appears because the set-points<br />

variation in order to keep the average cold-side temperature <strong>of</strong> air preheaters above the<br />

condensation point. An improvement in boiler operation through the six years can also<br />

be appreciated.<br />

Figure 6.105 shows the evolution <strong>of</strong> impact caused by boiler components. The main<br />

characteristic <strong>of</strong> this graph is the seasonal oscillation appearing during the first three<br />

years, which corresponds to the boiler aggregated efficiency variation when economizer<br />

by-pass was used. Besides, there is big point dispersion, caused by short time fouling<br />

221


Chapter 6<br />

variation and responsible for dispersion appearing in Figure 6.103. This impact has a<br />

standard deviation <strong>of</strong> 0.4672 %.<br />

222<br />

Impact (%)<br />

Impact (%)<br />

1.5<br />

1.0<br />

0.5<br />

-0.5<br />

-1.0<br />

-1.5<br />

Boiler efficiency impact due to set points<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

-0.5<br />

-1.0<br />

-1.5<br />

-2.0<br />

-2.5<br />

Figure 6.104. Impact on boiler efficiency caused by boiler set-points.<br />

Boiler efficiency impact due to boiler components<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.105. Impact on boiler efficiency caused by boiler components.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

6.5.3 Unit heat rate<br />

After considering cycle heat rate and boiler efficiency separately, in this paragraph<br />

unit heat rate, which comprises the other two, is studied. Since the influence <strong>of</strong> setpoints<br />

and components separated by cycle, boiler and cooling system has been studied<br />

previously, in this part a more general vision is shown and impact <strong>of</strong> set-points and<br />

components efficiencies <strong>of</strong> all plants will be considered together.<br />

The characteristic cyclic evolution <strong>of</strong> impact on unit heat rate caused by ambient<br />

conditions is shown in Figure 6.106. It has a standard deviation <strong>of</strong> 0.03737, which is<br />

lower than the standard deviation <strong>of</strong> cycle heat rate because impact on boiler efficiency<br />

partially compensates impact on cycle.<br />

Impact<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

-0.10<br />

-0.12<br />

Unit heat rate impact due to ambient conditions<br />

0.00<br />

24/07/1998 0:00<br />

-0.02<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.106. Impact on unit heat rate caused by ambient conditions.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.107 represents the impact on unit heat rate caused by fuel variation. Since<br />

the main influence <strong>of</strong> fuel is on the boiler, this graph is similar to Figure 6.101 but<br />

inverted and with a scale change. Standard deviation <strong>of</strong> this impact is 0.02078. Impact<br />

caused by set-points is plotted in Figure 6.108. It can be appreciated a slight cyclic<br />

evolution and a decreasing tendency. However, importance <strong>of</strong> this family <strong>of</strong> variables is<br />

not high, with a standard deviation <strong>of</strong> 0.01203.<br />

223


Chapter 6<br />

Impact<br />

Impact<br />

224<br />

0.14<br />

0.12<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

Unit heat rate impact due to fuel<br />

0.00<br />

24/07/1998 0:00<br />

-0.02<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.02<br />

-0.03<br />

-0.04<br />

-0.05<br />

Figure 6.106. Impact on unit heat rate caused by fuel.<br />

Unit heat rate impact due to set-points<br />

0.00<br />

24/07/1998 0:00<br />

-0.01<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.107. Impact on unit heat rate caused by set-points.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

After removing the effects <strong>of</strong> ambient conditions, fuel and set points, only impact<br />

due to components remains (neglecting residual term). Evolution <strong>of</strong> this impact is<br />

plotted on Figure 6.108. In the first year, a big heat rate reduction appears due to<br />

improvements made on boilers (preheaters replacement). Then, heat rate tends to<br />

increase due to components degradation, although there is a reduction around 2003 or<br />

2004. Besides, differences among units can be appreciated: unit 3 is the best in most<br />

cases and unit 1 is the worst from 2002 to 2004 included. However, there is an<br />

oscillating behaviour which makes difficult to see clearly the results. This effect appears<br />

because <strong>of</strong> the dependence <strong>of</strong> the efficiency indicator chosen for the cooling tower with<br />

the ambient conditions. Standard deviation <strong>of</strong> this impact is 0.05756.<br />

Impact<br />

0.3<br />

0.2<br />

0.1<br />

-0.1<br />

-0.2<br />

-0.3<br />

Unit heat rate impact due to components<br />

0.0<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.108. Impact on unit heat rate caused by components.<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

6.5.4 Conclusion<br />

In this section, it has been illustrated how the aggregation <strong>of</strong> impacts can be very<br />

useful to provide a first decomposition <strong>of</strong> the variation <strong>of</strong> the global efficiency<br />

indicators. Then, it is possible to decompose further up to all individual free <strong>diagnosis</strong><br />

variables. This zooming strategy can be made because impacts are additive and helps to<br />

clarify the <strong>analysis</strong>, but it is not mandatory for the application <strong>of</strong> the <strong>diagnosis</strong><br />

procedure, because the methodology provides directly the impacts corresponding to all<br />

225


Chapter 6<br />

free <strong>diagnosis</strong> variables. In this point, the task <strong>of</strong> dividing the variation <strong>of</strong> the efficiency<br />

indicators is finished: an exhaustive <strong>analysis</strong> <strong>of</strong> all free <strong>diagnosis</strong> variables and its<br />

impact has been done, residual terms have been checked and proved to be low enough,<br />

and impacts has been grouped in order to clarify the <strong>analysis</strong>. However, a problem<br />

remains unsolved because there is a dependence <strong>of</strong> cooling tower efficiency with<br />

ambient conditions that introduces oscillation. This question is tackled in the next<br />

section.<br />

6.6 Causality chains<br />

In previous sections, it has been detected that a dependence <strong>of</strong> efficiency indicators<br />

with other variables, such as ambient conditions, is an issue that makes the <strong>diagnosis</strong><br />

more difficult. The most relevant example <strong>of</strong> this is the dependence <strong>of</strong> cooling tower<br />

effectiveness with ambient conditions. It should be highlighted that problem is not<br />

caused by the <strong>diagnosis</strong> method itself but by the choice <strong>of</strong> the free <strong>diagnosis</strong> variables.<br />

If another variable more linked to the physical behaviour <strong>of</strong> the equipment had been<br />

chosen, this effect would have been avoided, but this variable would not correspond to a<br />

widely accepted definition. A possibility to solve this task is the use <strong>of</strong> <strong>causal</strong>ity chains,<br />

which is applied in this paragraph. Of course, the case <strong>of</strong> the cooling tower is not the<br />

only example <strong>of</strong> crossed effects, however, is the only which makes more difficult the<br />

clear interpretation <strong>of</strong> results. Obviously, the same solution could be made to other<br />

situations in order to increase the precision <strong>of</strong> the <strong>analysis</strong>.<br />

The concept <strong>of</strong> <strong>causal</strong>ity chains was developed in point 3.5. The idea is to share the<br />

impact caused by a <strong>diagnosis</strong> variable into several impacts, one corresponding to the<br />

intrinsic variation <strong>of</strong> this variable, and others induced by some <strong>of</strong> the rest <strong>of</strong> the free<br />

<strong>diagnosis</strong> variables. To obtain this decomposition, a relation linking the variations <strong>of</strong> the<br />

variable whose impact has to be decomposed to the other free <strong>diagnosis</strong> variables (or to<br />

the other variables <strong>of</strong> the thermal representation <strong>of</strong> the system) is needed. This <strong>analysis</strong><br />

provides a <strong>causal</strong>ity chain because, after the decomposition <strong>of</strong> the global efficiency<br />

indicators into a summation <strong>of</strong> impacts, each one corresponding to a free <strong>diagnosis</strong><br />

variable, each impact can be decomposed again into an intrinsic component and a<br />

summation <strong>of</strong> induced components. This <strong>analysis</strong> is very useful to solve the dilemma<br />

appearing in the definition <strong>of</strong> free <strong>diagnosis</strong> variables corresponding to efficiency<br />

226


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

components. On one hand, it is preferable to use a definition widely accepted, easy to be<br />

calculated from properties <strong>of</strong> flows entering and exiting the component and without the<br />

need <strong>of</strong> tuning <strong>of</strong> complex models. On the other hand, it is interesting to choose a<br />

variable very close to the physical behaviour <strong>of</strong> the components, in order to avoid<br />

induced effects. When it is difficult to choose a variable able to accomplish with the two<br />

conditions, the use <strong>of</strong> <strong>causal</strong>ity chains <strong>analysis</strong> can help to the <strong>diagnosis</strong>. The idea is to<br />

choose a variable with widely accepted definition, easy to be calculated directly from<br />

properties <strong>of</strong> flows entering and exiting the component, and then to use a more or less<br />

sophisticated model to decompose this impact into intrinsic and induced components.<br />

The choice <strong>of</strong> the complexity level <strong>of</strong> the model depends on the difficulty <strong>of</strong> the<br />

modelling and the precision required in the task <strong>of</strong> filtering induced effects. In general, a<br />

theoretical or experimental correlation linking the variable with other variables (mainly<br />

ambient conditions) is enough. However, the developed theory allows to introduce a<br />

model linking the efficiency indicator <strong>of</strong> the component to the variables <strong>of</strong> the<br />

thermodynamic representation <strong>of</strong> the system, so that it is suitable to be used with any<br />

model, whatever its complexity.<br />

In this section, the approach is used to practically eliminate the effects induced by<br />

ambient conditions in the cooling tower effectiveness. The model used is very simple<br />

and links directly the cooling tower effectiveness to the ambient temperature. First, this<br />

experimental correlation is determined by using the large amount <strong>of</strong> operation points<br />

available. Then, the impact <strong>of</strong> the cooling tower effectiveness is decomposed in an<br />

intrinsic term and in a term induced by ambient temperature. Finally, the evolution <strong>of</strong><br />

some aggregated impacts taking into account this new decomposition is shown.<br />

6.6.1 Determination <strong>of</strong> a correlation for cooling tower effectiveness.<br />

Figure 6.109 shows the close relation between cooling tower effectiveness and<br />

ambient temperature for the three units, which is responsible for the seasonal oscillation<br />

detected in the evolution <strong>of</strong> the tower efficiency indicators. Correlations for this<br />

dependence for each one <strong>of</strong> the three units are also shown. As it can be seen, linear<br />

regression is enough to achieve a good fitting <strong>of</strong> the points. Since lines adjusting the<br />

three units are quite close, it is possible to use one correlation for all <strong>of</strong> them:<br />

ε cooling _ tower = 0.007096⋅ Tamb<br />

+ 0.3476 R 2 = 0.7821 6.9<br />

227


Chapter 6<br />

Cooling tower effectiveness<br />

228<br />

0.7<br />

0.65<br />

0.6<br />

0.55<br />

0.5<br />

0.45<br />

0.4<br />

0.35<br />

0.3<br />

y = 0.0072x + 0.3421<br />

R 2 = 0.8016<br />

Cooling tower effectiveness and ambient temperature<br />

y = 0.0069x + 0.3499<br />

R 2 = 0.7735<br />

y = 0.0073x + 0.3498<br />

R 2 = 0.7875<br />

0.25<br />

0 5 10 15 20 25 30 35<br />

Ambient temperature (ºC)<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Lineal (Unit 1)<br />

Lineal (Unit 2)<br />

Lineal (Unit 3)<br />

Figure 6.109. Dependence <strong>of</strong> cooling tower effectiveness with ambient temperature.<br />

6.6.2 Decomposition <strong>of</strong> the impact <strong>of</strong> cooling tower effectiveness.<br />

By using the correlation obtained above and the <strong>causal</strong>ity chains theory, the impact<br />

caused by cooling tower can be decomposed into two terms: one intrinsic and the other<br />

induced by ambient temperature. Figure 6.110 shows the intrinsic impact on cycle heat<br />

rate due to cooling tower effectiveness. Although high dispersion is present due to the<br />

high uncertainty in cooling water temperature measurements, it can be seen how the<br />

oscillating behaviour has disappeared. Consequently, standard deviation <strong>of</strong> this impact<br />

has dropped to 0.008906.<br />

Impact on cycle heat rate caused by ambient temperature through cooling tower (in<br />

other words, the induced part <strong>of</strong> the impact caused by cooling tower effectiveness), is<br />

plotted in Figure 6.111. It can be appreciated the cyclic behaviour that made difficult the<br />

<strong>diagnosis</strong> <strong>of</strong> the cooling tower by using directly its effectiveness. This impact has a<br />

standard deviation <strong>of</strong> 0.01671.<br />

Finally, figure 6.112 represents the total impact on cycle heat rate caused by ambient<br />

temperature. It is calculated by adding the direct impact <strong>of</strong> ambient temperature plus the<br />

impact <strong>of</strong> this variable through cooling tower effectiveness. This total impact has a<br />

standard deviation <strong>of</strong> 0.03609.


Impact<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

Intrinsic cycle heat rate impact due to cooling tower<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Figure 6.110. Impact <strong>of</strong> cooling tower effectiveness on cycle heat rate, after filtering effects<br />

induced by ambient temperature.<br />

Impact<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

-0.03<br />

Cycle heat rate impact due to ambient temperature through cooling tower<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.111. Impact on cycle heat rate due to cooling tower effectiveness variation induced<br />

by ambient temperature.<br />

229


Chapter 6<br />

Impact<br />

230<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

-0.10<br />

-0.12<br />

Total cycle heat rate impact due to ambient temperature<br />

0.00<br />

24/07/1998 0:00<br />

-0.02<br />

06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.112. Total impact on cycle heat rate due to ambient temperature, including induced<br />

effect on cooling tower effectiveness.<br />

Although it is not illustrated here, this new decomposition also affects unit heat rate.<br />

Intrinsic and induced impacts <strong>of</strong> cooling tower effectiveness on unit heat rate have<br />

standard deviations <strong>of</strong> 0.01008 and 0.01895, respectively. Combined impact on unit<br />

heat rate due to ambient temperature (both directly and through cooling tower<br />

effectiveness) has a standard deviation <strong>of</strong> 0.02600.<br />

6.6.3 Aggregated impacts<br />

After the decomposition <strong>of</strong> impact caused by cooling tower in two terms, all<br />

aggregated impacts including cooling tower effectiveness or ambient temperature vary.<br />

Two examples are shown here. Figure 6.113 shows the aggregated impacts <strong>of</strong> cooling<br />

system components on cycle heat rate, but using only the intrinsic component <strong>of</strong> cooling<br />

tower effectiveness. Although high point dispersion is present, the oscillation appearing<br />

in figure 6.98 is not present. Impact tends to increase during the first two years, then<br />

decreases suddenly and tend to increase slowly again. Standard deviation <strong>of</strong> this<br />

aggregated impact is 0.009020.


Impact<br />

Impact<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01<br />

-0.02<br />

Intrinsic ycle heat rate impact due to cooling system<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.113. Impact caused by cooling system on cycle heat rate after filtering effects<br />

induced by ambient temperature on cooling tower effectiveness.<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

-0.05<br />

-0.10<br />

-0.15<br />

-0.20<br />

Intrinsic unit heat rate impact due to components<br />

0.00<br />

24/07/1998 0:00 06/12/1999 0:00 19/04/2001 0:00 01/09/2002 0:00 14/01/2004 0:00 28/05/2005 0:00<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Figure 6.114. Impact caused by plant components on unit heat rate, after filtering effects<br />

induced by ambient temperature on cooling tower effectiveness.<br />

231


Chapter 6<br />

Aggregated impact on unit heat rate due to components but eliminating the induced<br />

effect <strong>of</strong> ambient temperature on cooling tower effectiveness is plotted in figure 6.114.<br />

Again, oscillations previously present (Figure 6.108) have been substantially reduced.<br />

During the first year, a big reduction appears due to pre-heaters replacement. Then, a<br />

steady increment is present due to component degradation, with reductions in spring<br />

2003 and 2004, corresponding to maintenance operations. Standard deviation <strong>of</strong> this<br />

aggregated impact is 0.04801.<br />

6.7 Conclusion<br />

6.7.1 Conclusions on the method<br />

In this chapter, results obtained in the application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

to the <strong>diagnosis</strong> <strong>of</strong> 6 years <strong>of</strong> operation <strong>of</strong> a three unit coal fired power plant have been<br />

shown. First, evolution <strong>of</strong> the three global efficiency indicators selected (boiler<br />

efficiency and cycle and unit heat rate) have been plotted. Then, all free <strong>diagnosis</strong><br />

variables have been studied: evolution <strong>of</strong> the variable and its impacts and relation<br />

between variable and impact. This exhaustive exposition has demonstrated the<br />

capability <strong>of</strong> the approach and its consistency. Results have been summarized by using<br />

suitable parameters such as standard deviation <strong>of</strong> the impacts and impact factors.<br />

Afterwards, residual term appearing in the decomposition <strong>of</strong> global efficiency<br />

indicators into impacts has been studied. This indicator has demonstrated to be low<br />

enough to ensure good accuracy <strong>of</strong> the procedure. Besides, its value is substantially<br />

lower than the residual obtained by using constant impact factor, which justifies the<br />

small additional computational effort required for the approach.<br />

The use <strong>of</strong> aggregated impacts facilitates the interpretation <strong>of</strong> results by using a<br />

zooming approach: from the aggregation <strong>of</strong> groups <strong>of</strong> variables to the detailed <strong>analysis</strong>.<br />

It should be noted that this approach is not mandatory for the application <strong>of</strong> the<br />

methodology, but an option to improve its usability.<br />

Finally, the use <strong>of</strong> <strong>causal</strong>ity chains allows to decompose an impact into an intrinsic<br />

term and a summation <strong>of</strong> terms induced by other free <strong>diagnosis</strong> variables. This<br />

232


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

procedure has been successfully applied for the filtration <strong>of</strong> effects induced by ambient<br />

conditions on cooling tower effectiveness.<br />

To sum up, results provided in this chapter clearly demonstrate the applicability <strong>of</strong><br />

the <strong>diagnosis</strong> procedure proposed for the <strong>analysis</strong> <strong>of</strong> complex energy systems, using real<br />

operation data available in every power plant.<br />

6.7.2 Practical results about the working example<br />

Besides its interest regarding the procedure, this chapter constitutes a true example<br />

<strong>of</strong> <strong>diagnosis</strong> <strong>of</strong> a thermal system. So that, valuable information has been obtained,<br />

which can be very useful for plant operators. The aim <strong>of</strong> this last section is to<br />

summarize the main results presented throughout the chapter.<br />

Unit heat rate was higher in the first months and then suddenly decreased. During<br />

the first three years, it was higher in unit 2. In the two following years, the highest heat<br />

rate was in unit 1 and the lowest in unit 3. Finally, the heat rate <strong>of</strong> unit 1 decreased<br />

substantially in the last year (Figure 6.1). Evolution <strong>of</strong> cycle heat rate (Figure 6.2) has a<br />

cyclic tendency, with higher values in summer and lower in winter. Besides, this<br />

parameter shows the same differences among units as unit heat rate. Boiler efficiency<br />

has also stationary evolution (higher values in summer) but in this parameter it is<br />

difficult to appreciate differences among units.<br />

6.7.2.1 Steam cycle<br />

Ambient temperature originates variations <strong>of</strong> cycle heat rate as big as 0.15 units, if<br />

effects induced on cooling tower are included (Figure 6.112). Relative humidity causes<br />

a variation <strong>of</strong> more than 0.03 units (Figure 6.9). Influence <strong>of</strong> cycle components causes<br />

variations <strong>of</strong> cycle heat rate <strong>of</strong> more than 0.10 units (Figure 6.99); impact increases<br />

during the first years and decreases in springs <strong>of</strong> 1999 and 2004.<br />

Isoentropic efficiency <strong>of</strong> high pressure <strong>of</strong> unit 1 is about 3 points higher than in unit<br />

3, which implies an impact on cycle heat rate <strong>of</strong> -0.015 units (Figure 6.52). Maintenance<br />

operations in the first medium pressure steam turbine in units 1 and 3 have originated<br />

increments <strong>of</strong> 3 points in their isoentropic efficiency, which implies almost 0.0075<br />

points <strong>of</strong> reduction <strong>of</strong> cycle heat rate (Figure 6.54). The turbine with highest influence<br />

on cycle heat rate is the low pressure one, with differences as big as 0.15 points (Figure<br />

6.58). Due to degradation, impact increases during the first years; then, unit 1 has the<br />

highest impact and unit 3 the lowest. Afterwards, impacts <strong>of</strong> unit 2 and 1 decrease as<br />

much as 0.05 units due to maintenance actions. It should be noted that, due to the lack<br />

233


Chapter 6<br />

<strong>of</strong> measurements, this variable does not include only the low pressure turbine itself but<br />

also other aspects which have not been included in other variables. It would be very<br />

interesting to have more information in order to further decompose these impacts.<br />

Influence <strong>of</strong> set-points can originate a variation in cycle heat rate <strong>of</strong> almost 0.02<br />

points, although no clear tendencies can be appreciated (Figure 6.96). A by-pass <strong>of</strong> 5 th<br />

water heater <strong>of</strong> unit 2 was performed in 2004, which implied an impact on cycle heat<br />

rate <strong>of</strong> 0.015 units.<br />

6.7.2.2 Boiler<br />

Ambient conditions (mainly ambient temperature) can originate a variation in boiler<br />

efficiency <strong>of</strong> 2 points (Figure 6.100). In winter, losses increase and efficiency decreases,<br />

and in summer losses decrease and efficiency increases. Variation <strong>of</strong> fuel quality causes<br />

an impact with short term variation <strong>of</strong> 1 point and long term variation <strong>of</strong> 2 points<br />

(Figure 6.101).<br />

Substitution <strong>of</strong> air preheaters in the first months <strong>of</strong> 1999 originated a reduction in air<br />

infiltration <strong>of</strong> around 20%, which caused an increment <strong>of</strong> boiler efficiency <strong>of</strong> more than<br />

0.5% (Figure 6.86 and 6.87). Carbon in ashes ranges from almost 0 to more than 2 %,<br />

and was lower during the first two years (Figure 6.87). Each point <strong>of</strong> carbon in ashes<br />

represents almost 0.5 points <strong>of</strong> reduction <strong>of</strong> boiler efficiency (Figure 6.88). However, it<br />

should be kept in mind that this variable actually depends on fuel quality and boiler<br />

operation. For example, the set-point <strong>of</strong> oxygen in flue gases leaving the boiler, which<br />

has varied from 1.5 to 2.5 %, and was higher during the first years (Figure 6.37). Each<br />

additional point <strong>of</strong> oxygen content entails a reduction <strong>of</strong> 0.25 points in boiler efficiency<br />

(Figure 6.38).<br />

Average cold-side temperatures in preheaters vary from less than 100 to more than<br />

130 ºC (Figures 6.39 and 6.41). Each degree causes an increment <strong>of</strong> unit heat rate <strong>of</strong><br />

0.001 units (secondary air preheaters, Figure 6.40) or 0.0003 units (primary air<br />

preheaters, Figure 6.42), due to the use <strong>of</strong> high pressure steam. The amount <strong>of</strong> steam<br />

used for sootblowing can reach up to 3 kg/s (Figure 6.43) and each kg/s causes an<br />

increment <strong>of</strong> unit heat rate <strong>of</strong> 0.08 units (Figure 6.44). It should be noted that this flowrate<br />

<strong>of</strong> steam is an average value which results <strong>of</strong> periods without sootblowing and<br />

periods with high flow rate.<br />

234


7 Application <strong>of</strong> quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong> to the quantification <strong>of</strong> intrinsic<br />

and induced malfunctions.<br />

In the previous chapter, the quantitative <strong>causal</strong>ity <strong>analysis</strong> method has proved its<br />

ability to diagnose an actual power plant, with coherent results and negligible error. On<br />

the other hand, thermoeconomic <strong>analysis</strong> provides tools and concepts which can be<br />

more homogeneous in their formulation but do not always guarantee good results due to<br />

the problem <strong>of</strong> induced malfunctions. In Chapter 4, a connection <strong>of</strong> the two approaches<br />

has been formulated, based on the application <strong>of</strong> the quantitative <strong>causal</strong>ity <strong>analysis</strong> to<br />

the decomposition <strong>of</strong> the unit energy consumptions and final plant products, which are<br />

the free variables <strong>of</strong> the thermoeconomic model. The aim <strong>of</strong> this chapter is to prove the<br />

practical application <strong>of</strong> these theoretical developments.<br />

First, the chosen productive structure for the case <strong>of</strong> work is described. Afterwards,<br />

the methodology is applied to an example <strong>of</strong> <strong>diagnosis</strong>. Finally, the correspondence <strong>of</strong><br />

the physical and thermoeconomic representations <strong>of</strong> the system is checked.<br />

7.1 Description <strong>of</strong> the productive structure.<br />

7.1.1 Simplified physical structure<br />

The first step to define a productive structure is to have a physical representation,<br />

which should be complete enough to take into account all the components <strong>of</strong> interest<br />

and simple enough to avoid too complex definitions <strong>of</strong> fuels and products. This<br />

simplified physical structure is represented in figure 7.1, while the flows are listed in<br />

Table 7.1 and the components in Table 7.2. Some <strong>of</strong> these components are macro-<br />

235


Chapter 7<br />

components which include several physical devices. For example, the furnace and all<br />

the heat exchangers <strong>of</strong> the boiler are included in only one component, and the low<br />

pressure feeding water heaters, the deaerator and the turbo-pump are included in another<br />

one. Besides, all flows <strong>of</strong> minor importance, such as steam for seals have not been<br />

considered. Finally, since all the <strong>analysis</strong> has been made by using the gross power<br />

production and gross heat rate, the electricity consumption <strong>of</strong> all ancillary devices<br />

(pumps, fans and mills) has not been considered.<br />

Flow Description<br />

236<br />

1 Air entering forced draft fans<br />

2 Air exiting primary air fans<br />

3 Air entering primary air coil heaters<br />

4 Tempering air<br />

5 Air from primary air coil heaters to primary air pre-heaters<br />

6 Air exiting primary air pre-heaters<br />

7 Hot air to sulphur removing unit<br />

8 Primary air to boiler<br />

9 Air from forced draft fans to secondary air coil heaters<br />

10 Air from secondary air coil heaters to secondary air pre-heaters<br />

11 Secondary air to boiler<br />

12 Flue gases entering primary air pre-heaters<br />

13 Flue gases exiting primary air pre-heaters<br />

14 Flue gases entering secondary air pre-heaters<br />

15 Flue gases exiting secondary air pre-heaters<br />

16 Flue gases exiting induced draft fans<br />

17 Steam to primary air coil heaters<br />

18 Condensed steam exiting primary air coil heaters<br />

19 Steam to secondary air coil heaters<br />

20 Condensed steam exiting secondary air coil heaters<br />

21 Flows leaving the boiler to deaerator<br />

22 Coal entering the boiler<br />

23 Natural gas entering the boiler<br />

Table 7.1. A. Description <strong>of</strong> flows <strong>of</strong> the simplified physical structure.


