Publications

A Collection of Scientific Articles Relevant to EnHeGi's Technologies and Engineering Methodologies

Peer-Reviewed Journal Papers

  1. Osara J. A. The Thermodynamics of Degradation. Doctoral Thesis, (The University of Texas at Austin, 2017).*
  2. Osara J. A. and Bryant M. D. Thermodynamics of Lead-Acid Battery Degradation: Application of the Degradation-Entropy Generation Methodology. (Journal of the Electrochemical Society, 2019).
  3. Osara J. A. and Bryant M. D. Thermodynamics of Grease Degradation. (Tribology International, 2019).
  4. Osara J. A. and Bryant M. D. A Thermodynamic Model for Lithium-Ion Battery Degradation: Application of the Degradation-Entropy Generation Theorem. (Inventions, 2019).
  5. Osara J. A. and Bryant M. D. Thermodynamics of Fatigue: Degradation-Entropy Generation Methodology for System and Process Characterization and Failure Analysis. (Entropy, 2019).
  6. Osara J. A. Thermodynamics of Manufacturing Processes: The Workpiece and the Machinery (Inventions, 2019).
  7. Bryant M. D., Khonsari M. M., Ling F. F. On the Thermodynamics of Degradation. (Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 2008).
  8. Bryant M. D. Entropy and Dissipative Processes of Friction and Wear. (FME Transactions, 2009).
  9. Bryant M. D. On Constitutive Relations for Friction from Thermodynamics and Dynamics. (Journal of Tribology, 2016).

 

*Not a Journal paper but included for relevance.

 

Further Reading

  1. Ling F. F., Bryant, M. D., Doelling K. L. On Irreversible Thermodynamics for Wear Prediction. (Wear, 2002).
  2. Doelling K. L., Ling F. F., Bryant M. D. and Heilman B. P. An Experimental Study of the Correlation Between Wear and Entropy Flow in Machinery Components. (Journal of Applied Physics, 2000).

White Papers

ALL ARTICLES ARE EXCLUSIVELY FOR ENHEGI. UNAUTHORIZED COPYING IS PROHIBITED. ALL RIGHTS RESERVED.

<< UPDATED 06/29/19 - THE DEG STORY >>

  • THE DEG STORY

    \!/ CHAPTER 2: THE THEORY OF DEGRADATION OF EVERYTHING

     

    Dr. Jude Osara, PhD PE

    Published on Jan 6, 2019

     

    “I myself know nothing, except just a little, enough to extract an argument from another who is wise and to receive it fairly.” - Socrates

     

    “Jude, you should go home and get some rest”, I heard Prof. Bryant say as I opened my eyes. I had inadvertently dozed off in the middle of his story about renowned inventor Nikola Tesla. The peculiarity of my sleepiness is noteworthy considering it was midday and Prof. Bryant’s story telling is only rivaled by my father’s, in my experience. For me, the day had started about 32 hours ago, culminating in a doctoral dissertation defense. It was over an hour since the official “Congratulations Dr. Osara” from the defense committee, fully exercising their official rights to bestow the title and indicating their satisfaction with the work. Prof. Bryant and I were now having lunch at the famous Posse East and while I can vouch for finishing my meal, I cannot do the same for the Tesla story as it appeared caffeine depletion and postprandial somnolence (or “food coma”) had suddenly brought naptime. Or perhaps one could knowingly observe that I was nearing local minimum energy and maximum entropy state, and re-energizing in the form of sleep was the only natural path forward or degree of freedom, as it were. Indeed, this constraint imposed by nature on all things, natural and artificial, is the crux of this article. While this day’s event is a significant milestone in the DEG story as the first official presentation of the new DEG methodology, its significance is only truly seen in the trajectory that led to this sleepy moment, one that goes back a few years. In Chapter 1, Prof. Bryant tells the genesis. Here, I take you through the sequel.

     

    \!/ CLICK sub-sections below TO READ THE REST OF THE ARTICLE \!/

    • The Meeting

      Shortly after the decision to pursue a PhD degree, I had to choose a PhD research subject. My search for a relevant, yet interesting topic led to a meeting with Prof. Bryant and the start of my involvement in the DEG development. At the end of the first DEG discussion, I was almost certain this would be my PhD research subject, albeit only understanding about 20% of content discussed. The choice turned out much easier than anticipated due to a few key factors:

      • Prof. Bryant seemed 100% certain. “The DEG works”, he would say emphatically with a forward nod that added, “I guarantee it.” His enthusiasm was palpable and infectious.

      • Thermodynamics, I always found fascinating. This indeed would be enjoyable work, I envisaged.