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

Flow Description<br />

24 Steam entering high pressure turbine<br />

25 Steam exiting high pressure turbine<br />

26 Steam to 6 th feeding water heater<br />

27 Steam entering the reheater<br />

28 Steam from the reheater to the 1 st intermediate pressure turbine<br />

29 Steam exiting 1 st intermediate pressure turbine<br />

30 Steam to 5 th feeding water heater<br />

31 Steam entering 2 nd intermediate pressure turbine<br />

32 Steam exiting 2 nd intermediate pressure turbine<br />

33 Steam to turbo-pump<br />

34 Steam to deaerator<br />

35 Steam entering low pressure turbine<br />

36 Steam to 3 rd feeding water heater<br />

37 Steam to 2 nd feeding water heater<br />

38 Steam to 1 st feeding water heater<br />

39 Steam entering the condenser<br />

40 Condensed water exiting the condenser<br />

41 Water exiting the turbo-pump<br />

42 Tempering water<br />

43 Water entering 5 th water heater<br />

44 Water from 5 th to the 6 th water heater<br />

45 Water from 6 th water heater to boiler<br />

46 Condensed water exiting 6 th water heater<br />

47 Condensed water exiting 5 th water heater<br />

48 Condensed water to condenser<br />

49 Cooling water entering the condenser<br />

50 Cooling water exiting the condenser<br />

51 Power produced by high pressure turbine<br />

52 Power produced by 1 st intermediate pressure turbine<br />

53 Power produced by 2 nd intermediate pressure turbine<br />

54 Power produced by low pressure turbine<br />

Table 7.1. B. Description <strong>of</strong> flows <strong>of</strong> the simplified physical structure.<br />

237


Chapter 7<br />

238<br />

24<br />

27<br />

28<br />

H<br />

26<br />

51<br />

45 44<br />

O<br />

46<br />

52<br />

53<br />

I J K<br />

25 29 31<br />

32<br />

30 33 34 36 37 38<br />

42<br />

N<br />

43<br />

47<br />

41<br />

Figure 7.1. Simplified physical diagram <strong>of</strong> the cycle.<br />

22<br />

23<br />

7<br />

8<br />

6<br />

11<br />

D<br />

13<br />

12<br />

G<br />

F<br />

Figure 7.2. Simplified physical diagram <strong>of</strong> the boiler.<br />

35<br />

M<br />

18 20 21<br />

10 9<br />

E C<br />

16<br />

14<br />

15<br />

5<br />

17<br />

18<br />

19<br />

20<br />

21<br />

24<br />

28<br />

27<br />

41<br />

44<br />

3 2<br />

B A<br />

4<br />

1<br />

L<br />

39<br />

40<br />

54<br />

48<br />

50<br />

49


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

Component Description<br />

A Forced draft fans and primary air fans<br />

B Primary air coil heaters<br />

C Secondary air coil heaters<br />

D Primary air pre-heaters<br />

E Secondary air pre-heaters<br />

F Boiler<br />

G Induced draft fans<br />

H High pressure steam turbine<br />

I 1 st intermediate pressure steam turbine<br />

J 2 nd intermediate pressure steam turbine<br />

K Low pressure steam turbine<br />

L Condenser<br />

M Low pressure water heaters, deaerator and turbo-pump<br />

N 5 th feeding water heater<br />

O 6 th feeding water heater<br />

Table 7.2. Description <strong>of</strong> components <strong>of</strong> the physical structure.<br />

Some coments should be made in order to explain this physical structure. First, the<br />

effects <strong>of</strong> induced draft fans on flue gases have been artificially translated to primary<br />

and secondary air pre-heaters, so that component G is actually only a mixer. Second,<br />

mechanical and electrical losses produced in the generator have been included in the<br />

low pressure steam turbine, which means that the flow 54 is calculated by substracting<br />

the power produced by that turbine minus mechanical and electrical losses<br />

corresponding to all turbines.<br />

Another important point is the reference for exergy calculation. It has been chosen<br />

an ambient temperature, which obviously may vary from one example to another. The<br />

first consequence <strong>of</strong> this decision is that exergy <strong>of</strong> flow 1 is always equal to zero. Other<br />

consequences are commented in subsequent sections.<br />

7.1.2 Productive structure<br />

From the previous physical structure, a productive structure has been built. Since<br />

forced and primary air fans are out <strong>of</strong> the scope <strong>of</strong> the <strong>diagnosis</strong>, the component A has<br />

239


Chapter 7<br />

been divided into two parts (primary and secondary air) which have been merged with<br />

components B and C. As a result, 15 components are obtained (taking into account the<br />

environment). They are described in Table 7.3.<br />

There are two waste flows: flue gases leaving the 6 th component and exergy<br />

increment <strong>of</strong> the cooling water. The cost <strong>of</strong> the first one is imputed to the boiler, while<br />

the cost <strong>of</strong> the second is shared among feeding water heaters, boiler and turbines,<br />

according to ratios proportional to the specific entropy generated in each component:<br />

240<br />

Δsi<br />

ρi<br />

=<br />

Δs<br />

tot<br />

Flow 7 (hot air to desulfuration) may be considered as a product <strong>of</strong> the plant or as a<br />

residue. However, since the aim <strong>of</strong> the sulphur removing unit is to reduce the SO2<br />

emissions which are produced in the boiler, and its amount is low compared to other<br />

residues, it has been considered as a productive flow which goes to the boiler.<br />

The FPR table appears in Table 7.4, and the productive structure is plotted in Figure<br />

7.3.<br />

Component Description<br />

0 Ambient<br />

1 Primary air coil heaters (includes primary air fan and forced draft fan)<br />

2 Secondary air coil heaters (includes forced draft fan)<br />

3 Primary air pre-heaters (includes induced draft fan)<br />

4 Secondary air pre-heaters (includes induced draft fan)<br />

5 Boiler<br />

6 Mixer <strong>of</strong> flue gases<br />

7 High pressure steam turbine<br />

8 1 st intermediate pressure steam turbine<br />

9 2 nd intermediate pressure steam turbine<br />

10 Low pressure steam turbine<br />

11 Condenser<br />

12 Low pressure feeding water heaters, deaerator and turbo-pump<br />

13 5 th feeding water heater<br />

14 6 th feeding water heater<br />

Table 7.3. Components <strong>of</strong> the productive structure<br />

7.1


F0 F1 F2 F3 F4 F5 F6 F7 F8 F9 F10 F11 F12 F13 F14 R5 R7 R8 R9 R10 R12 R13 R14 TOTAL<br />

P0 E22+E23 E22+E23<br />

P1 E4+E5 E4+E5<br />

P2 E10 E10<br />

E7+E8<br />

P3 E7+E8<br />

-E5<br />

-E5<br />

P4 E11-E10 E11-E10<br />

E26-<br />

E30+E46<br />

E39+E48<br />

E31-<br />

E28-<br />

E24-<br />

E13<br />

E40<br />

E14-<br />

E12-E13<br />

E19-<br />

P5 E17-<br />

(3)<br />

E46<br />

-E47<br />

(2)<br />

-E40<br />

(1)<br />

E23-<br />

E30-<br />

E26-<br />

+E15<br />

E15<br />

E20<br />

E18<br />

E34-<br />

E31<br />

E27<br />

E35<br />

P6 E16 E16<br />

P7 E51 E51<br />

P8 E52 E52<br />

P9 E53 E53<br />

P10 E54 E54<br />

P11 ρ5P11 ρ7P11 ρ8P11 ρ9P11 ρ10P11 ρ12P11 ρ13P11 ρ14P11 E50-E49<br />

P12 E42+E43-<br />

E42+E43-<br />

E40<br />

E40<br />

P13 E44+E43 E44+E43<br />

P14 E45+E44 E45+E44<br />

ρ7P11 ρ8P11 ρ9P11 ρ10P11 ρ12P11 ρ13P11 ρ14P11<br />

E16+<br />

E26-<br />

E30+E46<br />

E39+E48<br />

E31-<br />

E28-<br />

E13+E15 E24-<br />

E12+E13 E14-<br />

E19-<br />

E17-<br />

TOTAL E51+E52+<br />

ρ5P11<br />

E46<br />

-E47<br />

(2)<br />

-E40<br />

(1)<br />

E23-<br />

E30-<br />

E26-<br />

(4)<br />

E15<br />

E20<br />

E18<br />

E53+E54<br />

E34-<br />

E31<br />

E27<br />

E35<br />

(1) = E35-E36- E37-E38- E39; (2) = E18+E20+E21+E33+E34+E36+E37+E38+ E47-E48; (3) = E12+E14+E17+E19+ E21+E24-E27+E28; (4) = E7+E8+E11+E22+ E23+E42 +E45<br />

Table 7.4. FPR Table.<br />

241


Chapter 7<br />

242<br />

B 0,5<br />

B 5,1<br />

B 5,2<br />

B 5,3<br />

B 5,4<br />

B 5,6<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

B 1,5<br />

B 2,5<br />

B 3,5<br />

B 4,5<br />

R 6,5<br />

B 5,11<br />

11<br />

B 5,7<br />

Figure 7.3. Productive structure.<br />

R 11,7<br />

B 5,8<br />

R 11,8<br />

B 5,9<br />

R 11,9<br />

B 5,10<br />

R 11,10<br />

B 5,12<br />

R 11,12<br />

B 5,13<br />

R 11,13<br />

B 5,14<br />

R 11,14<br />

R 11,5<br />

7<br />

8<br />

9<br />

10<br />

12<br />

13<br />

14<br />

B 7,0<br />

B 8,0<br />

B 9,0<br />

B 10,0<br />

B 12,5<br />

B 13,5<br />

B 14,5


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

7.2 Example <strong>of</strong> thermoeconomic <strong>diagnosis</strong><br />

In this section, the productive structure described in section 7.1 and the methodology<br />

developed in chapter 4 are applied to diagnose a real situation <strong>of</strong> the working example.<br />

First <strong>of</strong> all, some preliminary results are reported (free <strong>diagnosis</strong> variables, exergy flows<br />

and costs and unit exergy costs). Afterwards, the decomposition <strong>of</strong> unit exergy<br />

consumption increment, final plant product variations and malfunctions into intrinsic<br />

and induced is presented. Finally, malfunction costs are analyzed.<br />

7.2.1 Preliminary results<br />

Operation <strong>of</strong> unit 1 on April 26 th <strong>of</strong> 2004 has been selected as an example to be<br />

diagnosed in comparison with the reference point used along all this work (unit 3 on<br />

September 17 th 2003). This choice provides a <strong>diagnosis</strong> example where a lot <strong>of</strong> free<br />

<strong>diagnosis</strong> variables change. The values <strong>of</strong> all these variables in the actual and reference<br />

cases and their variations appear in Table 7.5.<br />

Table 7.6 shows the values <strong>of</strong> the exergy and exergy costs <strong>of</strong> the productive flows,<br />

both in the actual and in the reference state. Exergy and exergy costs <strong>of</strong> residues flows<br />

appear in Table 7.7.<br />

Unit exergy costs are listed in Table 7.8. It is worth to highlight the high value <strong>of</strong> the<br />

products <strong>of</strong> components 1 and 2, due to the high irreversibility caused by using a high<br />

quality flow (steam) to slightly increase the temperature <strong>of</strong> the air entering the<br />

preheaters.<br />

Finally, unit exergy consumptions appear in Table 7.9 (associated with productive<br />

flows) and Table 7.10 (associated with residues).<br />

243


Chapter 7<br />

244<br />

Description units Actual Reference Increment<br />

1 Ambient temperature ºC 17.6 22.2 -4.6<br />

2 Relative humidity % 60.1 58.1 2.0<br />

3 Wind speed m/s 3.65 3.72 -0.07<br />

4 Coal high heating value kJ/kg 16912 17587 675<br />

5 Carbon mass fraction in coal % 42.09 44.55 -2.46<br />

6 Hydrogen mass fraction in coal % 2.53 2.79 -0.26<br />

7 Moisture mass fraction in coal % 19.65 17.53 2.12<br />

8 Ash mass fraction in coal % 23.16 23.57 -0.41<br />

9 Sulphur mass fraction in coal % 5.01 4.45 0.56<br />

10 Nitrogen mass fraction in coal % 0.67 0.70 0.03<br />

11 Energy provided by natural gas % 0.016 0.545 -0.529<br />

12 Main steam temperature ºC 538.06 538.53 -0.47<br />

13 Reheated steam temperature ºC 537.33 539.78 -2.45<br />

14 Main steam pressure bar 157.51 156.91 0.60<br />

15 Gross electric power MW 351.560 334.420 17.140<br />

16 Oxygen in flue gases leaving the<br />

boiler<br />

17 Average cold-side temperature in<br />

secondary air preheaters<br />

18 Average cold-side temperature in<br />

primary air preheaters<br />

% 1.87 1.90 -0.03<br />

ºC 109.9 116.8 -6.9<br />

ºC 105.0 116.0 -11.0<br />

19 Sootblowing steam flow rate kg/s 1.800 1.115 0.685<br />

20 Air for flue gas desulfuration unit kg/s 0 3.228 -3.228<br />

21 Tempering and primary air relation kg/kg 0.2786 0.3653 -0.0867<br />

22 Primary air-coal ratio kg/kg 2.268 1.800 0.468<br />

23 Temperature difference <strong>of</strong> flue<br />

gases entering preheaters<br />

ºC -0.28 -0.54 0.26<br />

24 Fraction <strong>of</strong> flue gases through PAH - 0.252 0.216 0.036<br />

25 High pressure turbine isoentropic<br />

efficiency<br />

% 81.46 81.50 -0.04<br />

Table 7.5.A. Free <strong>diagnosis</strong> variables in the actual and reference states.


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

Description Unit Actual Reference Increment<br />

26 Intermediate pressure 1 isoentropic<br />

efficiency<br />

27 Intermediate pressure 2 isoentropic<br />

efficiency<br />

% 85.15 83.48 1.67<br />

% 82.84 86.56 -3.72<br />

28 Low pressure isoentropic efficiency % 81.18 83.23 -2.05<br />

29 Intermediate<br />

coefficient<br />

pressure 1 flow<br />

30 Intermediate<br />

coefficient<br />

pressure 2 flow<br />

kg 1/2 ·m 3/2 ·<br />

s -1 ·bar -1/2<br />

kg 1/2 ·m 3/2 ·<br />

s -1 ·bar -1/2<br />

31 Low pressure flow coefficient kg 1/2 ·m 3/2 ·<br />

s -1 ·bar -1/2<br />

12.623 13.553 -0.930<br />

24.643 25.034 -0.391<br />

52.672 53.358 -0.686<br />

32 TTD 6 th water heater ºC -2.19 -1.76 -0.43<br />

33 TTD 5 th water heater ºC 1.22 -0.64 1.86<br />

34 TTD 3 rd water heater ºC -1.00 1.09 -2.09<br />

35 TDCA 6 th water heater ºC 13.65 10.15 3.50<br />

36 TDCA 5 th water heater ºC 0.00 0.00 0.00<br />

37 Pressure drop in reheater bar 0.80 1.57 -0.77<br />

38 Pressure increment in turbo-pump bar 181.83 171.48 10.35<br />

39 Water losses kg/s 1.634 3.752 -2.118<br />

40 Condenser effectiveness - 0.7293 0.6720 0.0573<br />

41 Cooling water flow rate kg/s 10147.7 10496.5 -348.8<br />

42 Cooling tower effectiveness - 0.4858 0.5463 -0.0605<br />

43 Secondary<br />

effectiveness<br />

air preheater<br />

- 0.8816 0.8305 0.0511<br />

44 Primary air preheater effectiveness - 0.8449 0.8918 -0.0468<br />

45 Aggregated boiler effectiveness - 0.8441 0.8515 -0.0074<br />

46 Air infiltration in preheaters kg/kg 0.1493 0.1047 0.0446<br />

47 Carbon in ashes % 2.29 0.90 1.39<br />

Table 7.5.B. Free <strong>diagnosis</strong> variables in the actual and reference state.<br />

245


Chapter 7<br />

Flow<br />

246<br />

Actual Reference<br />

Exergy Exergy cost Exergy Exergy cost<br />

B0,5 1014825 1014825 951236 951236<br />

B1,5 426.48 4164.5 236.37 2094.0<br />

B2,5 295.28 3667.5 641.32 6785.4<br />

B3,5 17539 50843 12630 36856<br />

B4,5 41665 138813 40189 130434<br />

B5,1 1710.7 4164.5 862.42 2094.0<br />

B5,2 1506.5 3667.5 2794.6 6785.4<br />

B5,3 20885 50843 15179 36856<br />

B5,4 57021 138813 53719 130434<br />

B5,5 2344.9 5708.4 1880.0 4564.8<br />

B5,6 37914 92298 37432 90889<br />

B5,7 108293 263630 106854 259448<br />

B5,8 66722 162429 57306 139143<br />

B5,9 64546 157132 62562 151904<br />

B5,10 170480 415022 159118 186350<br />

B5,11 62937 153215 59074 143437<br />

B5,12 59951 145947 52577 127661<br />

B5,13 23772 57870 22042 53518<br />

B5,14 31736 77260 24874 60397<br />

B7,0 96022 266212 94669 262096<br />

B8,0 62414 163573 53209 14234<br />

B9,0 58759 158867 58319 153188<br />

B10,0 134365 426173 128223 395718<br />

B12,5 52927 175597 46609 154728<br />

B13,5 22446 66270 20727 61683<br />

B14,5 29908 86284 23547 67923<br />

Table 7.6. Exergy and exergy costs <strong>of</strong> the productive flows (kW).


Flow<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

Actual Reference<br />

Exergy Exergy cost Exergy Exergy cost<br />

R6,5 37728 92298 37373 90889<br />

R11,5 28637 89529 27504 86287<br />

R11,7 825.81 2581.8 844.19 2648.5<br />

R11,8 365.75 1143.5 347.91 1091.5<br />

R11,9 555.04 1735.2 409.32 1284.2<br />

R11,10 3567.1 11152 2986.0 9368.0<br />

R11,12 9483.9 29650 8627.5 27067<br />

R11,13 2686.8 8399.6 2602.5 8164.7<br />

R11,14 2886.6 9024.5 2398.9 7526.2<br />

Table 7.7. Exergy and exergy cost <strong>of</strong> waste flows (kW).<br />

Unit cost Actual Reference Unit cost Actual Reference<br />

*<br />

k 1 1 *<br />

P,0<br />

P,8<br />

*<br />

k 9.7648 8.8590 *<br />

P,1<br />

P,9<br />

*<br />

k 12.420 10.580 *<br />

P,2<br />

P,10<br />

*<br />

k 2.8989 2.9181 *<br />

P,3<br />

P,11<br />

*<br />

k 3.3317 3.2455 *<br />

P,4<br />

P,12<br />

*<br />

k 2.4344 2.4281 *<br />

P,5<br />

P,13<br />

*<br />

k 2.4464 2.4319 *<br />

P,6<br />

P,14<br />

k 2.7724 2.7685<br />

*<br />

P,7<br />

Table 7.8. Unit exergy costs.<br />

k 2.6208 2.6356<br />

k 2.7037 2.6267<br />

k 3.1718 3.0862<br />

k 3.1263 3.1373<br />

k 3.3177 3.3197<br />

k 2.9525 2.9760<br />

k 2.8849 2.8846<br />

247


Chapter 7<br />

248<br />

Actual Reference Increment Actual Reference Increment<br />

κ 1.4297 1.4494 -1.9751·10<br />

0,5<br />

-4<br />

κ 6.0083·10<br />

1,5<br />

-4<br />

3.6017·10 -4 2.4066·10 -4<br />

κ 4.1599·10<br />

2,5<br />

-4 9.7722·10 -4 -5.6122·10 -4<br />

κ 2.4709·10<br />

3,5<br />

-2 1.9245·10 -2 5.4635·10 -3<br />

κ 5.8698·10<br />

4,5<br />

-2 6.1238·10 -2 -2.5400·10 -3<br />

κ 4.0112 3.6486 3.6255·10<br />

5,1<br />

-1<br />

κ 5.1020 4.3575 7.4446·10<br />

5,2<br />

-1<br />

κ 1.1908 1.2018 -1.1045·10<br />

5,3<br />

-2<br />

κ 1.3686 1.3367 3.1896·10<br />

5,4<br />

-2<br />

κ 3.3035·10<br />

5,5<br />

-3 2.8647·10 -3 4.3878·10 -4<br />

κ 1.0049 1.0016 3.3164·10<br />

5,6<br />

-3<br />

κ 1.1278 1.1287 -9.1996·10<br />

5,7<br />

-4<br />

κ 1.0690 1.0770 -7.9846·10<br />

5,8<br />

-3<br />

κ 1.0985 1.0727 2.5738·10<br />

5,9<br />

-2<br />

κ 1.2688 1.2409 2.7842·10<br />

5,10<br />

-2<br />

κ 1.2842 1.2921 -7.8836·10<br />

5,11<br />

-3<br />

κ 1.1327 1.1280 4.6674·10<br />

5,12<br />

-3<br />

κ 1.0591 1.0634 -4.3422·10<br />

5,13<br />

-3<br />

κ 1.0611 1.0564 4.7226·10<br />

5,14<br />

-3<br />

κ 7.4565·10<br />

12,5<br />

-2 7.1021·10 -2 3.5439·10 -3<br />

κ 3.1622·10<br />

13,5<br />

-2 3.1583·10 -2 3.8682·10 -5<br />

κ 4.2135·10<br />

14,5<br />

-2 3.5879·10 -2 6.2561·10 -3<br />

Table 7.9. Unit exergy consumptions.<br />

Actual Reference Increment Actual Reference Increment<br />

θ 5.3152·10<br />

6,5<br />

-2 5.6947·10 -2 -3.7946·10 -3<br />

θ 4.0345·10<br />

11,5<br />

-2<br />

4.1909·10 -2 -1.5640·10 -3<br />

θ 8.6002·10<br />

11,7<br />

-3 8.9172·10 -3 -3.1699·10 -4<br />

θ 5.8601·10<br />

11,8<br />

-3 6.5387·10 -3 -6.7855·10 -4<br />

θ 9.4461·10<br />

11,9<br />

-3 7.0187·10 -3 2.4274·10 -3<br />

θ 2.6548·10<br />

11,10<br />

-2 2.3288·10 -2 3.2603·10 -3<br />

θ 1.7919·10<br />

11,12<br />

-1 1.8510·10 -1 -5.9175·10 -3<br />

θ 1.1970·10<br />

11,13<br />

-1 1.2556·10 -1 -5.8574·10 -3<br />

θ 9.6516·10<br />

11,14<br />

-2 1.0188·10 -1 -5.3653·10 -3<br />

Table 7.10. Unit exergy consumptions associated with the residues.<br />

7.2.2 Decomposition <strong>of</strong> unit exergy consumptions and plant<br />

products.<br />

Once some basic results <strong>of</strong> thermoeconomic <strong>analysis</strong> have been shown, the aim <strong>of</strong><br />

this section is to apply quantitative <strong>causal</strong>ity <strong>analysis</strong> to decompose the variations <strong>of</strong> the<br />

independent variables <strong>of</strong> the thermoeconomic model (unit exergy consumptions and


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

final plant products) into a summation <strong>of</strong> terms corresponding to the free <strong>diagnosis</strong><br />

variables <strong>of</strong> the thermodynamic model.<br />

The complete set <strong>of</strong> results including the effect <strong>of</strong> the 47 free <strong>diagnosis</strong> variables<br />

would be too detailed, so that impacts are grouped into intrinsic and induced by ambient<br />

conditions, fuel quality, set points and other components. The classification <strong>of</strong> the free<br />

<strong>diagnosis</strong> variables into ambient conditions, fuel quality, set-points and component<br />

parameters is clear and was made in the definition <strong>of</strong> the free <strong>diagnosis</strong> variables in<br />

Chapter 5. However, the classification <strong>of</strong> component parameters into intrinsic and<br />

induced depends on the component, so that it should be specified here.<br />

Steam coil-heaters (components 1 and 2) have no free <strong>diagnosis</strong> variables considered<br />

as intrinsic. In the case <strong>of</strong> air-preheaters (3 and 4), intrinsic variables are their<br />

effectiveness and air infiltration. Variables associated with the boiler (5) are: aggregated<br />

boiler effectiveness, unburned carbon in ashes, pressure drop in reheater, water losses<br />

and pressure increment in turbo-pump. It should be noted that the last two variables<br />

could be associated with more than one component, but for simplicity are assigned to<br />

the boiler. There are no component variables associated with the flue gas mixer (6).<br />

Free <strong>diagnosis</strong> variables corresponding to the turbines (7 to 10) are isoentropic<br />

efficiencies and flow coefficients. The condenser (11) includes not only its effectiveness<br />

but also the flow rate <strong>of</strong> cooling water and the effectiveness <strong>of</strong> the cooling tower.<br />

Finally, the variables associated to the feeding water heaters (12 to 14) are the<br />

corresponding temperature differences (TTD and TDCA).<br />

Decomposition <strong>of</strong> the unit exergy consumptions <strong>of</strong> all the components is plotted in<br />

Figure 7.4. Due to the high variation <strong>of</strong> unit exergy consumption <strong>of</strong> components 1 and<br />

2, most <strong>of</strong> decompositions are difficult to see, so that a normalized version <strong>of</strong> Figure 7.4<br />

is plotted in Figure 7.5. It should be noted that a residual term (‘error’) is included in<br />

order to take into account the error produced by the quantitative <strong>causal</strong>ity <strong>analysis</strong> in the<br />

decomposition. This term is negligible in most cases; perhaps, it may seem that error in<br />

the normalized variation <strong>of</strong> κ5,6 is not very low, but it should be noted that the absolute<br />

increment <strong>of</strong> this variable is negligible. Components where intrinsic effects can be<br />

appreciated more clearly are the turbines (except the high pressure one, which has a<br />

very low variation). The condenser can also be diagnosed easily but effect induced by<br />

ambient conditions has to be filtered. The effect <strong>of</strong> ambient temperature is also<br />

important in the coil heaters. Influence <strong>of</strong> fuel quality is important in all components<br />

related to the boiler (1 to 6). Set-points have high influence in high pressure turbine,<br />

249


Chapter 7<br />

boiler, flue gas mixer and preheaters. Finally, the role <strong>of</strong> effects induced by other<br />

components is important in feeding water heaters, high pressure turbine, boiler, flue gas<br />

mixer and steam coil heaters.<br />

Component<br />

Component<br />

250<br />

k5_14<br />

k5_13<br />

k5_12<br />

k5_11<br />

k5_10<br />

k5_9<br />

k5_8<br />

k5_7<br />

k5_6<br />

k_5<br />

k5_4<br />

k5_3<br />

k5_2<br />

k5_1<br />

k ij decomposition<br />

-0.4 -0.2 0 0.2 0.4 0.6 0.8 1<br />

Variation<br />

Figure 7.4. Decomposition <strong>of</strong> unit exergy consumptions.<br />

k ij decomposition<br />

k5_14<br />

k5_13<br />

k5_12<br />

k5_11<br />

k5_10<br />

k5_9<br />

k5_8<br />

k5_7<br />

k5_6<br />

k_5<br />

k5_4<br />

k5_3<br />

k5_2<br />

k5_1<br />

-100% -80% -60% -40% -20% 0% 20% 40% 60% 80% 100%<br />

Normalized variation<br />

Figure 7.5. Normalized decomposition <strong>of</strong> unit exergy consumptions.<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

The detailed decomposition <strong>of</strong> all the unit exergy consumptions <strong>of</strong> the boiler (the<br />

only component which has several productive fuels) appears in Figure 7.6. It can be<br />

appreciated the high importance <strong>of</strong> induced effects, which can be justified by the low<br />

detail <strong>of</strong> the thermoeconomic model <strong>of</strong> this section <strong>of</strong> the plant.<br />

Figure 7.7 shows the decomposition <strong>of</strong> the unit exergy consumptions associated with<br />

the residues. Most <strong>of</strong> the variation is due to ambient conditions and induced effects.<br />

Importance <strong>of</strong> ambient conditions is due by their influence on cooling system and on the<br />

exergy <strong>of</strong> flue gases (which would be higher if constant reference value were used).<br />

Induced effects appear because residual flows not only depend on the component where<br />

they are charged, but mainly on the component from which they come (condenser or<br />

flue gases mixer).<br />

Finally, plant product decomposition is plotted on Figure 7.8. It should be noted that<br />

only the set point <strong>of</strong> gross power produced affects the total product <strong>of</strong> the plant<br />

(summation <strong>of</strong> the product <strong>of</strong> the four turbines), while the other variables only affect the<br />

distribution <strong>of</strong> this total power among the turbines. For example, the variation <strong>of</strong><br />

ambient conditions makes the power produced by the low pressure turbine to increase<br />

and the power produced by the other turbines to decrease exactly the same amount.<br />

Component<br />

k ij decomposition<br />

-0.04 -0.03 -0.02 -0.01 0 0.01 0.02<br />

Variation<br />

k14_5<br />

k13_5<br />

k12_5<br />

Figure 7.6. Decomposition <strong>of</strong> the unit exergy consumptions <strong>of</strong> boiler.<br />

k5_5<br />

k4_5<br />

k3_5<br />

k2_5<br />

k1_5<br />

k0_5<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

251


Chapter 7<br />

Component<br />

Component<br />

252<br />

theta ij decomposition<br />

theta11_14<br />

theta11_13<br />

theta11_12<br />

theta11_10<br />

theta11_9<br />

theta11_8<br />

theta11_7<br />

theta11_5<br />

theta6_5<br />

-0.04 -0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04<br />

Variation<br />

Figure 7.7. Decomposition <strong>of</strong> unit exergy consumptions associated with wastes.<br />

E10_0<br />

E9_0<br />

E8_0<br />

E7_0<br />

Plant product decomposition<br />

-6000 -4000 -2000 0 2000 4000 6000 8000 10000 12000<br />

Variation (kW)<br />

Figure 7.8. Decomposition <strong>of</strong> plant products.<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

7.2.3 Decomposition <strong>of</strong> malfunctions.<br />

The decomposition <strong>of</strong> unit exergy consumptions analyzed in the previous section is<br />

the origin <strong>of</strong> the decomposition <strong>of</strong> malfunctions. It should be noted that, due to the<br />

malfunction definition, the distribution <strong>of</strong> each malfunction among the different free<br />

<strong>diagnosis</strong> causes, is exactly the same as the distribution <strong>of</strong> the corresponding unit<br />

exergy consumption. However, the magnitude <strong>of</strong> malfunctions varies because <strong>of</strong> the<br />

different value <strong>of</strong> the component’s product.<br />

Figure 7.9 shows how the highest malfunction takes place in the boiler, and there are<br />

also significant malfunctions in preheaters, in condenser and in some turbines. Besides,<br />

induced effects are only important in the boiler. All malfunctions associated with the<br />

boiler are detailed in Figure 7.10. Finally, Figure 7.11 shows the importance <strong>of</strong> ambient<br />

conditions and induced effects in malfunctions associated with wastes; which justifies<br />

the separate accounting <strong>of</strong> them.<br />

Component<br />

Malfunction decomposition<br />

MF14<br />

MF13<br />

MF12<br />

MF11<br />

MF10<br />

MF9<br />

MF8<br />

MF7<br />

MF6<br />

MF5<br />

MF4<br />

MF3<br />

MF2<br />

MF1<br />

-20000 -15000 -10000 -5000 0 5000 10000 15000<br />

Malfunction (kW)<br />

Figure 7.9. Decomposition <strong>of</strong> malfunctions.<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

253


Chapter 7<br />

Component<br />

254<br />

Component<br />

Malfunction decomposition (boiler)<br />

MF14_5<br />

MF13_5<br />

MF12_5<br />

MF5_5<br />

MF4_5<br />

MF3_5<br />

MF2_5<br />

MF1_5<br />

MF0_5<br />

-30000 -25000 -20000 -15000 -10000 -5000 0 5000 10000<br />

Malfunction (kW)<br />

Figure 7.10. Decomposition <strong>of</strong> malfunctions corresponding to the boiler.<br />

Malfunction decomposition (residues)<br />

MR11_14<br />

MR11_13<br />

MR11_12<br />

MR11_10<br />

MR11_9<br />

MR11_8<br />

MR11_7<br />

MR11_5<br />

MR6_5<br />

-5000 -4000 -3000 -2000 -1000 0 1000 2000 3000 4000<br />

Malfunction (kW)<br />

Figure 7.11. Decomposition <strong>of</strong> malfunctions associated with the residues.<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

7.2.4 Malfunction cost decomposition. MFI and MFD tables.<br />

Finally, malfunctions originate malfunction costs or, in other words, impact on fuel.<br />

This magnitude is decomposed in Figure 7.12. In this graph, the fuel impact due to plant<br />

product variation ( MF ) has also been considered, and it can be seen how this term is<br />

*<br />

0<br />

the highest contribution to fuel demand increment. Again the decomposition into causes<br />

is identical to the decomposition <strong>of</strong> unit exergy consumptions, except in the components<br />

with more than one productive fuel (only the boiler). The decomposition <strong>of</strong> malfunction<br />

cost <strong>of</strong> the boiler can be seen in Figure 7.13, and malfunction costs associated with the<br />

residues are represented in Figure 7.14.<br />

Information contained in Figure 7.12 and 7.14 can be seen numerically in the Table<br />

<strong>of</strong> Intrinsic and Induced Malfunctions (MFI) (Table 7.11). Columns <strong>of</strong> this table<br />

correspond to malfunction costs induced by ambient conditions, fuel quality and set<br />

points, intrinsic malfunction costs, malfunction costs induced by other components and<br />

malfunction costs associated with the error <strong>of</strong> the decomposition <strong>of</strong> unit exergy<br />

consumptions variation. The rows correspond to the component; first, there are one row<br />

for each component (included the environment), to take into account malfunction costs<br />

associated with productive flows, and then, there are several rows to take into account<br />

malfunction costs associated with the wastes.<br />

The total summation <strong>of</strong> the table corresponds to the increment in the fuel <strong>of</strong> the<br />

plant. Since the fuel impact formula is exact, due to the use <strong>of</strong> the column<br />

corresponding to the error, the fuel impact obtained in the MFI table is also exact. This<br />

fact can be easily checked by comparing the obtained result with the increment <strong>of</strong> fuel<br />

(B05) which can be calculated from Table 7.6 (1014825 – 951236 = 63589 kW).<br />

More detailed information can be seen in the Table <strong>of</strong> Malfunctions caused by the<br />

Free Diagnosis variables (MFD), which takes into account the influence <strong>of</strong> the 47 free<br />

<strong>diagnosis</strong> variables considered (Table 7.12). It should be noted that this decomposition<br />

has to be performed also for the unit exergy consumptions, plant products and<br />

malfunctions, but results have not been detailed in the text <strong>of</strong> this section.<br />

255


Chapter 7<br />

Component<br />

Component<br />

256<br />

MF*14<br />

MF*13<br />

MF*12<br />

MF*11<br />

MF*10<br />

MF*9<br />

MF*8<br />

MF*7<br />

MF*6<br />

MF*5<br />

MF*4<br />

MF*3<br />

MF*2<br />

MF*1<br />

MF*0<br />

Malfunction cost decomposition<br />

-20000 -10000 0 10000 20000 30000 40000 50000 60000 70000<br />

Malfunction cost (kW)<br />

Figure 7.12. Decomposition <strong>of</strong> malfunction cost.<br />

Malfunction cost decomposition (boiler)<br />

MF*14_5<br />

MF*13_5<br />

MF*12_5<br />

MF*5_5<br />

MF*4_5<br />

MF*3_5<br />

MF*2_5<br />

MF*1_5<br />

MF*0_5<br />

-30000 -25000 -20000 -15000 -10000 -5000 0 5000 10000 15000 20000<br />

Malfunction cost (kW)<br />

Figure 7.13. Decomposition <strong>of</strong> malfunction cost corresponding to boiler.<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error


Component<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

Malfunction cost decomposition (residues)<br />

MR*11_14<br />

MR*11_13<br />

MR*11_12<br />

MR*11_10<br />

MR*11_9<br />

MR*11_8<br />

MR*11_7<br />

MR*11_5<br />

MR*6_5<br />

-15000 -10000 -5000 0 5000 10000 15000<br />

Malfunction cost (kW)<br />

Figure 7.14. Decomposition <strong>of</strong> malfunction cost associated with wastes.<br />

ambient<br />

fuel<br />

set-points<br />

intrinsic<br />

induced<br />

error<br />

257


Chapter 7<br />

258<br />

Ambient Fuel Set-points Intrinsic Induced Error Total<br />

*<br />

MF 1223.14 -22.659 49533.3 9650.78 -11836.4 -2.5492 48545.6<br />

0<br />

*<br />

MF 222.653 -96.836 -35.981 0 120.975 -2.188 208.623<br />

1<br />

*<br />

MF 1035.71 67.2392 166.959 0 -116.14 8.51615 1162.28<br />

2<br />

*<br />

MF -3.6009 -1015.8 -1265.4 1814.81 120.651 9.79102 -339.605<br />

3<br />

*<br />

MF 19.1812 2918.05 6252.61 -6613 407.857 135.862 3120.58<br />

4<br />

*<br />

MF 2952.33 254.984 -6893 3276.61 8641.55 950.207 9182.66<br />

5<br />

*<br />

MF -0.8266 91.0295 244.101 0 -130.852 98.2792 301.731<br />

6<br />

*<br />

MF 48.5634 -0.3782 -163.21 65.3341 -171.924 9.59065 -212.019<br />

7<br />

*<br />

MF -0.3597 0.87055 25.0896 -970.14 -55.5988 -34.133 -1034.27<br />

8<br />

*<br />

MF -6.3476 0.59917 29.3314 3432.97 84.1649 113.402 3654.12<br />

9<br />

*<br />

MF 82.5177 9.03462 -108.73 8768.67 96.2304 -157.01 8690.71<br />

10<br />

*<br />

MF 7387.96 2.63096 41.8375 -7606.6 -343.452 -359.85 -877.461<br />

11<br />

*<br />

MF 379.023 -5.3794 239.811 -660.8 601.261 -24.316 529.596<br />

12<br />

*<br />

MF 2.05732 10.6732 -67.333 42.307 -204.739 -2.0657 -219.1<br />

13<br />

*<br />

MF -7.8586 2.26523 23.019 1.88763 253.555 -2.1588 270.709<br />

14<br />

*<br />

MR -11997.3 620.007 -1394.76 1714.93 2677.62 -921.647 -9301.17<br />

5<br />

*<br />

MR -375.99 -2.8782 23.9363 6.61792 246.967 7.53239 -93.8177<br />

7<br />

*<br />

MR -152.38 -0.643 27.752 -126.26 128.89 9.76867 -112.874<br />

8<br />

*<br />

MR -222.81 -0.4341 43.3401 425.302 190.28 6.88719 442.566<br />

9<br />

*<br />

MR -1416.2 -1.8208 916.273 467.459 1437.15 -95.897 1306.93<br />

10<br />

*<br />

MR -5114.4 23.3432 752.908 -55.497 3346.96 184.464 -862.268<br />

12<br />

*<br />

MR -1141 -38.768 3.03088 47.1471 723.605 26.3784 -379.557<br />

13<br />

*<br />

MR -1045.4 -35.944 2.33185 -15.851 675.817 24.1149 -394.96<br />

14<br />

Total -8131.4 2779.15 48397.2 13666.7 6894.39 -17.024 63589<br />

Table 7.11. Table <strong>of</strong> intrinsic and induced malfunctions (MFI).


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

1 2 3 4 5 6 7 8 9 10 11 12<br />

MF 1304.656 -81.3984 -0.11719 181.5281 -211.646 -87.8132 79.66839 -10.6376 23.57984 -0.70412 3.366404 12.52364<br />

MF 222.6118 0.042635 -0.00149 2.145792 -94.903 -30.0754 27.24925 -4.31793 10.9898 -0.31858 -7.6056 0.00714<br />

MF 1035.894 -0.19438 0.00678 -10.4247 78.99232 17.97042 -16.2873 3.56597 -10.3787 0.264946 3.536239 -0.03255<br />

MF -3.92788 0.338766 -0.01182 16.64351 -988.771 -314.126 284.6037 -44.908 114.1737 -3.31304 -80.1475 0.056729<br />

MF 19.6353 -0.47047 0.016409 -21.9582 2970.643 602.0518 -545.464 131.3773 -395.927 9.943224 167.3882 -0.07878<br />

MF 2983.942 -5.28968 -26.3177 3806.129 -3105.17 -5576.91 5077.544 -541.922 962.2613 -26.6761 -340.272 85.4972<br />

MF -0.79334 -0.03448 0.001203 -1.73837 90.15011 20.12212 -18.2314 3.440256 -8.97071 0.242391 6.015195 -0.00577<br />

MF 51.5227 -2.95721 -0.0021 3.250444 -3.76013 -1.57024 1.424599 -0.1892 0.417655 -0.01251 0.061131 6.048637<br />

MF -0.41234 0.047956 0.004674 -7.24017 8.405663 3.499798 -3.17519 0.422738 -0.93496 0.027963 -0.1353 -0.10013<br />

MF -6.76778 0.417618 0.002585 -4.00448 4.775935 1.944898 -1.7645 0.239274 -0.53667 0.015893 -0.07118 -0.02184<br />

MF 87.93334 -5.45626 0.040631 -62.9369 74.64949 30.53738 -27.7049 3.742832 -8.37112 0.248393 -1.13061 -6.1898<br />

MF 7876.316 -488.364 0.0098 -15.1803 18.4923 7.400846 -6.71433 0.923733 -2.09417 0.061549 -0.2587 -0.41862<br />

MF 405.3078 -26.2784 -0.00599 9.270373 -15.3591 -4.81416 4.367227 -0.73919 1.90572 -0.05125 0.041031 1.461923<br />

MF 1.976032 0.024608 0.056674 -87.7888 102.0474 42.44509 -38.5083 5.13125 -11.3561 0.33949 -1.63685 -1.71643<br />

*<br />

0<br />

*<br />

1<br />

*<br />

2<br />

*<br />

3<br />

*<br />

4<br />

*<br />

5<br />

*<br />

6<br />

*<br />

7<br />

*<br />

8<br />

*<br />

9<br />

*<br />

10<br />

*<br />

11<br />

*<br />

12<br />

*<br />

13<br />

*<br />

14<br />

*<br />

5<br />

*<br />

7<br />

*<br />

8<br />

MF -8.45582 0.584989 0.01224 -18.9597 21.99616 9.163727 -8.31379 1.106343 -2.44596 0.073175 -0.35475 -4.875<br />

MR -12796.9 801.9109 -2.32584 3606.894 -3014.69 -129.206 119.2644 0.979112 -237.156 5.221115 268.7008 30.73837<br />

MR -400.844 24.86334 -0.01189 18.41892 -22.0991 -8.95526 8.124593 -1.10623 2.488774 -0.07354 0.323622 1.859251<br />

MR -162.472 10.0912 -0.00202 3.127595 -3.91872 -1.53268 1.390494 -0.195 0.448227 -0.01305 0.050171 0.090507<br />

MR -237.576 14.76712 -0.0003 0.468611 -0.95113 -0.25601 0.232229 -0.04489 0.12327 -0.00318 -0.00295 0.197701<br />

MR -1510.07 93.82978 0.0013 -2.01559 0.116428 0.813217 -0.73799 0.021942 0.082511 0.000312 -0.10162 3.306296<br />

MR -5442.63 328.3106 -0.11961 185.3054 -165.962 -86.0118 78.03864 -8.70231 16.34891 -0.55045 4.877077 4.870434<br />

MR -1215.31 74.55702 -0.19923 308.6056 -360.075 -149.306 135.4572 -18.0959 40.12789 -1.19794 5.715356 4.722556<br />

MR -1113.55 68.30965 -0.1846 285.9528 -333.667 -138.348 125.5156 -16.7686 37.1859 -1.11008 5.295162 4.382153<br />

*<br />

9<br />

*<br />

11<br />

*<br />

12<br />

*<br />

13<br />

*<br />

14<br />

Total -8909.93 807.6526 -29.1475 8195.493 -4950.7 -5792.97 5275.979 -496.676 531.9625 -17.5854 33.65363 142.3236<br />

Table 7.12.A. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD).<br />

259


Chapter 7<br />

13 14 15 16 17 18 19 20 21 22 23 24<br />

*<br />

MF -42.0355 -23.8973 49588.01 -5.30926 21.74434 9.913834 -32.8979 5.901595 -21.9988 18.03994 0.252517 3.072405<br />

0<br />

*<br />

MF 0.026378 -0.01539 2.463902 -2.42385 -0.11868 119.1511 0.090475 59.99041 -93.448 -390.695 0.324015 268.6661<br />

1<br />

*<br />

MF -0.12026 0.070171 -11.2333 2.076838 80.27113 0.241327 -0.41249 0.148208 0.390187 165.9861 -0.13199 -70.2948<br />

2<br />

*<br />

MF 0.209588 -0.12229 19.57739 -25.2085 -0.94301 570.9798 0.718888 658.7732 -1188.48 -4108.78 1.721405 2806.082<br />

3<br />

*<br />

MF -0.29107 0.169837 -27.1883 78.25897 1856.005 0.584089 -0.99836 0.358712 0.944379 7466.796 -1.96794 -3119.98<br />

4<br />

*<br />

MF 373.4518 -229.607 -774.065 -177.842 -2850.22 -1290.75 1988.493 -2294.78 1563.445 -2896.81 -0.63453 -389.193<br />

5<br />

*<br />

MF -0.02133 0.012449 -1.99286 1.972558 -23.5978 58.61209 -0.07318 -41.9616 75.74553 328.3632 -0.27526 -152.677<br />

6<br />

*<br />

MF -3.75002 15.13221 -180.636 -0.09425 0.465197 0.211869 -0.69289 0.126315 -0.43111 0.345412 0.004502 0.065508<br />

7<br />

*<br />

MF 24.06949 0.011674 1.926221 0.210775 -0.95883 -0.43688 0.597043 -0.2603 0.922323 -0.74654 -0.01005 -0.13521<br />

8<br />

*<br />

MF 22.18058 -0.03826 6.419347 0.120062 -0.20551 -0.09452 1.028783 -0.05557 0.350807 -0.31702 -0.00564 -0.02984<br />

9<br />

*<br />

MF 39.55719 12.73926 -151.637 1.875653 -4.28527 -1.96349 2.297685 -1.16061 6.031169 -5.2941 -0.08839 -0.61504<br />

10<br />

*<br />

MF 79.70012 0.599622 -48.7041 0.465783 0.213486 0.091267 9.893751 0.059468 0.843001 -0.90879 -0.02164 0.024205<br />

11<br />

*<br />

MF -23.5673 -2.49211 251.4586 -0.39615 -10.5436 -4.77231 30.89781 -2.87038 4.59297 -2.51895 0.015926 -1.45559<br />

12<br />

*<br />

MF -28.4439 0.341022 -28.682 2.559178 -11.3017 -5.15038 7.777932 -3.06795 11.02429 -8.95628 -0.12192 -1.59458<br />

13<br />

*<br />

MF 1.279835 -6.30119 35.12431 0.551524 -2.5508 -1.16214 1.562717 -0.69252 2.43486 -1.96675 -0.0263 -0.3596<br />

14<br />

*<br />

MR -79.2753 -53.4627 314.2242 15.62484 -1786.54 -792.222 -54.6493 767.3913 -1608.6 2067.027 -0.95624 -214.067<br />

5<br />

*<br />

MR -2.42945 1.695775 26.49096 -0.55586 0.607801 0.281886 -4.205 0.163761 -1.44804 1.358567 0.026037 0.090569<br />

7<br />

*<br />

MR 1.979683 -0.40591 28.01453 -0.09895 -0.32248 -0.14495 -1.30139 -0.08805 -0.03708 0.105028 0.004532 -0.04352<br />

8<br />

*<br />

MR 3.890206 -0.44495 42.84743 -0.02483 -0.98046 -0.44392 -1.49184 -0.26688 0.45167 -0.25943 0.000921 -0.1355<br />

9<br />

*<br />

MR 169.3666 -7.60226 763.9547 -0.00241 -5.96152 -2.70094 -2.79989 -1.6223 3.05001 -1.88885 -0.00132 -0.82556<br />

10<br />

*<br />

MR -18.0079 -8.82312 637.6214 -4.04369 150.4698 68.22023 -110.349 40.93513 -85.3755 56.2809 0.224417 20.88462<br />

12<br />

*<br />

MR -2.68788 -0.91305 -19.07 -9.0333 36.28284 16.54437 -31.623 9.846951 -37.0636 30.46691 0.429467 5.128653<br />

13<br />

*<br />

MR -3.46628 -0.85009 -16.9174 -8.37085 33.56043 15.30317 -29.4584 9.108059 -34.314 28.21307 0.397958 4.744011<br />

14<br />

Total 511.6153 -304.203 50458 -129.688 -2518.91 -1239.71 1772.406 -794.024 -1400.96 2743.842 -0.83952 -842.652<br />

Table 7.12.B. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD).<br />

260


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

25 26 27 28 29 30 31 32 33 34 35 36<br />

*<br />

MF 6.273276 -434.192 889.8358 -939.518 -1498.4 69.78665 360.9716 -15.579 121.4205 190.0653 24.815 0<br />

0<br />

*<br />

MF 0.004045 -0.08998 0.233577 0.484826 -0.11091 0.000574 0.003017 -0.00886 0.018564 -0.04422 0.003794 0<br />

1<br />

*<br />

MF -0.01844 0.410234 -1.06492 -2.2104 0.505657 -0.00261 -0.01375 0.040383 -0.08464 0.20162 -0.0173 0<br />

2<br />

*<br />

MF 0.032138 -0.71495 1.855927 3.852272 -0.88126 0.004557 0.023972 -0.07038 0.147507 -0.35138 0.030146 0<br />

3<br />

*<br />

MF -0.04463 0.9929 -2.57744 -5.34989 1.223853 -0.00633 -0.03329 0.09774 -0.20485 0.487985 -0.04187 0<br />

4<br />

*<br />

MF 1.724259 13.76189 -43.2129 -173.263 1242.975 34.63315 106.7058 136.868 -79.8802 40.3699 -2.70955 0<br />

5<br />

*<br />

MF -0.00327 0.072778 -0.18892 -0.39214 0.089706 -0.00046 -0.00244 0.007164 -0.01502 0.035769 -0.00307 0<br />

6<br />

*<br />

MF 65.3341 7.004463 -17.3257 -35.5445 -158.086 -3.31513 -0.75549 0.179674 5.688694 3.090142 1.162612 0<br />

7<br />

*<br />

MF -0.0013 -1168.28 -0.82887 0.376965 198.1439 -41.3691 -2.67998 0.019077 -9.45699 -0.79757 -1.93275 0<br />

8<br />

*<br />

MF 0.012709 21.0578 3409.263 1.23515 101.7835 23.70569 -36.4824 0.063217 -4.78133 -10.4949 -0.97717 0<br />

9<br />

*<br />

MF -3.3483 112.5414 -302.978 8922.126 355.3157 5.062669 -153.452 7.150255 -28.2666 -142.379 -5.77692 0<br />

10<br />

*<br />

MF -0.15116 81.22939 -231.745 -518.419 372.6457 7.21368 -10.0845 0.194641 -18.3256 2.487351 -3.74524 0<br />

11<br />

*<br />

MF 0.69124 -43.9034 126.8996 40.07616 -130.146 -4.28096 3.614106 0.487232 14.31681 -660.804 2.925962 0<br />

12<br />

*<br />

MF -0.11648 -32.6147 -27.9251 -5.69511 -159.077 99.84244 -66.8793 -0.90604 42.30697 -18.4991 -5.56701 0<br />

13<br />

*<br />

MF 0.833511 4.023239 2.175287 6.909105 245.9394 -47.9819 -3.01856 -31.2331 103.7325 -1.50129 33.12075 0<br />

14<br />

*<br />

MR -0.0825 -180.34 480.9833 1013.749 -2763.5 -48.615 -13.504 -202.914 145.9169 -97.2432 29.82139 0<br />

5<br />

*<br />

MR 6.61792 -6.26146 16.49369 34.50674 -79.6634 -0.89133 -0.1364 -2.72778 3.326594 -3.22088 0.679864 0<br />

7<br />

*<br />

MR 0.092597 -136.463 7.858294 16.52252 10.20462 -4.73568 -0.30417 -0.1243 3.092182 -1.66428 0.631957 0<br />

8<br />

*<br />

MR 0.118034 1.118525 426.1643 24.46143 15.57404 -0.86244 -5.22309 -0.11813 -0.43154 -3.82089 -0.0882 0<br />

9<br />

*<br />

MR 2.023727 104.9565 -311.954 496.0681 674.7338 14.1166 -28.609 -2.8156 -25.005 -54.1859 -5.11033 0<br />

10<br />

*<br />

MR 2.112628 -79.2926 205.1851 471.1814 -459.946 -0.71841 -55.6019 -31.9603 19.81495 -55.4975 4.049629 0<br />