      • Did someone say degradation of everything? Does that mean I get to work on a number of completely different systems? Even less boring!

      Before me was a relevant and interesting doctoral research subject: thermodynamic modeling of real system degradation – formulate and experimentally verify DEG models for two or more systems. The decision was made.

       

      Soon afterwards, another realization hit: I had neither an inkling what the final product would look like nor an idea how I was going to get it done. Fortunately, Profs. Michael Bryant, Michael Khonsari and Frederick Ling had already done significant work in establishing the DEG theorem [1][2], having previously applied the proposed structured approach to friction and wear [2], details in chapter 1; so, there was sufficient material to study. Grease (the lubricant, not the movie) was selected first for experimental demonstration, with others to come later.

    • The 4.5-Year Hiatus...

      ... during which I learned more real-world engineering and partly developed what would become invaluable practical results-oriented, problem-solving intuition. While I was away on full-time employment, hence the hiatus and consequent increase in static inertia (for the Engineering Mechanics enthusiast), Prof. Bryant continued to publish works on system degradation including a chapter in Jonathan Singler's The Physics of Degradation in Engineered Materials and Devices. Without new experiments, he would derive governing thermodynamic relations for various systems and obtain DEG coefficients via direct analogy with previously established system-process characterization models, an approach later inherited in my dissertation study.

    • Impulse – "a force acting briefly on a body and producing a finite change of momentum"

      It was 1pm at Posse East (informal venue of several hundreds of routine science/engineering team lunch meetings annually, including ours). “Jude, we know we can do the work. But just knowing is not enough. We actually have to get it done. Let’s get it done”, Prof. Bryant concluded, as we rose for the stroll back to the department building. While I was away, his interest in batteries had heightened, and with the obvious significance of portable energy, we selected batteries – lithium-ion and lead-acid – for second DEG application. Due to my interest in fatigue analysis and recent work experience, we decided fatigue was next in line, but the potential challenges in characterizing grease and batteries left DEG’s fatigue application yet a distant thought at the time.

       

      Over the next 2.5 years, I would study over 200 book sections and technical papers, perform several grease and battery experiments, process hundreds of severely nonlinear data, consume several Posse East burgers and fries - with carbonated soft drinks for the 'wash down', write a doctoral thesis [1], establish EnHeGi and present DEG to a few potential industry users. Nevertheless, the highlight of this period remains the emergence of a universal real-time system degradation methodology, with variegated experimental verifications.

       

      Now for some scientific narrative.

    • Those That Came Before Us – THERMODYNAMICS

      In addition to presenting a background for new developments, all independent research studies start with a review of past and current works in the specific field. Thermodynamics is a well-established broad field of science that has long been and continues to be used in design and analysis of several engineering systems and processes, from power generation to transportation. This section briefly reviews relevant foundational thermodynamic principles, from equilibrium to non-equilibrium Thermodynamics as they relate to real-world systems and processes.

       

      Classical Thermodynamics [3-5]

      Applying energy conservation to the heat engine, the first law of thermodynamics states that the change in internal energy of a system is the difference between the heat supplied to it and the consequent work obtained from it. In mathematical notation,

       ∆U = Q W .       (1)

      The first law gives the best-case scenario for ALL things, notwithstanding the apparent obscurity of that esoteric deduction from equation (1). (Unfortunately, its demystification will make for an overly long article, hence the interested reader should consult a good thermodynamics textbook - see REFERENCES below). To account for waste heat that never gets converted into work but disorganizes the system, the second law of Thermodynamics introduces entropy as a measure of this disorganization which, if not reversed, results in permanent degradation of the system. Stated mathematically as the Clausius inequality (named after Rudolf Clausius, who first came up with the expression),

       ∆S ≥ (Q / T)b ,       (2)

      a system’s entropy change as it undergoes a process is greater than or equal to the entropy transfer across its boundaries b. This inequality has since been simplified into the more useful

       ∆S = (Q / T)b + S'        (3)

      which is specific and easier to evaluate and interpret: the total entropy change taking place in a closed system as it undergoes a process is the sum of entropy transfer across the system's boundaries and entropy generation within its boundaries. The second law prescribes the real-world scenario. In these two fundamental laws, the capabilities of thermodynamics are deeply rooted. Formulations such as equations (1) and (3) form the basis of several important engineering systems such as combustion – including jet propulsion – engines, steam power plants and air-conditioning/refrigeration systems. However, many everyday systems do not operate on energy conversion between heat and mechanical work, and this may have limited the application of thermodynamics in other areas of engineering, e.g. characterization of electrical/electronic systems.