12<br />

*<br />

MR 0.198116 -11.0485 33.74628 81.39292 -136.372 -20.3892 -15.062 -11.3096 47.14715 -12.1739 1.164193 0<br />

13<br />

*<br />

MR 0.18518 -12.7568 37.28025 78.64327 -203.423 -15.7952 -1.13587 -17.6908 37.25888 -7.56157 1.839978 0<br />

14<br />

Total 82.4874 -1758.79 4698.173 9511.194 -2370.47 65.40217 78.3398 -172.351 377.7364 -833.502 74.27587 0<br />

Table 7.12.C. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD).<br />

261


Chapter 7<br />

37 38 39 40 41 42 43 44 45 46 47 Error Total<br />

*<br />

MF -108.784 -282.609 100.4448 411.0331 -272.868 -753.927 -52.031 17.28126 -37.6365 -12.5834 30.53384 -2.54920 48545.6<br />

0<br />

*<br />

MF -0.07332 0.075883 -0.11489 -0.20379 0.135285 0.373788 -0.23687 98.75705 21.29866 0.080876 0.387442 -2.18796 208.623<br />

1<br />

*<br />

MF 0.334281 -0.34596 0.5238 0.929092 -0.61678 -1.70416 -84.4591 -0.32514 27.92661 -54.3816 -1.76641 8.51615 1162.28<br />

2<br />

*<br />

MF -0.58258 0.60294 -0.91287 -1.61921 1.074928 2.970003 -1.88213 1814.808 113.3507 0.642614 3.078493 9.79102 -339.605<br />

3<br />

*<br />

MF 0.809067 -0.83734 1.267765 2.248702 -1.49282 -4.12462 -7738.93 -0.78695 420.5048 1125.941 -4.27529 135.862 3120.58<br />

4<br />

*<br />

MF -1668.49 667.9024 -3108 222.2448 -147.539 -407.647 8148.206 -2283.06 528.1301 1831.378 6857.057 950.207 9182.66<br />

5<br />

*<br />

MF 0.059303 -0.06138 0.092925 0.164826 -0.10942 -0.30233 -38.3438 -65.7777 -1.46109 -24.4 -0.31337 98.2792 301.731<br />

6<br />

*<br />

MF 53.2862 -6.03813 2.118769 14.91682 -9.90265 -27.3608 -1.01303 0.337877 -0.64359 -0.27155 0.546736 9.59065 -212.019<br />

7<br />

*<br />

MF -0.02789 0.226712 -1.79539 -0.27639 0.183481 0.506953 2.173464 -0.72359 1.464501 0.5577 -1.21782 -34.1328 -1034.27<br />

8<br />

*<br />

MF -0.47473 12.24468 -3.15484 -2.12357 1.40975 3.895107 0.853662 -0.27842 0.939889 0.110455 -0.67358 113.402 3654.12<br />

9<br />

*<br />

MF 61.16528 69.26361 -6.70373 27.19185 -18.0515 -49.876 14.5489 -4.77152 14.34987 2.379321 -10.5864 -157.013 8690.71<br />

10<br />

*<br />

MF 3.083112 -0.05152 -30.4926 -2893.12 -199.868 -4513.61 2.17127 -0.68327 3.959853 -0.18634 -2.55347 -359.851 -877.461<br />

11<br />

*<br />

MF -17.792 894.4639 -95.4524 132.4601 -87.9347 -242.962 9.845576 -3.48734 -6.58229 6.461725 1.559772 -24.3156 529.596<br />

12<br />

*<br />

MF 1.819416 6.460572 -23.4371 -0.5452 0.361934 1.000015 26.00581 -8.65206 17.88714 6.564521 -14.7664 -2.06573 -219.100<br />

13<br />

*<br />

MF -65.4128 2.063589 -4.69888 -3.03869 2.017262 5.57365 5.734457 -1.90982 3.819091 1.484785 -3.18909 -2.1588 270.709<br />

14<br />

*<br />

MR 221.8016 106.0253 366.8655 -873.315 1827.436 7380.008 -4374.55 1365.797 414.24 -1012.03 605.9958 -921-647 -9301.17<br />

5<br />

*<br />

MR 14.95033 2.560436 12.88528 -28.2537 57.15216 229.6389 -3.56446 1.154072 -4.45804 -0.30233 3.098206 -1278.51 -93.8177<br />

7<br />

*<br />

MR -1.88286 1.464764 4.007028 -12.4797 23.47555 93.24106 -0.14857 0.036347 -0.92722 0.208163 0.526103 356.862 -112.874<br />

8<br />

*<br />

MR -2.51035 3.417351 4.616602 -18.5553 34.43957 136.482 0.977766 -0.34408 -0.51167 0.599428 0.078863 7.53239 442.566<br />

9<br />

*<br />

MR -40.9752 85.52718 8.717301 -99.9451 214.0292 867.2455 6.714348 -2.33672 -1.86946 3.626693 -0.33881 6.88719 1306.93<br />

10<br />

*<br />

MR 33.72449 -129.898 338.6576 -663.796 832.634 3032.801 -190.728 65.73884 12.87438 -91.0408 31.1642 -95.8971 -862.268<br />

12<br />

*<br />

MR 25.4445 6.487937 95.65031 -77.5644 168.982 686.3822 -87.7217 29.1226 -64.257 -20.9757 51.90887 184.464 -379.557<br />

13<br />

*<br />

MR 42.47164 4.43736 89.11387 -71.5708 154.9518 628.8398 -81.2192 26.96272 -59.564 -19.3999 48.09856 26.3784 -394.960<br />

14<br />

Total -1448.05 1443.384 -2249.8 -3935.21 2579.9 7067.449 -4437.59 1046.856 1402.835 1744.465 7594.353 24.1149 63588.99<br />

Table 7.12.D. Table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong> variables (MFD)<br />

262


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

7.3 General quantification <strong>of</strong> intrinsic and induced effects.<br />

In the previous section, the capability <strong>of</strong> the approach to perform a detailed<br />

thermoeconomic <strong>diagnosis</strong> <strong>of</strong> a thermal system has been demonstrated in a practical<br />

example.<br />

Besides, the importance <strong>of</strong> induced effects has been quantified, which is a very<br />

important result for assessing the suitability <strong>of</strong> a given thermoeconomic model to be<br />

used for <strong>diagnosis</strong>. However, it should be noted that it was only an example <strong>of</strong> only one<br />

situation; so that, results can only be seen as indicative. In order to test the accuracy <strong>of</strong><br />

the chosen productive structure, information <strong>of</strong> a set <strong>of</strong> situations as wide as possible<br />

should be used.<br />

In this section, a ‘representative’ example has been developed in order to serve as a<br />

general test. This example is a fictitious situation where the reference state is the same<br />

as in all previous <strong>analysis</strong>, and the work state is built by adding to the reference state an<br />

increment <strong>of</strong> all free <strong>diagnosis</strong> variables equal to their standard deviation (which can be<br />

seen in Table 6.1). This approach enables to take into account the typical variations <strong>of</strong><br />

all variables. Since it is a fictitious situation, the derivatives are calculated in the<br />

reference state and it has no sense to consider residuals.<br />

7.3.1 Decomposition <strong>of</strong> unit exergy consumptions and plant<br />

products.<br />

Decomposition <strong>of</strong> unit exergy consumptions is plotted in Figure 7.15, where it can<br />

be seen how the highest variations correspond to the coil heaters, characterized by the<br />

high influence <strong>of</strong> ambient conditions. In order to see the decomposition <strong>of</strong> the other unit<br />

exergy consumptions, these two cases have been removed from Figure 7.16: ambient<br />

conditions have an important influence on the condenser, turbines have negligible<br />

induced effects, fuel impact is relevant in the boiler and in the preheaters and set points<br />

affect mainly the secondary air preheater.<br />

A decomposition <strong>of</strong> all unit exergy consumptions <strong>of</strong> the boiler is plotted in figure<br />

7.17, where it can be seen how they are affected by all groups <strong>of</strong> variables. Figure 7.18<br />

shows how the unit exergy consumptions associated with the residues vary mainly due<br />

to ambient conditions and induced effects. Finally, Figure 7.19 shows the<br />

decomposition <strong>of</strong> the different components <strong>of</strong> plant product. Although the total<br />

263


Chapter 7<br />

summation only depends on the corresponding set-point, the distribution among the four<br />

turbines is affected by all variables. At this point it should be reminded that the <strong>analysis</strong><br />

developed in this work only considers the full load range <strong>of</strong> the plant operation (> 320<br />

MW); in other situation, this variation would be higher.<br />

Component<br />

Component<br />

264<br />

k ij decomposition<br />

-1.6 -1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0 0.2<br />

Impact<br />

k5_14<br />

k5_13<br />

k5_12<br />

k5_11<br />

k5_10<br />

k5_9<br />

k5_8<br />

k5_7<br />

k5_6<br />

k5_4<br />

k5_3<br />

k5_2<br />

k5_1<br />

Figure 7.15 Decomposition <strong>of</strong> unit exergy consumptions.<br />

k ij decomposition<br />

-0.14 -0.12 -0.1 -0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06<br />

Impact<br />

k5_14<br />

k5_13<br />

k5_12<br />

k5_11<br />

k5_10<br />

k5_9<br />

k5_8<br />

k5_7<br />

k5_6<br />

k5_4<br />

k5_3<br />

Figure 7.16. Decomposition <strong>of</strong> unit exergy consumptions.<br />

k_5<br />

k_5<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced


Component<br />

Component<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

k14_5<br />

k13_5<br />

k12_5<br />

k5_5<br />

k4_5<br />

k3_5<br />

k2_5<br />

k1_5<br />

k0_5<br />

k ij decomposition (boiler)<br />

-0.015 -0.01 -0.005 0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04<br />

Figure 7.17. Decomposition <strong>of</strong> unit exergy consumption associated with the boiler.<br />

theta11_14<br />

theta11_13<br />

theta11_12<br />

theta11_10<br />

theta11_9<br />

theta11_8<br />

theta11_7<br />

theta11_5<br />

theta6_5<br />

Variation<br />

theta ij decomposition<br />

-0.06 -0.04 -0.02 0 0.02 0.04 0.06 0.08<br />

Variation<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

Figure 7.18. Decomposition <strong>of</strong> unit exergy consumptions associated with the residues.<br />

265


Chapter 7<br />

Component<br />

266<br />

Plant product decomposition<br />

E10_0<br />

E9_0<br />

E8_0<br />

E7_0<br />

-6000 -4000 -2000 0 2000 4000 6000 8000<br />

Variation (kW)<br />

Figure 7.19. Decomposition <strong>of</strong> plant product.<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

7.3.2 Decomposition <strong>of</strong> malfunctions.<br />

Figure 7.20 shows the decomposition <strong>of</strong> malfunctions and how the high variation <strong>of</strong><br />

unit exergy consumptions <strong>of</strong> the coil heaters has become a low malfunction, due to the<br />

reduced value <strong>of</strong> the product <strong>of</strong> these components. On the other hand, malfunction <strong>of</strong><br />

the boiler is clearly the highest. Other components with high value <strong>of</strong> the malfunctions<br />

are the preheaters, the low pressure turbine and the condenser. It should be reminded<br />

that the decomposition <strong>of</strong> malfunctions is, by definition, exactly the same as that <strong>of</strong> the<br />

unit exergy consumptions.<br />

Separate <strong>analysis</strong> <strong>of</strong> the different contributions <strong>of</strong> the boiler malfunctions is<br />

performed in Figure 7.21. Besides, Figure 7.22 shows the malfunctions associated with<br />

the residues, caused mainly by ambient conditions and induced by other components.


Component<br />

Component<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

MF14<br />

MF13<br />

MF12<br />

MF11<br />

MF10<br />

MF9<br />

MF8<br />

MF7<br />

MF6<br />

MF5<br />

MF4<br />

MF3<br />

MF2<br />

MF1<br />

Malfunction decomposition<br />

-10000 -5000 0 5000 10000 15000 20000<br />

Figure 7.20. Decompositions <strong>of</strong> malfunctions.<br />

MF14_5<br />

MF13_5<br />

MF12_5<br />

Malfunction (kW)<br />

Malfunction decomposition (boiler)<br />

MF5_5<br />

MF4_5<br />

MF3_5<br />

MF2_5<br />

MF1_5<br />

MF0_5<br />

-10000 -5000 0 5000 10000 15000 20000<br />

Malfunction (kW)<br />

Figure 7.21. Decomposition <strong>of</strong> malfunctions associated with the boiler.<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

267


Chapter 7<br />

268<br />

Component<br />

MR11_14<br />

MR11_13<br />

MR11_12<br />

MR11_10<br />

MR11_9<br />

MR11_8<br />

MR11_7<br />

MR11_5<br />

MR6_5<br />

Malfunction decomposition (residues)<br />

-4000 -2000 0 2000 4000 6000 8000 10000<br />

Malfunction (kW)<br />

Figure 7.22. Decomposition <strong>of</strong> malfunctions associated with the residues.<br />

7.3.3 Decomposition <strong>of</strong> malfunction costs.<br />

Malfunctions presented in the previous sections originate malfunction costs, or<br />

impacts on fuel consumed by the plant. These malfunction costs are summarized in<br />

Figure 7.23, where the impact caused by variation <strong>of</strong> plant product ( MF ) has also been<br />

included. The main malfunction costs are caused by variations in final product, boiler,<br />

secondary air preheater, low pressure turbine and condenser. Effects induced by other<br />

components are low, except in the boiler and in the variation <strong>of</strong> plant product.<br />

Figure 7.24 shows the detailed decomposition <strong>of</strong> malfunction cost associated with<br />

the boiler, and Figure 7.25 plots the decomposition <strong>of</strong> malfunction cost associated with<br />

the residues. Importance <strong>of</strong> induced effects in the last figure justifies the use <strong>of</strong> the<br />

residues <strong>analysis</strong> to reduce induced effects in thermoeconomic <strong>diagnosis</strong>.<br />

*<br />

0<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced


Component<br />

Component<br />

Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

Malfunction cost decomposition<br />

MF*14<br />

MF*13<br />

MF*12<br />

MF*11<br />

MF*10<br />

MF*9<br />

MF*8<br />

MF*7<br />

MF*6<br />

MF*5<br />

MF*4<br />

MF*3<br />

MF*2<br />

MF*1<br />

MF*0<br />

-30000 -20000 -10000 0 10000 20000 30000 40000<br />

Malfunction cost (kW)<br />

Figure 7.23. Decomposition <strong>of</strong> malfunction costs.<br />

Malfunction cost decomposition (boiler)<br />

MF*14_5<br />

MF*13_5<br />

MF*12_5<br />

MF*5_5<br />

MF*4_5<br />

MF*3_5<br />

MF*2_5<br />

MF*1_5<br />

MF*0_5<br />

-25000 -20000 -15000 -10000 -5000 0 5000 10000 15000 20000 25000<br />

Malfunction cost (kW)<br />

Figure 7.24. Decomposition <strong>of</strong> malfunction cost associated with the boiler<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

269


Chapter 7<br />

component<br />

270<br />

Malfunction cost decomposition (residues)<br />

MR*11_14<br />

MR*11_13<br />

MR*11_12<br />

MR*11_10<br />

MR*11_9<br />

MR*11_8<br />

MR*11_7<br />

MR*11_5<br />

MR*6_5<br />

-15000 -10000 -5000 0 5000 10000 15000 20000 25000 30000<br />

Malfunction cost (kW)<br />

Figure 7.25. Decomposition <strong>of</strong> malfunction costs associated with the residues.<br />

ambient<br />

fuel<br />

set points<br />

intrinsic<br />

induced<br />

Results presented up to now show the decomposition <strong>of</strong> malfunction cost into terms<br />

caused by the several groups <strong>of</strong> free <strong>diagnosis</strong> variables (ambient, fuel, set-points,<br />

intrinsic and induced). This has allowed us to determine the relative importance <strong>of</strong> all <strong>of</strong><br />

them and to evaluate the suitability <strong>of</strong> the thermoeconomic model to keep reduced<br />

induced effects. In other words, we have analyzed the MFD and MFI tables ‘row by<br />

row’.<br />

However, to have a complete picture, it is interesting to analyze these tables also ‘by<br />

columns’; that is, to analyze the influence <strong>of</strong> a given free <strong>diagnosis</strong> variable<br />

characterizing the behaviour <strong>of</strong> a component on the malfunction cost <strong>of</strong> that component<br />

and <strong>of</strong> malfunction costs <strong>of</strong> the other components. Productive and residual flows should<br />

be also considered separately. This task is performed in Figures 7.26 and 7.27.<br />

In the case <strong>of</strong> isoentropic efficiencies <strong>of</strong> the turbines, impact assigned to the<br />

corresponding components match fairly well the total impact. In the case <strong>of</strong> flow<br />

coefficients and parameters <strong>of</strong> feeding water pre-heaters this correspondence is not very<br />

good, although influence <strong>of</strong> this variables is low (Figure 7.26). Pressure variation,<br />

uncontrolled water loses and unburned carbon show good behaviour, whereas other


Application <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> to determine intrinsic and induced malfunctions<br />

variables do not. In variables associated with the cooling system, the importance <strong>of</strong><br />

induced effects associated with the residues justifies the separate <strong>analysis</strong> <strong>of</strong> these<br />

flows.<br />

Fuel impact (kW)<br />

Impact (kW)<br />

5000<br />

0<br />

-5000<br />

-10000<br />

-15000<br />

-20000<br />

-25000<br />

10000<br />

5000<br />

0<br />

-5000<br />

-10000<br />

-15000<br />

delta P rec.<br />

Fuel impact <strong>of</strong> free <strong>diagnosis</strong> variables<br />

eff. HP eff. IP1 eff. IP2 eff. LP phi IP1 phi IP2 phi LP TDCA<br />

6<br />

delta P TP<br />

Free <strong>diagnosis</strong> variable<br />

TDCA<br />

5<br />

TDCA<br />

4<br />

TTD 5 TTD 6<br />

Figure 7.26. Fuel impact <strong>of</strong> free <strong>diagnosis</strong> variables (1)<br />

unc. losses<br />

Fuel impact <strong>of</strong> free <strong>diagnosis</strong> variables<br />

eff. cond.<br />

m. cooling<br />

eff. tower<br />

eff. SAH<br />

eff. PAH<br />

Free <strong>diagnosis</strong> variable<br />

eff. boiler<br />

infiltration<br />

unburned<br />

Figure 7.27. Fuel impact <strong>of</strong> free <strong>diagnosis</strong> variables (2).<br />

induced residues<br />

induced productive<br />

intrinsic residues<br />

intrinsic productive<br />

induced residues<br />

induced productive<br />

intrinsic residues<br />

intrinsic productive<br />

271


Chapter 7<br />

7.4 Conclusion<br />

In this chapter, theory developed in chapter 4 which connects quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong> and thermoeconomic <strong>diagnosis</strong> has been successfully applied to the case <strong>of</strong><br />

work.<br />

Besides the demonstration <strong>of</strong> the practical applicability <strong>of</strong> the approach, results<br />

provide valuable information on the suitability <strong>of</strong> the used productive structure to<br />

successfully reduce induced effects, by precisely quantifying the influence <strong>of</strong> ambient<br />

conditions, fuel quality, set-points and other components. For example, turbines are<br />

diagnosed correctly, whereas the <strong>analysis</strong> <strong>of</strong> the boiler should be improved.<br />

Furthermore, the separate <strong>analysis</strong> <strong>of</strong> residues flows has proved to reduce induced<br />

effects, so that, to clarify the <strong>diagnosis</strong> results.<br />

In conclusion, the formulation built shows that the thermoeconomic <strong>diagnosis</strong> can be<br />

understood as the highest level <strong>of</strong> <strong>causal</strong>ization in the quantitative <strong>causal</strong>ity <strong>analysis</strong>.<br />

Besides, it opens the way towards an exhaustive evaluation <strong>of</strong> the suitability not only <strong>of</strong><br />

productive structures but also <strong>of</strong> functions to describe flows different from exergy.<br />

272


8 Application <strong>of</strong> linear regression and<br />

neural networks.<br />

Once the suitability <strong>of</strong> quantitative <strong>causal</strong>ity <strong>analysis</strong> for the <strong>diagnosis</strong> <strong>of</strong> a real<br />

example has been demonstrated in chapter 6, the aim <strong>of</strong> this chapter is to analyze the<br />

possibility <strong>of</strong> applying linear regression and neural networks to the same task.<br />

The theoretical development <strong>of</strong> these approaches has been presented in Chapter 4.<br />

The idea is the same: to relate the variations <strong>of</strong> the free <strong>diagnosis</strong> variables to the impact<br />

that they have on the global efficiency indicators. In the quantitative <strong>causal</strong>ity method,<br />

this relation is related by using only a thermodynamic representation <strong>of</strong> the system. The<br />

other two approaches do not use a description <strong>of</strong> the system (they consider it as a black<br />

box); instead they need a large amount <strong>of</strong> operation data to experimentally determine<br />

these relations.<br />

In this chapter, these two approaches are used to determine the relations between<br />

free <strong>diagnosis</strong> variables and their impacts. First, linear regression (linear additive<br />

model) is applied and then neural networks are used. To adjust the linear regression<br />

coefficients and to train the neural networks, the large amount <strong>of</strong> operation data<br />

available is used. Since the temporal evolution <strong>of</strong> variables and impacts has been deeply<br />

presented previously, to validate the approaches and to compare them with the<br />

quantitative <strong>causal</strong>ity method, only relations between variables and impacts are studied.<br />

Results show that the <strong>methods</strong> based on experimental relations are capable to correctly<br />

quantify impacts caused by the most influent variables, but provide inaccurate results in<br />

the case <strong>of</strong> less relevant variables. Calculations have been performed by using Matlab<br />

(MathWorks, 2007; Demuth and Beale, 2002).<br />

273


Chapter 8<br />

8.1 Linear regression<br />

Relation between free <strong>diagnosis</strong> increments and their impacts is, in most cases, quite<br />

linear. So that, it is possible to determine impact factors that relate variations and<br />

impacts. In Chapter 6, these factors have been calculated by adjusting a line to the<br />

points <strong>of</strong> the graph variable increment versus impact obtained by the quantitative<br />

<strong>causal</strong>ity method. The aim <strong>of</strong> this paragraph is to apply lineal regression (linear additive<br />

models) to determine these coefficients directly by using plant data.<br />

A problem appears because regression is applicable under the hypothesis <strong>of</strong><br />

independence <strong>of</strong> the free variables, which is not true in an example like this, in which a<br />

lot <strong>of</strong> variables are present. To solve this question, a variable change is performed in<br />

order to eliminate the most important dependences.<br />

8.1.1 Variable change<br />

The correlation matrix indicates the degree <strong>of</strong> linear correlation between the free<br />

<strong>diagnosis</strong> variables. The <strong>analysis</strong> <strong>of</strong> this matrix shows that there are 12 examples <strong>of</strong> high<br />

correlation (absolute value above 0.6). They are summarized in Table 8.1.<br />

Variable 1 Variable 2 Correlation<br />

Coefficient<br />

2 Relative humidity 1 Ambient temperature -0.7735<br />

4 Coal HHV 5 Carbon mass fraction in coal 0.951<br />

4 Coal HHV 9 Sulphur mass fraction in coal 0.6354<br />

9 Sulphur mass fraction in coal 8 Ash mass fraction in coal 0.6604<br />

10 Nitrogen mass fraction in coal 5 Carbon mass fraction in coal 0.6302<br />

21 Tempering and primary air relation 1 Ambient temperature 0.6374<br />

24 Fraction <strong>of</strong> flue gases through PAH 20 Air for flue gas desulfuration 0.684<br />

36 TDCA 5 th water heater 35 TDCA 6 th water heater 0.7163<br />

37 Pressure drop in reheater 29 Intermediate pressure 1 flow<br />

coefficient<br />

0.6773<br />

38 Pressure increment in turbo-pump 14 Live steam pressure 0.6437<br />

42 Cooling tower effectiveness 1 Ambient temperature 0.8844<br />

42 Cooling tower effectiveness 2 Relative humidity -0.6288<br />

274<br />

Table 8.1. Pairs <strong>of</strong> variables highly correlated.


Application <strong>of</strong> linear regression and neural networks<br />

To avoid this situation, it is possible to make a variable change. The firs step is to<br />

determine experimental correlations linking variable 1 and variable 2. Then, these<br />

correlations are used to calculate theoretical values <strong>of</strong> variables 1 for each point.<br />

Afterwards, variables 1 are substituted by the difference between its actual value minus<br />

the theoretical value calculated by the correlations.<br />

With these new independent variables, it is possible to apply linear regression<br />

without problems. Finally, an inverse variable change should be made in order to obtain<br />

the coefficients corresponding to the initial set <strong>of</strong> variables.<br />

Correlations obtained to link variables 1 and variables 2 are detailed below:<br />

x = 75.04 −0.7494⋅ x<br />

R 2 = 0.5983 8.1<br />

dc,2 d ,1<br />

x = 886.7 + 378.3⋅<br />

x<br />

R 2 = 0.9043 8.2<br />

dc,5 d ,5<br />

x = 1.408 + 0.1480⋅<br />

x<br />

R 2 = 0.4361 8.3<br />

dc,9 d ,8<br />

x =− 0.3324 + 2.329⋅10 ⋅ x R 2 = 0.3971 8.4<br />

−2<br />

dc,10 d ,5<br />

x = 24.22 + 0.6158⋅<br />

x<br />

R 2 = 0.4063 8.5<br />

dc,21 d ,1<br />

x = 23.26 + 0.2508⋅<br />

x<br />

R 2 = 0.4678 8.6<br />

dc,24 d ,20<br />

x = 5.861+ 1.062⋅<br />

x<br />

R 2 = 0.5131 8.7<br />

dc,36 d ,35<br />

x =− 23.91+ 1.940⋅<br />

x<br />

R 2 = 0.4587 8.8<br />

dc,37 d ,29<br />

x =− 30.11+ 1.320⋅<br />

x<br />

R 2 = 0.4144 8.9<br />

dc,38 d ,14<br />

x = 0.3476 + 7.096⋅10 ⋅ x R 2 = 0.7821 8.10<br />

−3<br />

dc,42 d ,1<br />

It should be noted that relations between coal high heating value and coal sulphur<br />

mass fraction, and between cooling tower effectiveness and relative humidity have not<br />

been used. This is due because coal high heating value and cooling tower effectiveness<br />

have been already transformed by filtering their dependence with carbon mass fraction<br />

and ambient temperature, respectively.<br />

275


Chapter 8<br />

8.1.2 Results<br />

Initial variables have been transformed by using the relations obtained in the<br />

previous section; then, linear regression has been applied and finally the inverse <strong>of</strong> the<br />

transformation has been used to obtain the correlation coefficients corresponding to the<br />

original variables. Results are presented in Table 8.2. This table contains the value <strong>of</strong><br />

the impact factors calculated by using linear regression, as well as the error <strong>of</strong> these<br />

impact factors when compared to the average impact factors obtained by the<br />

quantitative <strong>causal</strong>ity method.<br />

276<br />

error ed<br />

error ed<br />

error ed<br />

Where<br />

ed − ed<br />

= 8.11<br />

lr av<br />

_ i, ηb<br />

i, ηb av<br />

edi,<br />

ηb<br />

i,<br />

ηb<br />

ed − ed<br />

= 8.12<br />

lr av<br />

_ iHR , c<br />

iHR , c<br />

av<br />

ediHR<br />

, c<br />

iHR , c<br />

ed − ed<br />

= 8.13<br />

lr av<br />

_ iHR , u<br />

iHR , u<br />

av<br />

ediHR<br />

, u<br />

iHR , u<br />

av<br />

edi is the average impact factor for the variable i obtained by applying the<br />

quantitative <strong>causal</strong>ity method and<br />

lr<br />

ed i is the impact factor calculated by using linear<br />

regression, both <strong>of</strong> them for boiler efficiency, cycle heat rate and unit heat rate. It should<br />

be noted that this error is calculated by using absolute value.<br />

Besides impact factor and its error, the standard deviations <strong>of</strong> the impacts are<br />

included, in order to show which ones are the most relevant variables. These standard<br />

deviations are calculated by using the quantitative <strong>causal</strong>ity method, so that they are the<br />

same as those shown in Table 6.1. The same colour scheme proposed in that table is<br />

conserved: variables whose impact standard deviation is above 20% <strong>of</strong> the maximum<br />

are labelled pink, those from 10 to 20% are labelled orange, those from 5 to 10% are<br />

yellow and variables from 1 to 5% are blue-green. The rest <strong>of</strong> the variables remain<br />

white.


Application <strong>of</strong> linear regression and neural networks<br />

Impact on boiler effficiency Impact on cycle heat rate Impact on unit heat rate<br />

Num. Description Units<br />

Error Stand.<br />

dev.<br />

Impact<br />

Error Stand.de<br />

v.<br />

Impact<br />

factor<br />

factor<br />

0.2338 4.484<br />

4.702<br />

0.1535 5.276<br />

Error Stand. Impact<br />

dev. factor<br />

6.223 0.4017 6.124<br />

1 Ambient temperature ºC 5.868<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

·10 -2<br />

0.3673 9.314<br />

8.055<br />

0.3465 7.844<br />

0.7914 0.01081 7.010<br />

2 Relative humidity % -2.361<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

0.4098 4.117<br />

2.051<br />

8.302 8.039<br />

0.2376 0.1150 2.639<br />

3 Wind speed m/s -5.030<br />

·10 -3<br />

·10 -3<br />

·10 -5<br />

·10 -4<br />

·10 -2<br />

2.367<br />

1.960<br />

-3.572<br />

0.1722 2.618<br />

0.5744 -3.327<br />

1.546<br />

4 Coal high heating value kJ/kg 9.026<br />

·10 -2<br />

·10 -2<br />

·10 -5<br />

·10 -3<br />

·10 -6<br />

·10 -2<br />

·10 -4<br />

0.02687 9.782<br />

5.799<br />

0.5499 2.238<br />

5 Carbon mass fraction in coal % -0.1356 0.1654 0.1898 6.146<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

1.189<br />

2.543<br />

6.894<br />

8.645<br />

6.259<br />

0.3102 5.301<br />

6 Hydrogen mass fraction in coal % -1.775 1.153<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -2<br />

·10 -3<br />

·10 -2<br />

1.257<br />

5.671<br />

7.317<br />

0.8467 9.181<br />

0.3249 8.647<br />

7 Moisture mass fraction in coal % -0.2029 1.097<br />

·10 -2<br />

·10 -2<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

5.657<br />

1.874<br />

3.124<br />

0.9526 7.228<br />

0.1695 0.1339 1.857<br />

8 Ash mass fraction in coal % -8.494<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

0.3624 1.134<br />

1.732<br />

0.8729 2.791<br />

8.544<br />

0.1759 2.169<br />

9 Sulphur mass fraction in coal % -4.278<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

·10 -2<br />

5.819 1.040<br />

1.197<br />

16.86 2.221<br />

6.670<br />

1.186 2.109<br />

10 Nitrogen mass fraction in coal % -7.777<br />

·10 -4<br />

·10 -2<br />

·10 -5<br />

·10 -3<br />

·10 -3<br />

·10 -2<br />

Table 8.2.A. Impact factors obtained by linear regression.<br />

277


Chapter 8<br />

Impact on boiler effficiency Impact on cycle heat rate Impact on unit heat rate<br />

Num. Description Units<br />

Error Stand.<br />

dev.<br />

Impact<br />

Error Stand.de<br />

v.<br />

Impact<br />

Error Stand.<br />

dev.<br />

Impact<br />

factor<br />

factor<br />

factor<br />

0.8093 1.633<br />

1.772<br />

1.579 8.518<br />

4.443<br />

0.4481 6.829<br />

11 Energy provided by natural gas % -2.988<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -3<br />

·10 -3<br />

1.090<br />

5.458<br />

-8.968<br />

1.009<br />

9.640<br />

-7.112<br />

17.46 3.566<br />

12 Live steam temperature ºC 3.355<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -4<br />

·10 -3<br />

0.3803 1.310<br />

-3.873<br />

0.2319 1.308<br />

-4.785<br />

8.633 4.750<br />

13 Reheated steam temperature ºC -1.571<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -3<br />

0.1997 1.830<br />

-1.081<br />

1.664<br />

3.847<br />

-1.223<br />

20.82 5.025<br />

14 Live steam pressure bar -6.181<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

0.8529 7.331<br />

1.066<br />

0.1654 3.804<br />

4.520<br />

0.6367 1.843<br />

15 Gross electric power MW 7.159<br />

·10 -4<br />

·10 -5<br />

·10 -3<br />

·10 -4<br />

·10 -2<br />

·10 -4<br />

3.388<br />

1.520<br />

1.330<br />

0.3161 7.474<br />

1.975<br />

% -0.2978 0.1261 6.832<br />

16 Oxygen in flue gases leaving the<br />

·10 -3<br />

·10 -2<br />

·10 -2<br />

·10 -4<br />

·10 -3<br />

·10 -2<br />

boiler<br />

6.830<br />

8.566<br />

1.112<br />

0.5678 2.762<br />

0.1741 6.490<br />

3.519<br />

-2.709<br />

ºC<br />

17 Average cold-side temperature<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

·10 -2<br />

in secondary air preheaters<br />

0.2910 2.082<br />

3.627<br />

0.6911 8.771<br />

1.985<br />

0.2049 5.3412<br />

-8.727<br />

ºC<br />

18 Average cold-side temperature<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -2<br />

·10 -3<br />

in primary air preheaters<br />

4.550<br />

5.789<br />

8.506<br />

0.1223 9.412<br />

-1.1459<br />

1.480 1.232<br />

19 Sootblowing steam flow rate kg/s -7.574<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

0.1665 5.396<br />

5.112<br />

0.5867 2.141<br />

0.1233 0.1403 1.002<br />

-1.398<br />

kg/s<br />

20 Air for flue gas desulfuration<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

unit<br />

Table 8.2.B. Impact factors obtained by linear regression.<br />

278


Application <strong>of</strong> linear regression and neural networks<br />

Impact on boiler effficiency Impact on cycle heat rate Impact on unit heat rate<br />

Num. Description Units<br />

Error Stand.<br />

dev.<br />

Impact<br />

Error Stand.de<br />

v.<br />

Impact<br />

Error Stand.<br />

dev.<br />

Impact<br />

factor<br />

factor<br />

factor<br />

0.2389 3.455<br />

6.151<br />

14.90 2.314<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

0.4996 1.764<br />

5.715<br />

8.327 2.216<br />

21 Tempering and primary air kg/kg -1.670<br />

relation<br />

·10 -2<br />

0.1305 0.1028 3.618<br />

·10 -5<br />

22 Primary air-coal ratio kg/kg -0.2010 0.4202 5.299 9.078<br />

·10 -3<br />

·10 -3<br />

·10 -5<br />

·10 -4<br />

·10 -2<br />

0.3081 5.261<br />

3.657<br />

0.4848 2.365<br />

8.212<br />

3.600 8.549<br />

-6.716<br />

ºC<br />

23 Temperature difference <strong>of</strong> flue<br />

·10 -4<br />

·10 -5<br />

·10 -4<br />

·10 -6<br />

·10 -3<br />

·10 -4<br />

gases entering preheaters<br />

10.12 6.461<br />

5.821<br />

11.36 4.786<br />

9.591<br />

9.942 1.568<br />

-1.118<br />

-<br />

24 Fraction <strong>of</strong> flue gases through<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -2<br />

·10 -2<br />

PAH<br />

0.1018 9.927<br />

-4.920<br />

9.051<br />

8.491<br />

-4.570<br />

1.198 3.567<br />

-3.153<br />

%<br />

25 High pressure turbine<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

isoentropic efficiency<br />

5.793<br />

7.764<br />

-2.546<br />

5.127<br />

8.200<br />

-2.243<br />

0.8888 2.219<br />

-7.782<br />

%<br />

26 Intermediate pressure 1<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -5<br />

isoentropic efficiency<br />

4.327<br />

3.766<br />

-3.299<br />

3.830<br />

5.301<br />

-2.871<br />

0.7077 1.719<br />

-2.489<br />

%<br />

27 Intermediate pressure 2<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

isoentropic efficiency<br />

4.925<br />

3.989<br />

-1.304<br />

4.351<br />

2.828<br />

-1.165<br />

0.8452 1.577<br />

-5.774<br />

%<br />

28 Low pressure isoentropic<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -2<br />

·10 -3<br />

efficiency<br />

0.7393 1.233<br />

1.424<br />

1.506 7.434<br />

-1.678<br />

6.051 5.467<br />

-6.569<br />

kg 1/2 ·m 3/2<br />

29 Intermediate pressure 1 flow<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·s -1 ·bar -1/2<br />

coefficient<br />

Table 8.2.C. Impact factors obtained by linear regression.<br />

279


Chapter 8<br />

Impact on boiler effficiency Impact on cycle heat rate Impact on unit heat rate<br />

Num. Description Units<br />

Error Stand.<br />

dev.<br />

Impact<br />

Error Stand.de<br />

v.<br />

Impact<br />

Error Stand.<br />

dev.<br />

Impact<br />

factor<br />

factor<br />

factor<br />

2.956 4.509<br />

3.610<br />

2.972 3.943<br />

3.160<br />

60.10 1.462<br />

-1.371<br />

kg 1/2 ·m 3/2<br />

30 Intermediate pressure 2 flow<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -2<br />

·s -1 ·bar -1/2<br />

coefficient<br />

43.76 4.230<br />

1.040<br />

74.18 3.695<br />

7.987<br />

930.0 1.278<br />

-5.327<br />

31 Low pressure flow coefficient kg 1/2 ·m 3/2<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -3<br />

·s -1 ·bar -1/2<br />

2.356<br />

2.600<br />

1.061<br />

0.1189 1.616<br />

7.937<br />

8.376 8.663<br />

32 TTD 6 th water heater ºC 2.750<br />

·10 -3<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -3<br />

0.1463 3.040<br />

3.882<br />

2.652<br />

9.141<br />

3.604<br />

5.995 1.057<br />

33 TTD 5 th water heater ºC 9.586<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

0.3537 2.523<br />

1.696<br />

0.2874 2.226<br />

1.422<br />

10.93 7.914<br />

34 TTD 3 rd water heater ºC -3.105<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

1.748 2.932<br />

1.476<br />

0.6409 2.564<br />

7.707<br />

132.7 8.936<br />

35 TDCA 6 th water heater ºC -1.747<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

·10 -5<br />

·10 -3<br />

0.1387 9.612<br />

1.035<br />

0.1654 8.501<br />

8.874<br />

1.372 3.026<br />

36 TDCA 5 th water heater ºC 6.686<br />

·10 -4<br />

·10 -4<br />

·10 -4<br />

·10 -5<br />

·10 -4<br />

·10 -5<br />

0.1955 3.579<br />

4.306<br />

0.1344 3.503<br />

4.534<br />

0.7979 9.723<br />

37 Pressure drop in reheater bar 2.145<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -3<br />

1.165<br />

7.535<br />

3.473<br />

0.2787 1.039<br />

4.261<br />

28.55 5.117<br />

2.589<br />

bar<br />

38 Pressure increment in turbo-<br />

·10 -3<br />

·10 -2<br />

·10 -4<br />

·10 -3<br />

·10 -4<br />

·10 -4<br />

·10 -3<br />

pump<br />

Table 8.2.D. Impact factors obtained by linear regression.<br />

280


Application <strong>of</strong> linear regression and neural networks<br />

Impact on boiler effficiency Impact on cycle heat rate Impact on unit heat rate<br />

Num. Description Units<br />

Error Stand.<br />

dev.<br />

Impact<br />

Error Stand.de<br />

v.<br />

Impact<br />

Error Stand.<br />

dev.<br />

Impact<br />

factor<br />

factor<br />

factor<br />

0.1022 5.756<br />

3.734<br />

0.1506 2.823<br />

-1.419<br />

0.5539 1.870<br />

39 Water losses kg/s 3.485<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

-0.1390 0.3055 1.246<br />

-0.1296 0.2685 1.102<br />

40 Condenser effectiveness - -0.2262 3.941 3.505<br />

·10 -2<br />

·10 -2<br />

·10 -3<br />

0.3546 1.190<br />

-1.391<br />

0.3228 1.050<br />

-1.288<br />

2.640 3.910<br />

41 Cooling water flow rate kg/s -2.096<br />

·10 -2<br />

·10 -5<br />

·10 -2<br />

·10 -5<br />

·10 -3<br />

·10 -5<br />

-0.2671 0.2878 2.206<br />

-0.2477 0.2462 1.946<br />

42 Cooling tower effectiveness - -0.4303 2.580 1.009·<br />

·10 -2<br />

·10 -2<br />

10 -2<br />

-0.2015 0.2296 8.441<br />

11.36 6.703<br />

- 6.961 0.09682 0.2488 1.948<br />

43 Secondary air preheater<br />

·10 -3<br />

·10 -5<br />

·10 -2<br />

effectiveness<br />

-0.1289 0.7586 1.668<br />

34.75 1.403<br />

-1.478<br />

- 3.113 0.4388 4.904<br />

44 Primary air preheater<br />

·10 -3<br />

·10 -5<br />

·10 -2<br />

·10 -2<br />

effectiveness<br />

1.140<br />

-0.6308 6.177<br />

2.070<br />

0.4243 0.1179 3.074<br />

45 Aggregated boiler effectiveness - 23.95 2.443<br />

·10 -2<br />

·10 -3<br />

·10 -3<br />

·10 -2<br />

·10 -2<br />

0.2648 3.869<br />

1.353<br />

0.6955 1.593<br />

7.467<br />

0.2338 9.969<br />

46 Air infiltration in preheaters % -3.424<br />

·10 -3<br />

·10 -3<br />

·10 -4<br />

·10 -5<br />

·10 -2<br />

·10 -2<br />

9.010<br />

2.924<br />

1.665<br />

0.3095 1.775<br />

0.2518 2.197<br />

47 Carbon in ashes % -0.4726 4.507<br />

·10 -3<br />

·10 -2<br />

·10 -2<br />

·10 -4<br />

·10 -4<br />

·10 -2<br />

Table 8.2.E. Impact factors obtained by linear regression.<br />

281


Chapter 8<br />

The most important conclusion obtained from Table 8.2 is that variables with high<br />

impact (high value <strong>of</strong> standard deviation) presents low error while low impact variables<br />

presents a high error. In conclusion, the method is suitable to analyze only major<br />

impacts.<br />

This fact should not be surprising if the concept <strong>of</strong> the method is considered. The<br />

approach is based on the determination <strong>of</strong> sensitivity coefficients which multiply the<br />

independent variables by minimising the error in the approximation <strong>of</strong> a dependent<br />

variable. So that, variables which have high impact are determined precisely, while the<br />

other do not.<br />

This relation between error and standard deviation can be seen more clearly by using<br />

graphs. Figure 8.1 shows the case <strong>of</strong> boiler efficiency, Figure 8.2 corresponds to cycle<br />

heat rate and Figure 8.3 shows results <strong>of</strong> unit heat rate. X axis represents the normalized<br />

value <strong>of</strong> the standard deviations, calculated by dividing each one <strong>of</strong> them into the<br />

maximum one. Errors are represented on axis y, where logarithmic scale has been used<br />

due to the wide range <strong>of</strong> this variable. Pink points represent the errors obtained by using<br />

linear regression combined with variable change to reduce problems induced by nonindependence<br />

<strong>of</strong> variables, while blue points correspond to errors produced when linear<br />

regression is directly used, without applying variable change. Since variable change<br />

only affects a few variables, in most cases points pink and blue are coincident.<br />

In the three graphs, it can be seen how errors are low for the most important<br />

variables. In boiler efficiency, error remains lower than 0.1 for variables whose impact<br />

is above 40% <strong>of</strong> the maximum. In unit and cycle heat rate, error is lower than 0.4 for<br />

variables with impact above 10% <strong>of</strong> the maximum. In all cases, error increases<br />

dramatically in variables with low influence.<br />

These graphs demonstrate that variable change performed to reduce linear<br />

dependence between variables originates a good improvement in the results. It is<br />

expected that the use <strong>of</strong> this technique for more variables (establishing a maximum level<br />

lower than the 60% used here) would lead to better results.<br />

282


Error<br />

Error<br />

1000<br />

100<br />

10<br />

0.1<br />

0.01<br />

Application <strong>of</strong> linear regression and neural networks<br />

Error and impact in boiler efficiency<br />

1<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

100<br />

10<br />

0.1<br />

0.01<br />

0.001<br />

Normalised impact standard deviation<br />

Direct<br />

Variable change<br />

Figure 8.1. Error in the value <strong>of</strong> impact factors and impact standard deviation<br />

in boiler efficiency.<br />

Error and impact in cycle heat rate<br />

1<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Normalzed impact standard deviation<br />

Direct<br />

Variable change<br />

Figure 8.2. Error in the value <strong>of</strong> impact factors and impact standard deviation<br />

in cycle heat rate.<br />

283


Chapter 8<br />

Error<br />

284<br />

100<br />

10<br />

0.1<br />

0.01<br />

0.001<br />

Figure 8.3. Error in the value <strong>of</strong> impact factors and impact standard deviation<br />

in unit heat rate.<br />

8.2 Neural networks<br />

Error and impact in unit heat rate<br />

1<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Normalized impact standard deviation<br />

Direct<br />

Variable change<br />

An approach similar as that one presented in the previous section is applied here, but<br />

using neural networks instead <strong>of</strong> linear regression. The idea is to use a large amount <strong>of</strong><br />

plant data to train a neural network whose inputs are the free <strong>diagnosis</strong> variables and the<br />

output is the global efficiency indicator. Since, in our case <strong>of</strong> work, there are three<br />

global efficiency indicators, three neural networks are needed.<br />

Once the networks are trained, they can be used to calculate the impact caused by<br />

each one <strong>of</strong> the free variables by substracting the value provided by the net when the<br />

corresponding free <strong>diagnosis</strong> variable has the value <strong>of</strong> the actual point, minus the value<br />

provided by the net when this free variable takes the value <strong>of</strong> the reference points.<br />

Values <strong>of</strong> the other variables should be the same for the two points. They can be those<br />

<strong>of</strong> the reference point, those <strong>of</strong> the actual point or an intermediate value.<br />

Three neural networks, each one corresponding to a global efficiency indicator have<br />

been built and trained by using plant information available from the working example.


Application <strong>of</strong> linear regression and neural networks<br />

Afterwards, the capability <strong>of</strong> these networks for the <strong>diagnosis</strong> is analyzed by comparing<br />

the sensitivity for a change in an input shown by them with the results provided by the<br />

quantitative <strong>causal</strong>ity method and linear regression.<br />

8.2.1 Neural network definition and training<br />

As previously said, three networks have been created and trained, one for each<br />

global efficiency indicator. All <strong>of</strong> them have the same architecture.<br />

There are 47 inputs for the free <strong>diagnosis</strong> variables and only one output. The number<br />

<strong>of</strong> neurons in the hidden layer has been fixed in 3, because the form <strong>of</strong> the function<br />

required is not complex. As it has been seen in chapter 6, when impact is plotted against<br />

free <strong>diagnosis</strong> variable increment, a straight or slightly curved shape appears in most<br />

cases. If more neurons were used, results provided by the net would be fluctuating. The<br />

output function selected for these neurons is sigmoid. Each one <strong>of</strong> them has 47 inputs<br />

corresponding to all inputs <strong>of</strong> the net. Outputs <strong>of</strong> these 3 neurons are the inputs <strong>of</strong> the<br />

output neuron. The output function <strong>of</strong> this output neuron is linear.<br />

The three networks have been trained by using 70% <strong>of</strong> the operation points from the<br />

case study already used in chapter 6 and in section 8.1. Evolution <strong>of</strong> the residual during<br />

the training process is plotted in Figures 8.4. 8.5 and 8.6. It should be noted that this<br />

residual corresponds not to the absolute value <strong>of</strong> the objective function but to its<br />

normalized value.<br />

Since the transformation <strong>of</strong> the free <strong>diagnosis</strong> variables has provided good results in<br />

the application <strong>of</strong> linear regression, this technique has also been applied in the neural<br />

networks. So that, the nets have not been trained using the original set <strong>of</strong> variables, but<br />

using the transformed set. Consequently, this transformation should be used as well<br />

when nets are used for simulation.<br />

Finally, Figures 8.7, 8.8 and 8.9 compare the actual and the simulated value for all<br />

the working points available. Pink points correspond to training points and blue points<br />

are the other points, used only for test. Graphs show that the degree <strong>of</strong> accuracy <strong>of</strong> both<br />

point families is the same, so that the nets have not been over-trained.<br />

285


Chapter 8<br />

286<br />

Error<br />

Error<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

10 1<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -6<br />

Performance is 5.47821e-005<br />

0 10 20 30 40 50<br />

100 Epochs<br />

60 70 80 90 100<br />

Figure 8.4. Training process for boiler efficiency neural network.<br />

Performance is 6.11472e-006<br />

0 10 20 30 40 50<br />

100 Epochs<br />

60 70 80 90 100<br />

Figure 8.5. Training process for cycle heat rate neural network.


Calculated boiler efficiency (%)<br />

87<br />

86<br />

85<br />

84<br />

83<br />

82<br />

81<br />

80<br />

Error<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

Application <strong>of</strong> linear regression and neural networks<br />

Performance is 2.05167e-005<br />

0 10 20 30 40 50<br />

100 Epochs<br />

60 70 80 90 100<br />

Figure 8.6. Training process for unit heat rate neural network.<br />

Real and calculated boiler efficiency<br />

79<br />

79 80 81 82 83 84 85 86 87<br />

Real boiler efficiency (%)<br />

Figure 8.7. Comparison <strong>of</strong> real and calculated values <strong>of</strong> boiler efficiency.<br />

Train<br />

Test<br />

287


Chapter 8<br />

Calculated cycle heat rate<br />

Calculated unit heat rate<br />

288<br />

2.7<br />

2.6<br />

2.5<br />

2.4<br />

2.3<br />

2.2<br />

Real and calculated cycle heat rate<br />

2.1<br />

2.1 2.2 2.3 2.4<br />

Real cycle heat rate<br />

2.5 2.6 2.7<br />

3.2<br />

3.1<br />

3<br />

2.9<br />

2.8<br />

2.7<br />

Figure 8.8. Comparison <strong>of</strong> real and calculated cycle heat rate.<br />

Real and calculated unit heat rate<br />

2.6<br />

2.6 2.7 2.8 2.9<br />

Real unit heat rate<br />

3 3.1 3.2<br />

Figure 8.9. Comparison <strong>of</strong> real and calculated unit heat rate.<br />

Train<br />

Test<br />

Train<br />

Test


Application <strong>of</strong> linear regression and neural networks<br />

8.2.2 Results<br />

To evaluate the suitability <strong>of</strong> the neural networks for the <strong>diagnosis</strong> problem,<br />

sensitivity <strong>of</strong> them when the free <strong>diagnosis</strong> variables are varied is considered and<br />

compared to results provided by the quantitative <strong>causal</strong>ity method and by linear<br />

regression.<br />

Since 47 free <strong>diagnosis</strong> variables multiplied times three global efficiency indicators<br />

constitute a too large amount <strong>of</strong> graphs, only a fraction <strong>of</strong> them are considered. For each<br />

global efficiency indicator, the 6 most influent free <strong>diagnosis</strong> variables are analyzed<br />

(Figures 8.10, 8.12 and 8.14). Besides, 6 more variables are studied in order to take into<br />

account the capability <strong>of</strong> the approach to take into account the impact <strong>of</strong> less influent<br />

variables (Figures 8.11, 8.13 and 8.15). They have been chosen according to the<br />

following criteria: two variables whose influence (standard deviation <strong>of</strong> its impact) is<br />

just below 20% <strong>of</strong> the maximum, two below 10% <strong>of</strong> the maximum and two below 5%<br />

<strong>of</strong> the maximum. In the case <strong>of</strong> cycle heat rate, only 4 variables have an influence<br />

higher than 20% <strong>of</strong> the maximum, and only 3 variables have an influence between 10%<br />

and 20% <strong>of</strong> the maximum. So that, in Figure 8.13 are plotted the variable which has an<br />

impact just above 20% <strong>of</strong> the maximum, three variables whose influence is just below<br />

10% <strong>of</strong> the maximum and two variables with influence just below 5%.<br />

Results shown in these graphs demonstrate that neural networks (and also linear<br />

regression) are suitable for the more influent variables, but error increases when<br />

influence <strong>of</strong> variables decreases. It should be noted that divergence increases strongly<br />

for variables with influence around 1%, as it can be expected from figures 8.1 to 8.3.<br />

Besides, although the use <strong>of</strong> variable change entails a good improvement, variables<br />

which have dependence with others present worse results than the others. For example,<br />

ambient temperature in unit heat rate.<br />

Variables such as condenser effectiveness and cooling tower effectiveness and<br />

cooling water flow rate demonstrate the capability <strong>of</strong> neural networks and quantitative<br />

<strong>analysis</strong> quantification <strong>methods</strong> to deal with non-linear behaviours, in opposition to<br />

approaches based on constant impact factors.<br />

Finally, it should be highlighted that, if more neurons had been included in the<br />

networks, oscillating tendency would appear in the variables with the lowest influence.<br />

289


Chapter 8<br />

Impact (%)<br />

Impact (%)<br />

Impact (%)<br />

290<br />

-4000 -3000 -2000<br />

1<br />

0.5<br />

0<br />

-1000-0.5 0<br />

-1<br />

-1.5<br />

-2<br />

-2.5<br />

-3<br />

-3.5<br />

-4<br />

1000 2000<br />

HHV variation (kJ/kg)<br />

0.5<br />

0<br />

-30 -20 -10 0 10 20<br />

-0.5<br />

1<br />

-1<br />

-1.5<br />

Ambient temperature variation (ºC)<br />

1.5<br />

1<br />

0.5<br />

-1 -0.5<br />

0<br />

-0.5<br />

0 0.5 1<br />

-1<br />

-1.5<br />

-2<br />

Hydrogen in coal variation (%)<br />

0<br />

-6 -4 -2 0 2 4 6 8<br />

-0.5<br />

Figure 8.10. Variable increment and impact for the six variables most influent on boiler<br />

efficiency: coal HHV, aggregated boiler effectiveness, ambient temperature, moisture in<br />

coal, hydrogen in coal and carbon in ashes.<br />

Impact (%)<br />

Impact (%)<br />

Impact (%)<br />

-0.1 -0.05<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5 0<br />

-1<br />

-1.5<br />

-2<br />

-2.5<br />

0.05 0.1<br />

Aggregated effectiveness variation<br />

1<br />

0.5<br />

-1<br />

-1.5<br />

Coal moisture variation (%)<br />

-2 -1<br />

0<br />

0<br />

-0.5<br />

1 2 3 4<br />

Unit 1 Unit 2 Unit 3 Linear regression Neural network<br />

1<br />

0.5<br />

-1<br />

-1.5<br />

-2<br />

Carbon in ashes variation (%)


Impact (%)<br />

Impact (%)<br />

Impact (%)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

-30 -20 -10<br />

0<br />

-0.1 0 10 20 30<br />

-0.2<br />

-0.3<br />

-0.4<br />

Tempering and primary air relation variation (%)<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

-30 -20<br />

0<br />

-10-0.05 0 10 20 30 40<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

Average cold side temp. in PAP variation (ºC)<br />

0.06<br />

0.04<br />

0.02<br />

-1<br />

0<br />

-0.5-0.02 0 0.5 1 1.5 2<br />

-0.04<br />

-0.06<br />

-0.08<br />

-0.1<br />

Sulphur in coal variation (%)<br />

Application <strong>of</strong> linear regression and neural networks<br />

-10<br />

0<br />

-0.2<br />

0 10 20 30 40<br />

Figure 8.11. Relation between variable increment and its impact on boiler efficiency for<br />

several variables: tempering and primary air relation, air infiltration in preheaters, average<br />

cold-side temperature in primary air preheaters, primary air-coal ratio, sulphur mass fraction<br />

in coal and low pressure isoentropic efficiency.<br />

Impact (%)<br />

Impact (%)<br />

0.4<br />

0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1<br />

Air infiltration variation (%)<br />

-0.5<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-0.05 0<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

-0.35<br />

-0.4<br />

0.5 1<br />

Primary air-coal ratio variation (kg/kg)<br />

0.15<br />

0.1<br />

0.05<br />

-25 -15<br />

0<br />

-0.05<br />

-5 5 15 25<br />

-0.1<br />

-0.15<br />

-0.2<br />

Low pressure turbine efficiency variation (%)<br />

Unit 1 Unit 2 Unit 3 Linear regression Neural network<br />

Impact (%)<br />

291


Chapter 8<br />

Impact<br />

Impact<br />

Impact<br />

292<br />

0.1<br />

0.05<br />

-30 -20 -10<br />

0<br />

0<br />

-0.05<br />

10 20<br />

-0.1<br />

-0.15<br />

-0.2<br />

Ambient temperature variation (ºC)<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.3 -0.2 -0.1<br />

0<br />

-0.02 0 0.1 0.2<br />

-0.04<br />

Cooling tower effectiveness variation<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-3000 -2000<br />

0<br />

-1000 -0.01 0 1000 2000<br />

-0.02<br />

-0.03<br />

Cooling water flow rate variation (kg/s)<br />

-20 -10<br />

0<br />

-0.05 0 10 20<br />

Figure 8.12. Variable increment and impact for the six variables most influent on cycle heat<br />

rate: ambient temperature, low pressure turbine isoentropic efficiency, cooling tower<br />

effectiveness, condenser effectiveness, cooling water flow rate and high pressure turbine<br />

isoentropic efficiency.<br />

Impact<br />

Impact<br />

Impact<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

Low pressure turbine efficiency variation (%)<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.2 -0.1<br />

0<br />

-0.01 0 0.1 0.2 0.3<br />

-0.02<br />

-0.03<br />

-0.04<br />

Condenser effectiveness variation<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-8 -6 -4<br />

0<br />

-2 -0.01 0 2 4<br />

-0.02<br />

-0.03<br />

High pressure turbine efficiency variation (%)<br />

Unit 1 Unit 2 Unit 3 Linear regression Neural network


Impact<br />

Impact<br />

Impact<br />

0.03<br />

0.02<br />

0.01<br />

-20 -10<br />

0<br />

0<br />

-0.01<br />

10 20 30<br />

-0.02<br />

-0.03<br />

Relative humidity variation (%)<br />

0.02<br />

0.01<br />

0<br />

-4 -2 0 2 4 6 8<br />

-0.01<br />

-0.02<br />

-0.03<br />

IP 2 turbine efficiency variation (%)<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-3000 -2000 -1000 0 1000 2000<br />

-0.005<br />

Coal HHV variation (kJ/kg)<br />

Application <strong>of</strong> linear regression and neural networks<br />

0<br />

-12 -7 -2<br />

-0.01<br />

3 8<br />

Figure 8.13. Relation between variable increment and its impact on cycle heat rate for<br />

several variables: relative humidity, intermediate pressure 1 isoentropic efficiency,<br />

intermediate pressure 2 isoentropic efficiency, gross electric power, coal high heating value<br />

and carbon mass fraction in coal.<br />

Impact<br />

Impact<br />

Impact<br />

0.03<br />

0.02<br />

0.01<br />

-0.02<br />

IP 1 turbine efficiency variation (%)<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

-20<br />

0<br />

-10 -0.005 0 10 20 30<br />

-0.01<br />

-0.015<br />

Gross electric power variation (MW)<br />

0.005<br />

0<br />

-10 -8 -6 -4 -2 0 2 4<br />

-0.005<br />

-0.01<br />

-0.015<br />

Carbon mass fraction variation (%)<br />

Unit 1 Unit 2 Unit 3 Linear regression Neural network<br />

293


Chapter 8<br />

Impact<br />

Impact<br />

Impact<br />

294<br />

0.3<br />

0.2<br />

0.1<br />

-20 -10<br />

0<br />

0<br />

-0.1<br />

10 20<br />

-0.2<br />

-0.3<br />

Low pressure turbine efficiency variation (%)<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-3000 -2000<br />

0<br />

-1000 -0.02 0 1000 2000<br />

-0.04<br />

Coal HHV variation (kJ/kg)<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-6 -4 -2<br />

-0.02<br />

0 2 4 6 8<br />

-0.04<br />

Moisture mass fraction in coal variation (%)<br />

0<br />

-30 -20 -10 0 10 20<br />

-0.05<br />

Figure 8.14. Variable increment and impact for the six variables most influent on unit heat<br />

rate: low pressure turbine isoentropic efficiency, ambient temperature, coal high heating<br />

value, cooling tower effectiveness moisture mass fraction in coal and condenser<br />

effectiveness.<br />

Impact<br />

Impact<br />

Impact<br />

0.1<br />

0.05<br />

-0.1<br />

-0.15<br />

Ambient temperature variation (ºC)<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

-0.3 -0.2 -0.1<br />

0<br />

-0.02 0 0.1 0.2<br />

-0.04<br />

-0.06<br />

Cooling tower effectiveness variation<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-0.2 -0.1 0 0.1 0.2 0.3<br />

-0.02<br />

-0.04<br />

Condenser effectiveness variation<br />

Unit 1 Unit 2 Unit 3 Linear regression Neural network


Impact<br />

Impact<br />

Impact<br />

0.02<br />

0<br />

-10 -8 -6 -4 -2 0 2 4<br />

-0.02<br />

-0.04<br />

-0.06<br />

Carbon in coal variation (%)<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-3 -2<br />

0<br />

-1 0<br />

-0.01<br />

1 2 3 4 5<br />

-0.02<br />

Blowing steam variation (kg/s)<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

-10 -5<br />

0<br />

0<br />

-0.005<br />

5 10 15<br />

-0.01<br />

6 th heater hot side temp. diff. variation (ºC)<br />

Application <strong>of</strong> linear regression and neural networks<br />

-20 -10<br />

0<br />

-0.01 0 10 20 30<br />

Figure 8.15. Relation between variable increment and its impact on unit heat rate for several<br />

variables: carbon mass fraction in coal, relative humidity, sootblowing steam flow rate,<br />

intermediate pressure 3 turbine isoentropic efficiency, hot side temperature difference in 6 th<br />

-0.02<br />

-0.03<br />

water heater and average cold-side temperature in primary air preheaters.<br />

Impact<br />

Impact<br />

Impact<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

Relative humidity variation (%)<br />

0.03<br />

0.02<br />

0.01<br />

-4<br />

0<br />

-2<br />

-0.01<br />

0 2 4 6 8<br />

-0.02<br />

-0.03<br />

-0.04<br />

IP 2 turbine efficiency variation (%)<br />

0.01<br />

0.005<br />

0<br />

-30 -20 -10 0 10 20 30 40<br />

-0.005<br />

-0.01<br />

Average cold-side temp. in PAP variation (ºC)<br />

Unit 1 Unit 2 Unit 3 Linear regression Neural network<br />

295


Chapter 8<br />

8.3 Conclusion<br />

In this chapter, two <strong>diagnosis</strong> <strong>methods</strong> based respectively on linear regression and<br />

neural networks have been applied to the working example in order to test their<br />

suitability to the <strong>diagnosis</strong> <strong>of</strong> an actual example.<br />

These two <strong>methods</strong> are based on a purely experimental determination <strong>of</strong> relations<br />

between free <strong>diagnosis</strong> variables and global efficiency indicators, using a large amount<br />

<strong>of</strong> operation data but without introducing any information about the system under study.<br />

It has been seen that the two approaches provide good result for the more influent<br />

variables, but error increases substantially for low influence variables. It should be<br />

highlighted that the interest <strong>of</strong> a variable not only depends <strong>of</strong> its influence but also <strong>of</strong> its<br />

easiness <strong>of</strong> correction. For example, a set-point may have a small impact, but it is<br />

interesting to quantify it in order to advise plant operators to take care about it. This<br />

capability to determine the impact <strong>of</strong> all variables is an important advantage <strong>of</strong> the<br />

quantitative <strong>causal</strong>ity method.<br />

The problem <strong>of</strong> variable independence is a key issue. It has been demonstrated how<br />

the variable change helps to improve results when some degree <strong>of</strong> lineal dependence<br />

among free <strong>diagnosis</strong> variables is present. If the level for considering two variables as<br />

not independent had been fixed below 60%, accuracy <strong>of</strong> results for variables with low<br />

influence would have increased, but more work would have been needed to determine<br />

more correlations. So that, equilibrium should be reach in this point.<br />

Although linear regression and the use <strong>of</strong> neural networks have several points in<br />

common, they are differenced by the capability <strong>of</strong> the second <strong>of</strong> them to deal with non<br />

linear behaviour. This advantage is irrelevant in most variables, but can be decisive in<br />

others. It should be noted that this fact is due because linear additive models are used in<br />

linear regression. It might be possible to develop more complex models based on linear<br />

regression but considering other functions (eg. 2 nd order polynomials), which would be<br />

able to deal with non-linear behaviour but would also entail more complexity <strong>of</strong><br />

calculation and interpretation).<br />

Finally, it should be noted that here it has been compared the capability <strong>of</strong> the three<br />

approaches for the direct <strong>diagnosis</strong> <strong>of</strong> all the system considered. Probably, the best<br />

solution is to apply hybrid configurations based on a general <strong>diagnosis</strong> structure based<br />

on the quantitative <strong>causal</strong>ity <strong>analysis</strong>, where linear regression and/or neural networks<br />

are included for diagnose specific parts. Causality chain theory provides the<br />

296


Application <strong>of</strong> linear regression and neural networks<br />

methodological structure to build this kind <strong>of</strong> complex <strong>diagnosis</strong> systems. A first step<br />

towards this solution has been made in point 6.6, where linear regression has been<br />

applied to determine the relation between cooling tower effectiveness and ambient<br />

temperature.<br />

297


Chapter 8<br />

298


9.1 Synthesis<br />

9 Conclusion<br />

Chapter 2 is a review <strong>of</strong> several <strong>methods</strong> and systems for <strong>diagnosis</strong> <strong>of</strong> power plants.<br />

It starts with a short exposition <strong>of</strong> thermoeconomic concepts and tools; especially<br />

malfunction, malfunction cost and the fuel impact formula. Afterwards, proposals <strong>of</strong><br />

several authors to overcome the problem <strong>of</strong> induced effects (the main drawback <strong>of</strong> the<br />

application <strong>of</strong> thermoeconomic <strong>analysis</strong> to <strong>diagnosis</strong>) are presented. Besides, other<br />

authors have developed <strong>diagnosis</strong> methodologies which avoid the use <strong>of</strong><br />

Thermoeconomics and are based only on thermodynamic representations <strong>of</strong> the<br />

systems. Among these proposals, the <strong>diagnosis</strong> algorithm <strong>of</strong> Correas (2001) is<br />

explained with more detail, because it is the basis <strong>of</strong> the <strong>diagnosis</strong> method developed in<br />

the next chapter. Finally, some systems for plant monitoring and <strong>diagnosis</strong> are<br />

presented.<br />

The quantitative <strong>causal</strong>ity <strong>analysis</strong> is developed in Chapter 3. First <strong>of</strong> all, the<br />

<strong>diagnosis</strong> methodology and the nomenclature are presented. Afterwards, a method to<br />

determine the influence <strong>of</strong> measurement errors on the <strong>diagnosis</strong> results is explained.<br />

Since the methodology is based on the linearization <strong>of</strong> equations, the effect <strong>of</strong> nonlinear<br />

behaviour is also analysed. The aim <strong>of</strong> the method is to compare two situations in<br />

order to share the variation <strong>of</strong> a global efficiency indicator <strong>of</strong> the thermal system into a<br />

summation <strong>of</strong> terms each one due to a different free <strong>diagnosis</strong> variable; so that,<br />

guidelines to the choice <strong>of</strong> these free variables are also given. Finally, the fundamentals<br />

<strong>of</strong> the application <strong>of</strong> <strong>causal</strong>ity chains are stated, in order to introduce several sets <strong>of</strong> free<br />

<strong>diagnosis</strong> variables. A simple example is developed in order to clarify the main<br />

concepts and <strong>methods</strong> explained.<br />

299


Chapter 9<br />

In Chapter 4, the quantitative <strong>causal</strong>ity <strong>analysis</strong> is related and compared to other<br />

<strong>diagnosis</strong> <strong>methods</strong>. In the first part, this <strong>analysis</strong> is connected to thermoeconomic<br />

<strong>diagnosis</strong>. The key point is to apply quantitative <strong>causal</strong>ity <strong>analysis</strong> to decompose the<br />

independent variables <strong>of</strong> the thermoeconomic model (unit exergy consumptions and<br />

plant product) into a summation <strong>of</strong> terms corresponding to the free <strong>diagnosis</strong> variables<br />

<strong>of</strong> the thermodynamic model. The combination <strong>of</strong> this approach and the concept <strong>of</strong><br />

malfunction cost and the fuel impact formula allows to perform a double decomposition<br />

<strong>of</strong> the fuel impact: by components <strong>of</strong> the thermoeconomic model and by free <strong>diagnosis</strong><br />

variables. This double summation leads to the table <strong>of</strong> intrinsic and induced<br />

malfunctions (MFI) and the table <strong>of</strong> malfunctions induced by the free <strong>diagnosis</strong><br />

variables (MFD).<br />

The second part is focused on the application <strong>of</strong> <strong>diagnosis</strong> <strong>methods</strong> based on linear<br />

regression and neural networks. The idea is also to look for the connection between the<br />

variation <strong>of</strong> a free <strong>diagnosis</strong> variable and its impact on the global efficiency indicator,<br />

but this search is developed by using only experimental results and without any model<br />

<strong>of</strong> the system, neither thermodynamic nor thermoeconomic. These empirical techniques<br />

may constitute all the <strong>diagnosis</strong> method, or can be used to simulate a part <strong>of</strong> the system<br />

and integrated with the quantitative <strong>causal</strong>ity <strong>analysis</strong>, thus constituting a hybrid system.<br />

The test bench to the methodologies presented in the previous chapters is presented<br />

in Chapter 5. This is a conventional pulverized coal-fired power plant, composed by<br />

three units <strong>of</strong> 350 MW. Along with the plant description, the main problems affecting<br />

their components are reviewed, in order to define suitable free <strong>diagnosis</strong> variables.<br />

Although all the plant is considered, attention is focused on the steam cycle, while the<br />

boiler is analyzed with low detail. 47 free <strong>diagnosis</strong> variables have been defined,<br />

including ambient conditions, parameters <strong>of</strong> fuel quality, set-points and indicators <strong>of</strong><br />

components behaviour. This chapter constitutes a demonstration <strong>of</strong> the methodology to<br />

be followed to define a <strong>diagnosis</strong> system, starting from the physical description <strong>of</strong> the<br />

power plant to be diagnosed and the identification <strong>of</strong> possible malfunctions, up to the<br />

definition <strong>of</strong> free <strong>diagnosis</strong> variables and global efficiency indicators.<br />

In Chapter 6, the anamnesis, or the repeated <strong>diagnosis</strong>, <strong>of</strong> the three units <strong>of</strong> the<br />

working example during 6 years is presented. The evolution <strong>of</strong> all free <strong>diagnosis</strong><br />

variables and their impact, as well as relation between free variables and impacts is<br />

systematically analyzed in order to demonstrate the capability <strong>of</strong> the approach to deal<br />

with a real system. Residual term, or the error produced due to the non-linear behaviour,<br />

300


Conclusion<br />

is proved to be low enough. Finally, <strong>causal</strong>ity chains are applied in order to filter the<br />

dependence <strong>of</strong> cooling tower effectiveness with ambient temperature.<br />

The theory developed in Chapter 4 to connect quantitative <strong>causal</strong>ity <strong>analysis</strong> and<br />

thermoeonomic <strong>diagnosis</strong> is applied to the working example in Chapter 7. First, the<br />

chosen productive structure is described. Afterwards, the method is applied to a real<br />

example <strong>of</strong> <strong>diagnosis</strong>. Finally, a fictitious general example which considers an average<br />

variation <strong>of</strong> all the free <strong>diagnosis</strong> is developed in order to test the suitability <strong>of</strong> the<br />

productive structure to diagnose the system. Results show how effects induced by<br />

ambient conditions, fuel, set-points and other components are negligible in turbines, but<br />

not so low in the boiler. Besides, the use <strong>of</strong> separate <strong>analysis</strong> for the wastes also allows<br />

to clarify results.<br />

The aim <strong>of</strong> Chapter 8 is to test the suitability <strong>of</strong> linear regression and neural<br />

networks to diagnose the working case, and to compare these empirical approaches and<br />

quantitative <strong>causal</strong>ity <strong>analysis</strong>. Results show that linear regression and neural network<br />

quantify correctly the effects <strong>of</strong> the most influent variables, but fail when dealing with<br />

variables <strong>of</strong> secondary importance. Besides, neural networks and quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong> are able to deal with non-linear behaviour, while linear regression based on<br />

linear additive models does not. Finally, it is convenient to apply variable change to<br />

improve results when interdependence among free <strong>diagnosis</strong> variables appears.<br />

9.2 Contributions<br />

In Chapter 2, a revision <strong>of</strong> the <strong>diagnosis</strong> <strong>methods</strong> for thermal systems has been<br />

made. Thermoeconomics fundamentals and its application to <strong>diagnosis</strong> has been<br />

reviewed, as well as the methodologies developed to overcome the problem <strong>of</strong> induced<br />

malfunctions. Besides, other <strong>methods</strong> with the same purpose but not based on<br />

Thermoeconomics are analyzed. As a result, a general picture on <strong>diagnosis</strong> <strong>of</strong> thermal<br />

systems is obtained, in order to identify the main <strong>diagnosis</strong> tendencies and to serve as a<br />

basis to improve them or to develop new ones.<br />

Starting from an existing <strong>diagnosis</strong> algorithm (Correas, 2001), a complete <strong>diagnosis</strong><br />

methodology is developed in Chapter 3. First <strong>of</strong> all, a new nomenclature is introduced,<br />

which clarify the formulation in order to include new features <strong>of</strong> the <strong>diagnosis</strong> method.<br />

Besides, this nomenclature shows clearly that, the <strong>diagnosis</strong> problem has its own<br />

301


Chapter 9<br />

mathematical entity, but its implementation has a lot <strong>of</strong> common points with other<br />

problems such as simulation, optimization and performance test (Correas, 2001); this<br />

point contributes to reduce the marginal effort needed to develop a <strong>diagnosis</strong> system<br />

from a performance test code.<br />

Equations to relate the deviation in measurements with the variation in <strong>diagnosis</strong><br />

results have been developed. Due to the practical interest <strong>of</strong> the <strong>diagnosis</strong> problem, this<br />

result is very important because it allows to quantify the reliability <strong>of</strong> a <strong>diagnosis</strong> result<br />

depending on the uncertainty <strong>of</strong> available measurements. Besides, this methodology<br />

gives valuable information which can help to improve plant instrumentation depending<br />

on the accuracy required for the <strong>diagnosis</strong>. There is another source <strong>of</strong> error, which is<br />

due to the linealization <strong>of</strong> equations made in the <strong>methods</strong>. Proposals are made in order<br />

to reduce this error, which will prove to be successful in Chapter 6.<br />

The choice <strong>of</strong> the definition <strong>of</strong> free <strong>diagnosis</strong> variables should represent a<br />

equilibrium between the need <strong>of</strong> a widely accepted definition and the need to use<br />

parameters as close as possible to the physical behaviour <strong>of</strong> the components. Guidelines<br />

are listed to choose properly these definitions, which complement the <strong>analysis</strong> <strong>of</strong> the<br />

equipment parameters analyzed for the working example in Chapter 5. For situations<br />

when it is not easy to find a variable with these advisable properties, a theory <strong>of</strong><br />

<strong>causal</strong>ity chains has been developed. This formulation allows to use simultaneously<br />

more than one level <strong>of</strong> free <strong>diagnosis</strong> variables, and to trace quantitatively the path <strong>of</strong><br />

<strong>causal</strong>ity from the root causes, through intermediate indicators up to the global<br />

efficiency indicator <strong>of</strong> the whole plant.<br />

Finally, the idea <strong>of</strong> total fuel impact over a period <strong>of</strong> time has been stated. It extends<br />

the concept <strong>of</strong> <strong>diagnosis</strong> (comparison <strong>of</strong> only two situations) to a defined time span (e.g.<br />

period between plant revisions or between steam blowing), while introducing models <strong>of</strong><br />

the evolution <strong>of</strong> component degradation.<br />

In the first part <strong>of</strong> Chapter 4, the quantitative <strong>causal</strong>ity <strong>analysis</strong> has been connected<br />

with thermoeconomic <strong>diagnosis</strong>. This integration is very important from the conceptual<br />

point <strong>of</strong> view, because thermoeconomic <strong>analysis</strong> can be seen in the highest level <strong>of</strong><br />

<strong>causal</strong>ity in the <strong>analysis</strong> <strong>of</strong> thermal systems, while quantitative <strong>causal</strong>ity <strong>analysis</strong> is able<br />

to determine the true causes <strong>of</strong> the malfunctions pinpointed by thermoeconomic<br />

indicators. It should be noted that the use <strong>of</strong> thermodynamic variables to complement<br />

thermoeconomic <strong>diagnosis</strong> in order to eliminate induced effects has been previously<br />

used by several authors (T<strong>of</strong>folo and Lazzaretto, 2004; Valero et al., 1999). However,<br />

302


Conclusion<br />

the formulation developed here constitutes a systematic approach which takes into<br />

account the crossed influence <strong>of</strong> all variables <strong>of</strong> the thermal system analyzed. On the<br />

other hand, the understanding <strong>of</strong> these relations allows to determine precisely the effects<br />

induced in each components by the influence <strong>of</strong> ambient conditions, fuel quality, setpoints<br />

and other components. This <strong>analysis</strong> is not only interesting for the <strong>diagnosis</strong> <strong>of</strong><br />

particular situations, but represents a systematic tool to asses the suitability <strong>of</strong> a given<br />

productive structure for the <strong>analysis</strong> <strong>of</strong> a thermal system. In other words, it enables to<br />

perform the physical <strong>diagnosis</strong> <strong>of</strong> the thermoeconomic <strong>diagnosis</strong>.<br />

In the second part <strong>of</strong> Chapter 4, the capabilities <strong>of</strong> linear regression and neural<br />

networks for the <strong>diagnosis</strong> problem have been explored. These techniques constitute a<br />

method to link the variation <strong>of</strong> the free <strong>diagnosis</strong> variables and the global efficiency<br />

indicator based on operational experience, without the need <strong>of</strong> building a model. These<br />

relations can be direct, or the system can be decomposed in several blocks. Besides,<br />

these <strong>methods</strong> can be combined with quantitative <strong>causal</strong>ity <strong>analysis</strong> by using hybrid<br />

approaches.<br />

Chapter 5 is an example <strong>of</strong> the methodology to be followed in the specification <strong>of</strong> a<br />

<strong>diagnosis</strong> system. The components <strong>of</strong> the power plant have to be analyzed: possible<br />

malfunction and degradation mechanisms, models and indicators <strong>of</strong> their behaviour.<br />

Measurements available are an important point because they limit the depth <strong>of</strong> the<br />

<strong>diagnosis</strong> achievable. Finally, the level <strong>of</strong> detail required in the <strong>analysis</strong> has to be<br />

considered. With this information, the thermodynamic model can be built and the free<br />

<strong>diagnosis</strong> variables and the global efficiency indicator (or indicators), can be<br />

determined.<br />

Quantitative <strong>causal</strong>ity <strong>analysis</strong> is applied to the <strong>diagnosis</strong> <strong>of</strong> the three units <strong>of</strong> the<br />

example power plant during more than 6 years (Chapter 6). This exhaustive <strong>analysis</strong> <strong>of</strong><br />

a real example during such long time span allows to prove the capability <strong>of</strong> the<br />

approach to quantify the influence <strong>of</strong> seasonal variations, changes in fuel, modification<br />

in operations and component degradation and repairing. It should be noted that all free<br />

<strong>diagnosis</strong> variables considered have been analyzed in order to prove that the method is<br />

able to deal with all <strong>of</strong> them, not only the most influent ones. Besides, graphs<br />

representing the relation <strong>of</strong> variation <strong>of</strong> the free <strong>diagnosis</strong> variables and its impact<br />

demonstrate the coherence <strong>of</strong> the approach. Global indicators have also been proposed<br />

and applied in order to summarize the results.<br />

303


Chapter 9<br />

Impacts on global efficiency indicators are additive, which allows to group them in<br />

families in order to summarize results. This fact enables the use <strong>of</strong> a zooming strategy<br />

which starts with the evolution <strong>of</strong> the global efficiency indicators, continues with<br />

families <strong>of</strong> causes and finalizes with the individual free <strong>diagnosis</strong> variables. However, it<br />

should be noted that this approach is not mandatory for the application <strong>of</strong> the <strong>diagnosis</strong><br />

method; it is only a property <strong>of</strong> it which can facilitate the interpretation <strong>of</strong> results.<br />

Another feature <strong>of</strong> the method which can improve the quality <strong>of</strong> the results is the use <strong>of</strong><br />

<strong>causal</strong>ity chains. This technique has been applied to separate the variation <strong>of</strong> cooling<br />

tower effectiveness into an intrinsic term and other term induced by ambient<br />

temperature.<br />

Finally, errors produced by the method due to linealization have been analyzed for<br />

this large amount <strong>of</strong> real examples. Results prove that this error is very low in a real<br />

application.<br />

The formulation developed in Chapter 4 to connect quantitative <strong>causal</strong>ity <strong>analysis</strong><br />

and thermoeconomic <strong>analysis</strong> is applied to the working example in Chapter 7. Results<br />

show exhaustively the influence <strong>of</strong> the free <strong>diagnosis</strong> variables on the malfuncions, and<br />

demonstrate the capability <strong>of</strong> the approach to precisely quantify the effects intrinsic and<br />

induced by ambient conditions, fuel quality, set-points and other components. The<br />

approach has been applied not only to a <strong>diagnosis</strong> example but also to determine the<br />

general influence <strong>of</strong> all the variables, taking into account their typical deviation. It has<br />

been demonstrated that thermoeconomic <strong>analysis</strong> can be integrated with quantitative<br />

<strong>causal</strong>ity <strong>analysis</strong>, as the highest level <strong>of</strong> <strong>causal</strong>ity.<br />

Finally, linear regression and neural networks have been applied to the <strong>diagnosis</strong> <strong>of</strong><br />

thermal systems in Chapter 8. Advantages and drawbacks <strong>of</strong> these two <strong>methods</strong> and<br />

quantitative <strong>causal</strong>ity <strong>analysis</strong> have been evaluated on the basis <strong>of</strong> the <strong>analysis</strong> <strong>of</strong> a real<br />

system. Results show that pure empirical approaches are able to identify precisely the<br />

influence <strong>of</strong> the main variables, but do not guarantee accuracy for the others.<br />

9.3 Perspectives<br />

In this thesis, the <strong>diagnosis</strong> algorithm has been improved up to the quantitative<br />

<strong>causal</strong>ity <strong>analysis</strong>. This method has been applied to a working example, connected with<br />

thermoeconomic <strong>analysis</strong> and compared with other approaches such as linear regression<br />

304


Conclusion<br />

and neural networks. However, some points remain open and more additional research<br />

lines have appeared.<br />

The idea <strong>of</strong> <strong>causal</strong>ity chains has been formulated but its application has been<br />

reduced to a simple example <strong>of</strong> the cooling tower. This formulation opens the way to<br />

integrate models <strong>of</strong> the components in order to improve the <strong>diagnosis</strong>. Like in the<br />

problem <strong>of</strong> simulation, more exact models lead to better results, but there would be a<br />

point after which it is not worth to look for more accuracy. This point is probably<br />

determined by the error provided by the method and error determined by available<br />

measurements.<br />

Regarding measurements availability and accuracy, it should be noted that the<br />

availability <strong>of</strong> a large amount <strong>of</strong> plant data has been a key point in the development <strong>of</strong><br />

this thesis. However, it should be noted that if information with more quality would be<br />

available (e.g. in a new power plant), point dispersion would be lower. Besides, it would<br />

be interesting to analyze the possibility to perform real-time <strong>diagnosis</strong>. In this situation,<br />

oscillating tendencies (caused, for example, by tanks), would be eliminated by<br />

introducing suitable free <strong>diagnosis</strong> variables (e.g. the levels <strong>of</strong> those tanks). Finally, the<br />

formulation developed to quantify the influence <strong>of</strong> measurement errors on <strong>diagnosis</strong><br />

results can be applied to calculate the <strong>diagnosis</strong> uncertainty and to suggest<br />

improvements in instrumentation.<br />

Up to now, <strong>diagnosis</strong> compares two situations with the final objective <strong>of</strong> taking<br />

decisions on maintenance and repairing. Information on remaining life <strong>of</strong> components<br />

and models <strong>of</strong> degradation would improve the usefulness <strong>of</strong> <strong>diagnosis</strong> results. This idea<br />

could be extended not only to components (long term), but also to optimize operation in<br />

the short term (e.g. steam sootblowing).<br />

A <strong>diagnosis</strong> system needs the development <strong>of</strong> a thermodynamic representation <strong>of</strong> the<br />

system, including mass and energy balances, definitions <strong>of</strong> components effectiveness<br />

and other ancillary equations, as well as the corresponding Jacobian matrix. Automation<br />

<strong>of</strong> this process would reduce substantially the cost <strong>of</strong> development <strong>of</strong> these systems, and<br />

thus, their economic pr<strong>of</strong>itability.<br />

An original formulation has been developed to connect quantitative <strong>causal</strong>ity<br />

<strong>analysis</strong> and thermoeconomic <strong>diagnosis</strong>. This method opens a promising way to<br />

systematically evaluate the suitability <strong>of</strong> different productive structures (and even<br />

magnitudes different from exergy) to the <strong>diagnosis</strong> <strong>of</strong> a given system. In the long term,<br />

305


Chapter 9<br />

this connection perhaps would lead to develop methodologies to build structures able to<br />

reproduce the physical behaviour <strong>of</strong> any thermal system.<br />

Linear regression and neural networks have proved to be able to quantify the<br />

influence <strong>of</strong> the main variables. In order to improve these results, decomposition <strong>of</strong> the<br />

structure in several blocks should be made. The use <strong>of</strong> blocks modelled empirically<br />

would also be introduced in the quantitative <strong>causal</strong>ity <strong>analysis</strong> in order to build hybrid<br />

systems (It should be noted that a first step in this direction has been made in this work<br />

by applying linear regression to build a simple model <strong>of</strong> the cooling tower). With this<br />

point, and with the connection <strong>of</strong> thermoeconomic <strong>analysis</strong>, it can be seen how<br />

quantitative <strong>causal</strong>ity <strong>analysis</strong> can not only be seen as an alternative <strong>diagnosis</strong> method<br />

but also as a global method which can be used to connect others. Finally, neural<br />

networks are very suitable to model complex time-dependent functions, such as<br />

degradation <strong>of</strong> components or evolution <strong>of</strong> fouling, which opens a promising way to<br />

introduce these features in a <strong>diagnosis</strong> system.<br />

306


References<br />

ABB Switzerland Ltd. Power Technology Systems. 2005a. BoilerMax - Boiler Startup<br />

Optimization. www.abb.com/powergeneration<br />

ABB Switzerland Ltd. Power Technology Systems. 2005b. Lifecycle Optimizer. Optimax Asset<br />

Lifecycle Products. www.abb.com/powergeneration<br />

ABB Switzerland Ltd. Power Technology Systems. 2005c. PowerFit - Unit Commitment for<br />

Power Generation and Desalination. Optimax Operation Products.<br />

www.abb.com/powergeneration<br />

ABB Switzerland Ltd. Power Technology Systems. 2005d. Gas Turbine Compressor Wash<br />

Optimizer. Optimax Operation Products. www.abb.com/powergeneration<br />

Alconchel JA, Valero A, Abadía J. 1989. Exergy simulation <strong>of</strong> real operating steam power<br />

plants. Thermoeconomic Analysis <strong>of</strong> Improvement <strong>of</strong> Energy Systems. Eds. C. Ruixian and M.J.<br />

Moran, pp. 485-492. Pergamon Press. Oxford.<br />

Alconchel JA. 1988. Modelización exergética de ciclos de vapour en centrals térmicas. PhD<br />

Thesis, Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> Zaragoza [in Spanish]<br />

Alstom Power - Electrical & Control Systems. 2007. P320 - Power plant flexible automation<br />

system: ALSPA P320. www.power.alstom.com<br />

Álvarez T, Valero A, Montes JM. 2006. Thermoeconomic <strong>analysis</strong> <strong>of</strong> a fuel cell hybrid powe<br />

system from the fuel cell experimental data. Energy 31, pp 1358-1370<br />

Al-Waked R, Behnia M. 2006. CFD simulation <strong>of</strong> wet cooling towers. Applied Thermal<br />

Engineering 26 pp. 382-395.<br />

Antoine M. 2005. Parameter Estimation for Diagnosis and Optimization in Power Plants.<br />

Proceedings <strong>of</strong> PowerGen Europe 2005, Milan, Italy, June 28-30<br />

Arauzo I, Royo J, Valero A, Lacarta M, 1993. Strategies for auxiliary system consumption<br />

optimization in coal powerplants. 10th Annual International Pittsburgh coal Conference. Coal-<br />

Energy and the Environment. Session: Combustion Systems in Powerplants. September 20-24.<br />

Pittsburg, PA USA<br />

Arauzo I, Cortés C, 1995a. Diagnosis <strong>of</strong> Milling Systems Performance Based on Operating<br />

Window and Mill Consumption Correlation. Proceedings <strong>of</strong> the 30th Intersociety Energy<br />

Conversion Engineering Conference, Vol. 2, pp. 397-402. Orlando, Florida, USA, July-August<br />

307


References<br />

Arauzo I, Cortés C, 1995b. Expert System for Boiler Efficiency Deviations Asessment.<br />

Conference papers <strong>of</strong> Power-Gen Europe'95, Penwell, Vol. 6, pp. 183-194. Amsterdam,<br />

Netherlands<br />

Arauzo I, Cortés C, 1997. Diagnosis <strong>of</strong> Electrostatic Precipitator Performance Based on<br />

Operational Data end Empirical Models. Conference proceedings <strong>of</strong> Power-Gen Europe'97,<br />

Penwell, Vol. IV, pp. 323-339. Madrid, Spain<br />

Arranz A, Cruz A, Sanz-Bobi MA, Ruíz P, Coutiño J. 2007. DADICC: Intelligent system for<br />

anomaly detection in a combined cycle gas turbine plant. Expert Systems with Applications (in<br />

Press)<br />

ASME.1964. ASME Performance Test Codes. Steam Generation Units. PTC 4.1. The American<br />

Society <strong>of</strong> Mechanical Engineering<br />

ASME. 1996. Simplified Performance Test <strong>of</strong> Steam Turbines ASME/ANSI PTC-6, Ed. By<br />

ASME, N.Y.<br />

Auxiesa. 1980. Teruel Power Plant Thermal kit.<br />

Barbucci P, Bellagamba S, Benelli G. 2002. Advanced monitoring and <strong>diagnosis</strong> tool to<br />

improve gas turbine operation. Proceedings <strong>of</strong> PowerGen Europe 2002, Milano, Italy, June 11-<br />

13<br />

Bartels F, Simon S, Tippkötter T, Wessel B, Frach F. 2003. Increasing Steam Generator<br />

Performance through Selective on-load Cleaning. Proceedings <strong>of</strong> PowerGen Europe 2003,<br />

Düsseldorf, Germany, May 6-8<br />

Bauch V, Killich M. 2005. Diagnostics for Condition-based maintenance on steam turbines and<br />

generators. Proceedings <strong>of</strong> PowerGen Europe 2005, Milan, Italy, June 28-30<br />

Bella O, Cortés C, Valero A, Tomás A, 1994. Intelligent sootblowing system coal-fired utility<br />

boilers. Conference papers <strong>of</strong> Power-Gen Europe'94. Vol. 8, pp. 355-362, May, Cologne,<br />

Germany<br />

Blokh, AG. 1984. Heat Transfer in Steam Boiler Furnaces. Hemisphere Publishing<br />

Corporation.<br />

Carrasco A, Azcue JM, Sanabrias JI, Giménez A, Cortés A, Díez LI, Iranzo I, Fabiani L. 2001.<br />

On-line Management System for the Advanced Control <strong>of</strong> Utility Boilers Performance. 6th<br />

International Conference on Technologies and Combustion for a Clean Environment. VOL. 2,<br />

pag 921 - 927. Oporto, Portugal, July 9-12<br />

Cooke DH. 1985. On prediction <strong>of</strong> <strong>of</strong>f-design multistage turbine pressures by Stodola's ellipse.<br />

Journal <strong>of</strong> Engineering for Gas Turbines and Power. Transactions <strong>of</strong> the ASME. Vol. 107,<br />

pp.596-606<br />

Copado A, Rodríguez F. 2002. CESAR-SIRE: advanced s<strong>of</strong>ware for boiler efficiency and Nox<br />

optimisation. Fuel 81, pp. 619-626<br />

Correas L, Martínez A, Valero A. 1999. Operation <strong>diagnosis</strong> <strong>of</strong> a combined cycle based on the<br />

structural theory <strong>of</strong> thermoeconomics. ASME-IMECE, Nashville, TE, november 1999<br />

Correas L. 2001. Diagnóstico termoeconómico de la operación de un ciclo combinado. PhD<br />

Thesis, Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> Zaragoza.[in Spanish]<br />

308


References<br />

Correas L. 2004. On the thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy system<br />

malfunctions. Suitability to real-time monitoring. Int.J. Thermodynamics, Vol.7, (No.2), pp.85-<br />

94, June.<br />

Cortés C, Valero A, Tomás A, Abadía J, Arnal N, Torres C. 1989. A system for slagging control<br />

in a coal-fired utility boiler. ASME. AES, Vol. 10-1 Analysis and Design <strong>of</strong> Energy Systems:<br />

Computer-Aided Engineering. Eds, R.A. Bajura, C.H. Marston and G. Tsatsaroins, ASME Book<br />

No. H0537A, pp. 77-84. New York<br />

Cortés C, Bella O, Valero A, Tomás A. 1993. Ash fouling monitoring and sootblowing<br />

optimisation in a pulverised coal fired utility boiler. Proceedings <strong>of</strong> the 10th Annual<br />

International Pittsburgh Coal Conference. Coal-Energy and the Environment. Session:<br />

Combustion Systems. pp. 629-634, September 20-24, Pittsburg, PA USA<br />

Cortés C, Arauzo I, Díez L, Quíle A, González JF. 1998. Improvements in PF boiler<br />

Instrumentation for Advanced Monitoring and Modelling <strong>of</strong> Thermal Performance. Proceedings<br />

<strong>of</strong> Power-Gen Europe'98, Milan. Italiy. June.<br />

Cortés C, Díez L, Campo A. 2001. Modeling Large-Size Boilers as a Set <strong>of</strong> Heat Exchangers:<br />

Tips & Tricks. ASME 2001 Heat Transfer Division, Vol. 4, pp. 41-48. ISBN 0-7918-3561-8.<br />

American Society <strong>of</strong> Mechanical Engineers<br />

Cotton KC and Westcott JC. 1960. Methods for measuring steam turbine-generator<br />

performance, ASME 60-WA-139<br />

Crosa G, Torelli A. 2005. Performance Monitoring <strong>of</strong> Ansaldo Energia V94.3A2 Gas Turbines<br />

by Dynamic Simulator. Proceedings <strong>of</strong> PowerGen Europe 2005, Milan, Italy, June 28-30<br />

Davari B, de Boer R. 2004. Benefits <strong>of</strong> Closed-loop optimal control for combined cycle and<br />

cogeneration power-plants. Proceedings <strong>of</strong> PowerGen Europe 2004, Barcelona, Spain, May 25-<br />

27<br />

Decoussemaeker P, Basler B, Douillard C. 2005. Operational effectiveness and risk<br />

management with remote diagnostic expert support. Proceedings <strong>of</strong> PowerGen Europe 2005,<br />

Milan, Italy, June 28-30<br />

Demuth H, Beale M. 2002. Neural Network Toolbox User's Guide. For use with MATLAB®.<br />

Version 4. The MathWorks, Inc<br />

Díez LI, Cortés C, Arauzo I, González JF. 1998. Advanced Predictive Methods for Improving<br />

Thermal Performance <strong>of</strong> Power plant. Conference papers <strong>of</strong> Latin Power'98, Penwell, 1998, pp.<br />

109-120. Buenos Aires, Argentina. June<br />

Díez LI, Cortés C, Arauzo I, Valero A. 2001. Combustion and Heat Transfer Monitoring in<br />

Large Utility Boilers. Int. J. Therm. Sci. 40, pp. 489-496<br />

Díez L. 2002. Monitorización y simulación en tiempo real de calderas de potencia de carbon<br />

pulverizado. PhD Thesis, Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> Zaragoza. [in<br />

Spanish].<br />

Díez L, Cortés C, Campo A, 2005. Modelling <strong>of</strong> pulverized coal boilers: review and validation<br />

<strong>of</strong> on-line simulation techniques. Applied Thermal Engineering, Vol 25, pp. 1516-1533<br />

Drobnic B, Oman J, Tuma M. 2006. A numerical model for the analyses <strong>of</strong> heat transfer and<br />

leakages in a rotary air preheater. International Journal <strong>of</strong> Heat and Mass Transfer 49, pp.<br />

5001-5009.<br />

309


References<br />

El-Sayed Y. 2007. Fingerprinting the Malfunction <strong>of</strong> Devices. Int. J. <strong>of</strong> Thermodynamics, Vol.<br />

10 (No. 2), pp. 79-85, June.<br />

Erlach B, Serra L, Valero A. 1998. Structural Theory as standard for Thermoeconomics.<br />

ECOS'98. Efficiency, cost, optimization, simulation and environmental aspects <strong>of</strong> energy<br />

systems and processes. Volume I. 8-10 Juillet. Université Henri Poincaré Nancy. Institut<br />

National Polytechnicque de Lorraine. Nancy, France. pp. 229-312.<br />

Frangopoulos CA. 1983. Thermoeconomic functional <strong>analysis</strong>: a method for optimal design or<br />

improvement <strong>of</strong> complex thermal systems. PhD Thesis, Georgia Institute <strong>of</strong> Technology.<br />

Frangopoulos CA. 1986. Thermoeconomic functional <strong>analysis</strong> and optimization. Energy 7 pp.<br />

563-71<br />

Gálvez A, Cuesta D, Meralo E, Chamberlain J, Matesanz J. 2004. Coal Quality Impact<br />

Evaluation: Applications and Benefits. Proceedings <strong>of</strong> PowerGen Europe 2004, Barcelona,<br />

Spain, May 25-27.<br />

García-Peña F, Gálvez A, Correas L, Casero P. 2000. Advanced operation <strong>diagnosis</strong> for power<br />

plants. Performance monitoring and cost assessment in Puertollano IGCC. PowerGen Europe<br />

2000, Helsinki, June.<br />

García-Peña F, Correas L, Millán JL. 2001. Improving O&M at IGCC Puertollano through<br />

thermo-economic <strong>diagnosis</strong>. PowerGen Europe 2001, Brussels, May.<br />

GE Energy. 2005a. System 1 ®. Optimization and Diagnostic Platform.<br />

www.gepower.com/system1<br />

GE Energy. 2005b. Machine Performance. Condition Monitoring and Diagnostics. System 1 ®.<br />

Optimization and Diagnostic Platform. www.gepower.com/system1<br />

Gluch J, Krzyzanowski J, Czwiertnia, Luniewicz B. 1998. New Methods in the Thermal<br />

Diagnosis <strong>of</strong> Power Equipment. Proceedings <strong>of</strong> PowerGen Europe 1998, Milan, Italy, June 9-<br />

11.<br />

Gómez-Yagüe P, Albaladejo JL, Otero P, Copado A. 1999. On-line supervision system <strong>of</strong> unit<br />

energetic efficiency for cost reduction in power stations. Proceedings <strong>of</strong> PowerGen Europe<br />

1999, Frankfurt, Germany, June 1-3.<br />

Grimwade J, Gornm R. 2005. Remote Asset Management <strong>of</strong> Power Generation Plant through<br />

Information Integration. A case study: RWE npower. Proceedings <strong>of</strong> PowerGen Europe 2005,<br />

Milan, Italy, June 28-30.<br />

Guallar J, Valero A. 1988. Thermoeconomic study <strong>of</strong> a sugar factory. ASME. Thermoeconomic<br />

Analysis <strong>of</strong> Chemically Reactive Systems. Eds. W.F. Wepfer, G. Tsatsaronis, R.A. Bajura.<br />

ASME Book No. 600449, pp. 87-94, New York.<br />

Heo G, Chang SH, 2005. Algebraic approach for the <strong>diagnosis</strong> <strong>of</strong> turbine cycles in nuclear<br />

power plants. Nuclear Engineering and Design 235, pp. 1457-1467.<br />

Heo G, Chang SH, Choi SS, Choi GH, Jee MH. 2005. Advisory system for the <strong>diagnosis</strong> <strong>of</strong> lost<br />

electric output in nuclear power plants. Expert Systems with Applications 29, pp 747-756.<br />

Herbert R, Davies D. 2005. Multiple State Vibration Alerts for Turbine Generators.<br />

Proceedings <strong>of</strong> PowerGen Europe 2005, Milan, Italy, June 28-30.<br />

310


Husain. 1984. Steam Turbines-Theory and Design, Mc Graw Hill.<br />

References<br />

Ioanilli E, Gaibazzi I, Maziga G, Zanella GL. 1998. NDT as a tool to optimise the maintenance<br />

planning <strong>of</strong> feedwater heaters in power plants. Proceedings <strong>of</strong> PowerGen Europe 1998, Milan,<br />

Italy, June 9-11.<br />

Iranzo I, Cortés C, Sainz I, Arauzo I, Díez L, 1999. Combustion Assessment Model for an<br />

existing 350 MW Pulverized coal Boiler. Proceedings <strong>of</strong> Power-Gen Europe'99, Penwell,<br />

Frankfurt, Germay, June.<br />

Iranzo I, Domingo E, Cortés C, Arauzo I, 2001. Combustion characterisation <strong>of</strong> a pulverised<br />

coal utility boiler based on CFD techniques. 6th International Conference on Technologies and<br />

Combustion for a Clean Envieronment (Clean Air 2001) Vol. 1, pp. 141 - 177. Oporto, Portugal,<br />

9-12 July.<br />

Jang JSR, Sun CT, Mizutani E. 1997. Neuro-Fuzzy and S<strong>of</strong>t Computing. A Computational<br />

Approach to Learning and Machine Intelligence. Prentice Hall. ISBN 0-13-261066-3.<br />

Jarc CA, Lang R. 1998. On-line, closed-loop boiler optimization to improve boiler efficiency<br />

for four boilers at Chemopetrol's Litvinov plant. Proceedings <strong>of</strong> PowerGen Europe 1998, Milan,<br />

Italy, June 9-11.<br />

Kakaç S, 1991. Boilers, Evaporators and Condensers. John Wiley & Sons Inc.<br />

Kim SM, Joo YJ. 2005. Implementation <strong>of</strong> on-line performance monitoring system at<br />

Seoincheon and Sinincheon combined cycle power plant. Energy (30), pp. 2383-2401.<br />

Klimnstra J. 2004. Power Plant Operation Optimisation through Business Partnership.<br />

Proceedings <strong>of</strong> PowerGen Europe 2004, Barcelona, Spain, May 25-27.<br />

Kotas TJ. 1985. The exergy method for thermal plants <strong>analysis</strong>. Butterworths. London.<br />

Kubiak J, García-Gutiérrez A, Urquiza G. 2002. The <strong>diagnosis</strong> <strong>of</strong> turbine component<br />

degradation-case histories. Applied Thermal Engineering 22, pp. 1955-1963.<br />

Lazaro MA. 2001. Torres de refrigeración atmosféricas. Influencia de las condiciones<br />

ambientales en el rendimiento de un ciclo de vapor. MSc Thesis, Department <strong>of</strong> Mechanical<br />

Engineering, University <strong>of</strong> Zaragoza. [in Spanish].<br />

Lazzaretto A, T<strong>of</strong>folo A, 2006. A critical review <strong>of</strong> the thermoeconomic <strong>diagnosis</strong><br />

methodologies for the location <strong>of</strong> causes <strong>of</strong> malfunctions in energy systems. Journal <strong>of</strong> Energy<br />

Resources Technology. Vol. 128, pp. 335-342.<br />

Lazzaretto A, T<strong>of</strong>folo A, Reini M, Taccani R, Zaleta A, Rangel V, Verda V. 2006. Four<br />

approaches compared on the TADEUS (thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy<br />

utility systems) test case. Energy 31, pp. 1586-1613.<br />

Lerch F, Royo J, Serra L. 1999. Structural theory and thermoeconomic <strong>diagnosis</strong>--Part II:<br />

Application to an Actual Power Plant. Proceedings <strong>of</strong> ECOS 99, pp 374-379. Tokyo, Japan.<br />

López N, Rapún JL, Rodríguez M. 1999. Component diagnostic system for identifying power<br />

loss problems. Proceedings <strong>of</strong> PowerGen Europe 1999, Frankfurt, Germany, June 1-3.<br />

Lozano MA, Valero A. 1987. Application <strong>of</strong> the exergetic costs theory to a steam boiler in a<br />

thermal generating station. ASME. AES, Vol. 3-2 Analysis and Design <strong>of</strong> Advanced Energy<br />

311


References<br />

Systems: Applications. Eds. M.J. Moran, S.S. Stecco and G.M. Reistad, ASME Book No.<br />

G0377B, pp. 41-51. New York.<br />

Lozano MA, Valero A. 1993a. Thermoeconomic <strong>analysis</strong> <strong>of</strong> a gas turbine cogeneration system.<br />

ASME WAM; 1993a.<br />

Lozano MA, Valero A. 1993b. Theory <strong>of</strong> the Exergetic Cost. Energy. Vol. 18, No. 9, pp 939-<br />

960.<br />

Lozano MA, Bartolome JL, Valero A, Reini M. 1994. Thermoeconomic <strong>diagnosis</strong> <strong>of</strong> energy<br />

systems. Proceedings <strong>of</strong> the Third Florence World Eenrgy Research Symposium, Florence,<br />

Italy.<br />

The MathWorks. 2007. Statistics Toolbox User's Guide. For use with MATLAB®. Version 6.<br />

Chapter 8: Regression Analysis. The MathWorks, Inc.<br />

Mendivil P, Álvarez T. 2002. IBERDROLA CENTRE for Monitoring, Diagnostic and<br />

Simulation <strong>of</strong> Combined Cycle Power Plants (CMDS). Proceedings <strong>of</strong> PowerGen Europe 2002,<br />

Milano, Italy, June 11-13.<br />

Mills AF. 1992. Heat Transfer. Irwin.<br />

Montgomery DC, Peck EA. 1992. Introduction to linear regression <strong>analysis</strong>. Wiley<br />

Interscience. ISBN 0-471-53387-4.<br />

Morton S. 2002. Innovative S<strong>of</strong>tware Applications pave the way to improved plant operation.<br />

Proceedings <strong>of</strong> PowerGen Europe 2002, Milano, Italy, June 11-13.<br />

Muñoz M, Valero A, Lozano MA, Alconchel JA, 1986. Simulación del sistema de refrigeración<br />

de una CT eléctrica y su ahorro energético. Energía. Julio-Agosto, pp. 121-127. [in Spanish]<br />

Muñoz M, Valero A. 1989. Thermoeconomic Analysis <strong>of</strong> a cogeneration plant. Thermodynamic<br />

Analysis and Improvement <strong>of</strong> Energy Systems. Eds. C. Ruixian and M.J. Moran, pp. 210-219.<br />

Pergamon Press. Oxford, 1989<br />

Otero P, Gómez P, Albaladejo JL, Menéndez E, Rodríguez-Barea F, Cañadas L. 1999.<br />

Efficiency and Environmental improvement programme in Compostilla Power Station.<br />

Proceedings <strong>of</strong> PowerGen Europe 1999, Frankfurt, Germany, June 1-3.<br />

Pablo Moreno S.A. 2006. Características de SOLCEP frente a otros programas similares.<br />

www.pmsa.retemail.es [in Spanish]<br />

Parenti R, Penno C, Perotti F, Rebizzo A, Ronconi D. 1998. Steam Turbine vibration diagnostic<br />

system based on neural hologram. Proceedings <strong>of</strong> PowerGen Europe 1998, Milan, Italy, June 9-<br />

11.<br />

Rangel VH. 2005. Thermoeconomic <strong>diagnosis</strong> <strong>of</strong> large industrial boilers: Microscopic<br />

representation <strong>of</strong> the exergy cost theory. PhD Thesis. Department <strong>of</strong> Mechanical Engineering.<br />

Unversity <strong>of</strong> Zaragoza.<br />

Ray P, Howard J. 1999. Using ULTRAMAX ® Modeling and Dynamic Optimization s<strong>of</strong>tware<br />

integrated into the APMS SCADA control system <strong>of</strong> unit #2 at Eastern Generation's West<br />

Burton Power Station to reduce Nox and improve overall boiler performance. Proceedings <strong>of</strong><br />

PowerGen Europe 1999, Frankfurt, Germany, June 1-3.<br />

312


References<br />

Rebizzo A, Garbin R, Caruso L. 2002. Remote Monitoring and Diagnostics - The "Global<br />

Approach". Proceedings <strong>of</strong> PowerGen Europe 2002, Milano, Italy, June 11-13.<br />

Rebizzo A, dal Mut G, Oldrati A. 2003. How can mointoring and RCM improve the availability<br />

and performance <strong>of</strong> combined cycle power plants. Proceedings <strong>of</strong> PowerGen Europe 2003,<br />

Düsseldorf, Germany, May 6-8.<br />

Reini M, Taccani R. 2004. On the thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy system<br />

malfunctions. The role <strong>of</strong> the fuel impact formula. Int.J. Thermodynamics, Vol.7, (No.2), pp.61-<br />

72, June.<br />

Reini M. 1994. Analisi e Sviluppo dei Metodi Termoeconomici per lo studio degli Impianti di<br />

Conversione dell'Energia. PhD thesis, Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong><br />

Padova (Italy) [in Italian].<br />

Reismann O, Brunne T, Grusla S, S<strong>of</strong>frion V. 2005. Benefit Analysis <strong>of</strong> Combustion<br />

Optimization with a Neural-Network for Vattenfall's 6x500 MW Jaenschwalde Plant.<br />

Proceedings <strong>of</strong> PowerGen Europe 2005, Milan, Italy, June 28-30.<br />

Remiro JA, Lozano MA. 2007. Control del Rendimiento y Diagnóstico Termoeconómico de<br />

Centrales Termoeléctricas. Información Tecnológica, Vol 18 (1), pp 87-96.<br />

Romeo LM, Gareta R.2006. Hybrid System for fouling control in biomass boilers.Engineering<br />

Applications <strong>of</strong> Artificial Intelligence 19, pp 915-925<br />

Royo J, Valero A. 1995. Towards a unified description <strong>of</strong> the energy behavior <strong>of</strong><br />

thermomechanical systems. ASME. AES, Vol. 35. Thermodynamics and the design, <strong>analysis</strong> and<br />

improvement <strong>of</strong> energy systems. Ed. R.J. Krane, ASME Book No. H01045, pp. 127-134. San<br />

Francisco, California, 1995<br />

Royo J, Valero A, Zaleta A. 1997. The dissipation temperature: a tool for the <strong>analysis</strong> <strong>of</strong><br />

malfunctions in the thermomechanical systems. Energy Convers. Mgmt. Vol. 38, No. 15-17,<br />

pp.1557l-1566.<br />

Rusinowski H, Stanek W, Szapajkco G. 2007. An Advanced System <strong>of</strong> Thermal Diagnosis with<br />

the Application <strong>of</strong> Neural Modeling <strong>of</strong> the Boilers. Proceedings <strong>of</strong> ECOS 2007, Padova, Italy.<br />

Salisbury JK. 1974. Steam Turbines and their cycles. Huntington, N.Y., Robert E. Krieger<br />

Publishing Company.<br />

Schneidel M, Bender K, Taud R. 2004. Use and Experience <strong>of</strong> Gas Turbine On-line<br />

Diagnostics. Proceedings <strong>of</strong> PowerGen Europe 2004, Barcelona, Spain, May 25-27.<br />

Schwarcz P, Lozano MA, Von Spakovsky MR, Valero A. 1997. Diagnostic Analysis <strong>of</strong> a PFBC<br />

Power Plant using a Thermoeconomic Methodology. TAIES'97: Thermodynamic Analysis and<br />

Improvement <strong>of</strong> Energy Systems. Proceedings <strong>of</strong> International Conference. Eds. R. Cai, M.J.<br />

Moran, S. Zhang, Y. Xiao. World Pubs. Corp., pp. 240-249, ISBN 7-5062-3264-Z/TK.6. June<br />

10-13, Beijing, China.<br />

Sciubba E. 2004. Hybrid Semi-Quantitative Monitoring and Diagnostics <strong>of</strong> Energy Conversion<br />

Processes. Int.J. Thermodynamics, Vol. 7 (No.2), pp.95-106, June-2004<br />

Serra L, Lozano A, Valero A and Torres C. 1995. On average and marginal costs in<br />

thermoeconomics. International Symposium ECOS'95. Efficiency, Costs, Optimization,<br />

Simulation and Environmental Impact <strong>of</strong> Energy Systems. Eds. A. Gogus, A. Ozturk and G.<br />

Tsatsaronis, pp. 428-435, ISBN 975-7475-08-4, July 11-15. Istanbul, Turkey.<br />

313


References<br />

Serra L, Torres C. 2006. Structural theory <strong>of</strong> thermoeconomics. EOLSS Publishers, Oxford UK,<br />

2006, www.eolss.net<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006a. SPPA-P3000 Process<br />

Optimization. Combustion optimizer. Increase efficiency, lower emissions. www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006b. SPPA-P3000 Process<br />

Optimization. Sootblower optimizer. Cost-optimized sootblowing. www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006c. SPPA-P3000 Process<br />

Optimization. Fuel Optimizer. Optimum operating despite fluctuating heating values.<br />

www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006d. SPPA-P3000 Process<br />

Optimization. Low Loss Start. Start-up with minimum losses. www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006e. SPPA-P3000. Solutions for<br />

Process Optimization. Siemens Power Plant Automation. www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006f. SPPA-P3000 Process<br />

Optimization. Best Point. Plant performance monitoring and optimization. www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006g. SPPA-P3000 Process<br />

Optimization. Temperature Optimizer. Mobilization <strong>of</strong> temperature reserves.<br />

www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2006h. SPPA-P3000 Process<br />

Optimization. Low Throttling. Reduce throttling losses. www.siemens.de/ic<br />

Siemens AG Power Generation. Instrumentation & Controls. 2007. SPPA-D3000 Reduce costs<br />

by optimizing combustion in gas turbines. www.siemens.com/sppa-d3000<br />

Spalding DB. 1983. Combustión y Transferencia de Masa. Compañía Editorial Continental,<br />

S.A.<br />

Spencer RC, Cotton KC, Cannon CN. 1974. A method for predingting the performance <strong>of</strong> steam<br />

turbine generators 16.500 kW and larger. ASME Power Division Paper No.62-WA-209.<br />

Teruel E, Cortés C, Díez L, Arauzo I, 2005. Monitoring and prediction <strong>of</strong> fouling in coal-fired<br />

utility boilers using neural networks. Chemical Engineering Science, Vol. 60, pp. 5039-5052.<br />

Thulen P, Peper S. 2003. Intelligent Automation Solutions to Turn Power Generation Green.<br />

Proceedings <strong>of</strong> PowerGen Europe 2003, Düsseldorf, Germany, May 6-8.<br />

Tirone G, Arrighi L, Bonifacino L. 1998. Diagnostics based on thermodynamic <strong>analysis</strong> <strong>of</strong><br />

performance <strong>of</strong> steam turbines: case histories. Proceedings <strong>of</strong> PowerGen Europe 1998, Milan,<br />

Italy, June 9-11.<br />

T<strong>of</strong>folo A, Lazzaretto A. 2004. On the thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy<br />

system malfunctions. Indicators to diagnose malfunctions: application <strong>of</strong> a new indicator for the<br />

location <strong>of</strong> causes. Int.J. Thermodynamics, Vol.7, (No2), pp.41-49, June.<br />

T<strong>of</strong>folo A, Lazzaretto A. 2007. Energy System Diagnosis by a Fuzzy Expert System with<br />

Genetically Evolved Rules. Proceedings <strong>of</strong> ECOS 2007, Padova, Italy.<br />

314


References<br />

Torres C, Valero A, Cortés C. 1989. Application <strong>of</strong> symbolic exergoeconomics to thermal<br />

system simulation. ASME. Approaches to the design and optimization <strong>of</strong> thermal systems. Eds.<br />

W.J. Wepfer and M.J. Moran. ASME Book No. H00527, pp. 75-84. New York.<br />

Torres C, Valero A, Serra L, Royo J. 1999. Structural theory and thermoeconomic <strong>diagnosis</strong>--<br />

Part I: on malfunction and dysfunction <strong>analysis</strong>. Proceedings <strong>of</strong> ECOS 99, pp. 368-373. Tokyo,<br />

Japan. 1999.<br />

Torres C, Valero A, Serra L, Royo J. 2002. Structural theory and thermoeconomic <strong>diagnosis</strong>.<br />

Part I: on malfunction and dysfunction <strong>analysis</strong>. Energy Conversion and Management 43,<br />

pp.1503-1518.<br />

Torres C. 2006. Symbolic Thermoeconomic Analysis <strong>of</strong> Energy Systems. EOLSS Publishers,<br />

Oxford UK, 2006, www.eolss.net<br />

Torres C, Valero A, Rangel V, Zaleta A. 2006. On the Cost Formation Process <strong>of</strong> the Residues.<br />

Proceedings <strong>of</strong> the 19th International Conference on Efficiency, Cost, Optimization, Simulation<br />

and Environmental Impact <strong>of</strong> Energy Systems. ECOS06,Aghia Pelagia, Crete, Greece, 12-14<br />

July, Vol. I, pp 415-423<br />

Torres, C.; Valero, A.; Pérez, E; 2007. Guidelines to Develop S<strong>of</strong>tware for Thermoeconomic<br />

Analysis <strong>of</strong> Energy Systems. Part I: The Thermoeconomic Model. Part II: Application to<br />

Thermoeconomic Diagnosis. Proceedings <strong>of</strong> the 20th International Conference on Efficiency,<br />

Cost, Optimization, Simulation and Environmental Impact <strong>of</strong> Energy Systems. ECOS07,Padova,<br />

Italy, June 25-28, Vol. I, pp 435-451<br />

Tozer R, Lozano MA, Valero A, James R. 1996. Thermoeconomics applied to an air<br />

conditioning system with cogeneration. Proceedings <strong>of</strong> the Chartered Institution <strong>of</strong> Building<br />

Services A. Building Services Engineering Research and Tecnology. Vol. 17, No. 1, pp. 37-42.<br />

Tsatsaronis G, Pisa J, Lin L. 1990. The effect <strong>of</strong> assumptions on the detailed exergoeconomic<br />

<strong>analysis</strong> <strong>of</strong> a steam power plant design configuration--Part I: theoretical development. Part II:<br />

results and discussion. Proceedings <strong>of</strong> FLOWERS 90, Florence, May 28--June1.<br />

Uche J. 2001. Guia de usuario de la aplicación Ofi-res. CIRCE. Zaragoza, Spain. [in Spanish].<br />

Uche J, Serra L, Valero A. 2001. Thermoeconomic optimization <strong>of</strong> a dual-purpose power and<br />

desalination plant. Desalination 136, pp. 147-158.<br />

Uche J, Serra L, Valero A. 2006. Exergy Costs and Inefficiency Diagnosis <strong>of</strong> a Dual-Purpose<br />

Power and Desalination Plant. Transactions <strong>of</strong> the ASME, Journal <strong>of</strong> Energy Resources<br />

Technology, September, vol 128. 186-193<br />

Valero A, Lozano MA, Muñoz M. 1986a. A general theory <strong>of</strong> exergy savings--Part I: on the<br />

exergetic cost, Part II: on the thermoeconomic cost, Part III: energy savings and<br />

thermoeconomics; Computer-aided engineering <strong>of</strong> energy system, AES vols. 2-3. ASME WAM.<br />

Valero A, Lozano MA, Alconchel JA, Muñoz MA, Torres C. 1986b. GAUDEAMO: A system<br />

for energetic/exergetic optimization <strong>of</strong> coal power plants. ASME. AES, Vol. 2-1 Computer-Aided<br />

Engineering <strong>of</strong> Energy Systems Vol. 1: Optimization. Ed. R.A. Gaggioli, ASME Book No<br />

H0341 A, pp. 43-49. New York.<br />

Valero A, Guallar J, Muñoz M. 1987. The exergetic cost and related concepts: an application to<br />

a simple cogeneration plant. ASME. Second Law Analysis <strong>of</strong> Thermal Systems. Eds. M.J. Moran<br />

and E.Sciubba, ASME Book No. I00236, pp. 123-130. New York<br />

315


References<br />

Valero A, Torres C. 1988. Algebraic thermodynamic <strong>analysis</strong> <strong>of</strong> energy systems. ASME. AES,<br />

Vol. 7 Approaches to the Design and Optimization <strong>of</strong> Thermal Systems. Eds. E.J. Wepfer and<br />

M.J. Moran, ASME Book No. G00452, pp. 13-23. New York.<br />

Valero A, Torres C, Lozano MA. 1989. On the unification <strong>of</strong> thermoeconomic theories. ASME.<br />

HTD, Vol. 124 Simulation <strong>of</strong> Thermal Energy systems. Eds. R.F. Boehm and Y.M. El-Sayed,<br />

ASME Book No. H00527, pp. 63-74. New York.<br />

Valero A, Carreras A. 1990. On <strong>causal</strong>ity in organized energy systems: Part I: Purpose, cause,<br />

irreversibility, cost. Proceedings <strong>of</strong> FLOWERS 90, Florence, May 28--June 1.<br />

Valero A, Torres C. 1990. On <strong>causal</strong>ity in organized energy systems: Part II: symbolic<br />

exergoeconomics. Proceedings <strong>of</strong> FLOWERS 90, Florence, May 28--June 1.<br />

Valero A, Lozano MA, Torres C. 1990. On <strong>causal</strong>ity in organized energy systems: Part III:<br />

theory <strong>of</strong> perturbations. Proceedings <strong>of</strong> FLOWERS 90, Florence, May 28--June 1.<br />

Valero A, Guallar J. 1991. On processable exergy: A rigorous approach. Athens'91.<br />

International Conference on the Analysis <strong>of</strong> Thermal and Energy Systems. pp. 69-80, June 3-6,<br />

Athens, Greece.<br />

Valero A, Wimmert D, Torres C. 1991. Symbcost: a program for symbolic computation <strong>of</strong><br />

exergoeconomic cost parameters. ASME. AES, Vol. 21 Computer-Aided Energy-Systems<br />

Analysis, Eds. G. Tsatsaronis, R.A. Bajura, W.F. Kenney and G.M. Reistad. ASME Book No.<br />

G00568, pp. 13-21<br />

Valero A, Royo J. 1992. Second law efficiency and the relative free energy function. ASME.<br />

AES, Vol. 27, Thermodynamics and The Design, Analysis, and Improvements <strong>of</strong> Energy<br />

Systems. Eds. R.F. Boehm, ASME Book No. G00717, pp. 271-278, 1992<br />

Valero A, Torres C, Serra L, Lozano MA. 1992. A general theory <strong>of</strong> thermoeconomics--Part I:<br />

structural <strong>analysis</strong>,--Part II: the relative free energy function. Proceedings <strong>of</strong> ECOS’92,<br />

Zaragoza, Spain.<br />

Valero A, Serra L, Lozano MA. 1993. Structural theory <strong>of</strong> thermoeconomics. ASME. AES, Vol.<br />

30. Ed. H.J Richter, ASME Book No. H00874, pp. 189-198.<br />

Valero A. 1995. Thermoeconomics: the meeting point <strong>of</strong> thermodynamics, economics and<br />

ecology. International Conference Second Law Analysis <strong>of</strong> Energy Systems: Towards the 21st<br />

Century. Eds. E. Sciubba, M.J. Moran. Roma "La Sapienza", pp. 293-305, ISBN 88-86662-00-<br />

9.<br />

Valero A, Royo J, Lozano MA. 1995.The characteristic equation and second-law efficiency <strong>of</strong><br />

thermal energy systems. International Conference Second Law Analysis <strong>of</strong> Energy Systems:<br />

Towards the 21st Century. Eds. E. Sciubba, M.J. Moran. Roma "La Sapienza", pp. 99-112,<br />

ISBN 88-86662-00-9,<br />

Valero A, Cortés C. 1996. Ash fouling in coal-fired utility boilers. Monitoring and optimization<br />

<strong>of</strong> on-load cleaning. Progress in Energy and Combustion Science. Vol. 22, pp.189-200.<br />

Valero A, Lozano MA, Bartolomé JL. 1996. On line monitoring <strong>of</strong> power plants performance,<br />

using exergetic cost techniques. Applied Thermal Engineering. Vol. 16, No. 12, pp. 933-948.<br />

Valero A, Correas L, Serra L. 1999. On-line thermoeconomic <strong>diagnosis</strong> <strong>of</strong> thermal power<br />

plants. A. Bejan and E. Mamut (eds.), Thermodynamic Optimization <strong>of</strong> Complex Energy<br />

Systems, 117-136. Kluwer Academic Publishers.<br />

316


References<br />

Valero A, Torres C, Lerch F. 1999. Structural theory and thermoeconomic <strong>diagnosis</strong>, Part III:<br />

Intrinsic and induced malfunctions. Proceedings <strong>of</strong> ECOS'99. Tokyo (Japan): Tokyo Institute <strong>of</strong><br />

Technology.<br />

Valero A, Lerch F, Serra L, Royo J. 2002. Structural theory and thermoeconomic <strong>diagnosis</strong>.<br />

Part II: Application to an actual power plant. Energy Conversion and Management, 43, pp.<br />

1519-1535.<br />

Valero A, Correas L, Zaleta A, Lazzaretto A, Verda V, Reini M, Rangel V. 2004a. On the<br />

thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy system malfunctions. Part 1: the TADEUS<br />

problem. Energy 29 pp 1875-1887<br />

Valero A, Correas L, Zaleta A, Lazzaretto A, Verda V, Reini M, Rangel V. 2004b. On the<br />

thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy system malfunctions. Part 2. Malfunction<br />

definitions and assessment. Energy 29 pp 1889-1907<br />

Valero A, Correas L, Lazzaretto A, Rangel V, Reini M, Taccani R, T<strong>of</strong>folo A, Verda V, Zaleta<br />

A. 2004c. Thermoeconomic Philosopy Applied to the Operating Analysis and Diagnosis <strong>of</strong><br />

Energy Utility Systems. Int.J. Thermodynamics, Vol.7, No.2, pp.33-39, June.<br />

Valero A. 2006a. The thermodynamic process <strong>of</strong> cost formation.Encyclopedia <strong>of</strong> Life Support<br />

systems. EOLSS Publishers, Oxford UK, www.eolss.net<br />

Valero A. 2006b. Exergy accounting: Capabilities and drawbacks. Energy 31 pp 164-180.<br />

Valero A, Torres C. 2006a. Thermoeconomic Analysis. EOLSS Publishers, Oxford UK, 2006,<br />

www.eolss.net<br />

Valero A, Torres C. 2006b. Application <strong>of</strong> thermoeconomics to operation <strong>diagnosis</strong> <strong>of</strong> energy<br />

plants. EOLSS Publishers, Oxford UK, 2006, www.eolss.net<br />

Verda V, Serra L, Valero A. 2002a. Zooming procedure for the thermoeconomic <strong>diagnosis</strong> <strong>of</strong><br />

highly complex energy systems. Int.J. Applied Thermodynamics, Vol.5 (No.2), pp.75-83, June.<br />

Verda V, Serra L, Valero A. 2002b. Effects <strong>of</strong> the productive structure on the results <strong>of</strong> the<br />

thermoeconomic <strong>diagnosis</strong> <strong>of</strong> energy systems. Int.J. Applied Thermodynamics, Vol.5 (No.3),<br />

pp.127-137, September.<br />

Verda V, Valero A, Serra L, Rangel V, Zaleta A, Lazzaretto A, T<strong>of</strong>folo A, Reini M, Taccani R,<br />

Donatini F, Trucato E. 2003. On the thermoeconomic approach to the <strong>diagnosis</strong> <strong>of</strong> energy<br />

system malfunctions. Part 3: Approaches to the Diagnosis problem. Proceedings <strong>of</strong> ECOS 2003,<br />

Copenhagen, Denmark, June 30-July 2.<br />

Verda V. 2004. Thermoeconomic <strong>analysis</strong> and <strong>diagnosis</strong> <strong>of</strong> energy utility systems. From<br />

<strong>diagnosis</strong> to prognosis. Int.J. Thermodynamics, Vol.7, (No.2), pp.73-83, June.<br />

Verda V, Serra L, Valero A. 2004. The effects <strong>of</strong> the control system on the thermoeconomic<br />

<strong>diagnosis</strong> <strong>of</strong> a power plant. Energy 29 pp 331-359.<br />

Verda V, Serra L, Valero A. 2005a. Thermoeconomic Diagnosis: Zooming Strategy Applied to<br />

Highly Complex Energy Systems. Part I: Detection and Location <strong>of</strong> Anomalies. Transactions <strong>of</strong><br />

the ASME.Journal <strong>of</strong> Energy Resources Technology. Vol. 127, pp. 42-58. March.<br />

Verda V, Serra L, Valero A. 2005b. Thermoeconomic Diagnosis: Zooming Strategy Applied to<br />

Highly Complex Energy Systems. Part II: On the Choice <strong>of</strong> the Productive Structure.<br />

317


References<br />

Transactions <strong>of</strong> the ASME.Journal <strong>of</strong> Energy Resources Technology. Vol. 127, pp. 42-58.<br />

March.<br />

Verhage AJL, Jansen HJ, Stevens PMP. 1999. Predictive <strong>diagnosis</strong> for combustion chamber<br />

damage and flame intestabilities in gas turbines. Proceedings <strong>of</strong> PowerGen Europe 1999,<br />

Frankfurt, Germany, June 1-3.<br />

Von Spakovsky MR, Evans RB. 1987. The optimal design and performance <strong>of</strong> thermal system<br />

and their components. Analysis and design <strong>of</strong> advanced energy system fundamentals, AES vol.<br />

3-1. NY: ASME; p. 1-18.<br />

Von Spakovsky MR, Evans RB. 1990. The foundations <strong>of</strong> engineering functional <strong>analysis</strong> (Part<br />

I and II). A future for energy, FLOWERS 1990, Stecco and Moran, Florence.<br />

Walsh P, Fletcher P. 1998. Gas Turbine Performance, ASME Press, Blackwell Science.<br />

Weisenstein W, Ganz C, Stothert A, Morton S. 2002. Data representation and processing for<br />

optimising power plant operation. Proceedings <strong>of</strong> PowerGen Europe 2002, Milano, Italy, June<br />

11-13.<br />

White FM. 1979. Fluid Mechanics. McGraw-Hill Kogakusha.<br />

Zaleta A. 1997. Conceptos sobre el diagnóstico y la evaluación termoeconomica de turbinas de<br />

vapor. PhD thesis, Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> Zaragoza, 1997 [in<br />

Spanish].<br />

Zaleta A, Rangel V, Royo J. 2003. Exergoeconomic fuel-impact <strong>analysis</strong> for steam turbines<br />

sections in power plants. Int.J. Thermodynamics, Vol.6, (No.3), pp.133-141, September.<br />

Zaleta A, Gallegos A, Rangel VH, Valero A. 2004a. A reconciliation method based on a module<br />

simulator. An approach to the <strong>diagnosis</strong> <strong>of</strong> energy system malfunctions. Int.J. Thermodynamics,<br />

Vol.7, (No.2), pp.51-60, June.<br />

Zaleta A, Royo J, Rangel V, Torres E. 2004b. Thermo-characterization <strong>of</strong> power systems<br />

components: a tool to diagnose their malfunctions. Energy 29 pp 361-377.<br />

Zhang C, Chen S, Zheng C, Lou X. 2007. Thermoeconomic <strong>diagnosis</strong> <strong>of</strong> a coal fired power<br />

plant. Energy Conversion and Management 48, pp. 405-419.<br />

318

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!