       

      A first-law analysis tells us the best a system can do so we can decide early if the venture is worth initiating. A second-law analysis tells us the actual expected output from a given combination of input system parameters. Therefore, it is not far-fetched that a combined first and second law analysis would consistently anticipate, and can be used to tune or optimize a system’s performance, both in the design stage and during operation.

       

      A major feature of classical thermodynamics is the frequent use of mathematical manipulations, which may have attracted quite a few theoretical mathematicians and physicists. Thermodynamic models, taking advantage of phenomenological evolution common to all real systems (with very high accuracy when experiments are controlled), are typically of the first order. In all of theoretical thermodynamics, Prof. Herbert Callen's thermodynamic postulates [7] are peradventure the farthest-reaching and most experimentally verified. I find it interesting that he would conduct a simple experiment every now and then, perhaps to prove his formulations to experimental thermodynamicists. The next 3 sub-sections adapt Callen's definitions to everyday systems.

       

      Of States and Paths – Equilibrium and Non-Equilibrium

      (coming soon)

       

       

      Energy Minimum and Entropy Maximum

      (coming soon)

       

       

      Free Energies or Thermodynamic Potentials

       

      "... a battery with an 80% drop in internal energy is more useful (has more electrochemical potential or free energy) in supplying electric charge through direct interaction than a freshly cut diamond, so an internal energy analysis conducted for both components is subject to misinterpretation." – Jude Osara (Thermodynamics of Degradation, 2017)

      (coming soon)

       

       

      Non-Equilibrium Thermodynamics [6-8]

      Often given credit for the advent of modern thermodynamics, Theophile de Donder introduced the first entropy balance for chemical reactions (a more specific version of equation (3)), simplifying the Clausius inequality (equation (2)) and setting the foundation for analytical Irreversible Thermodynamics. Another significant contribution of his work is the chemical affinity which has since established thermodynamic analysis of chemical reactions, and is directly related to a battery's voltage. Prof. de Donder’s former PhD student and future Nobel laureate Ilya Prigogine extended these formulations to non-reacting systems (equation (3)), further establishing and broadening the applicable scope of Irreversible Thermodynamics. Prof. Prigogine focused his research on far from equilibrium interactions which Classical Thermodynamics appeared hitherto unable to characterize. His expansive study of fluctuating and unstable systems led to his discovery of dissipative structures, special systems formed from internal self-reorganizing response to severe energy dissipation. (Our experimental battery measurements recently revealed inadvertent emergence of a dissipative structure; more on this in a future technical paper).

       

      Extending equilibrium thermodynamics to non-equilibrium thermodynamics of macroscopic systems, Prof. Prigogine showed via his minimum entropy production theorem that irreversible entropy produced by a steady state process YdX is simply JdX where J=Y/T is thermodynamic force and dX is thermodynamic flow. This equation is often found in textbooks and research works that apply Thermodynamics to real system modeling. However, Prigogine also listed three basic assumptions for the validity of this expression:

      1. Linear phenomenological laws;

      2. Validity of Onsager’s reciprocity relations;

      3. Phenomenological coefficients may be treated as constants.

      Of particular note are the first and third assumptions requiring that the physical law governing the process be linear and that process coefficients be constant, conditions that readily favor Onsager’s reciprocity, the second assumption. In applying existing scientific laws to real-world applications especially nonlinear processes, it is well known that these assumptions limit validity of principles such as the minimum entropy production theorem to linear transformations and can reduce the accuracy of real system analysis results.

       

      Fluctuations and Instability

        "Chaos gives rise to order." – Ilya Prigogine (Time, Structure and Fluctuations – Nobel Lecture, Dec. 8, 1977)

       (coming soon)

       

    • Those That Are With Us – ENGINEERING Thermodynamics

      While science primarily deals with investigating, establishing and verifying natural laws/theories, engineering takes the next step of applying scientific discoveries to everyday applications. Here, I briefly review common thermodynamic systems, such as internal combustion engines and refrigerators, as well as a few recent and ongoing real-world applications of thermodynamics for system/process characterization and analysis.  Note that the applications listed below are those I have studied and are by no means exhaustive of existing thermodynamic characterizations.

       

      The Heat Engine ICEs, Turbines, Compressors, Heat Exchangers [1214]

      (coming soon)

       

      Prof. Adrian Bejan's Entropy Generation Minimization [1517]

      (coming soon)

       

      Prof. Cemal Basaran's Unified Mechanics Theory [1822]

      (coming soon)

       

      Prof. Erik Kuhn's Grease Rheological Energy Density [2328]

      (coming soon)

       

      Prof. Michael Khonsari et al's Fatigue Entropy [2938]

      (coming soon)

       

      Profs. Leonid A. Sosnovskiy and Sergei S. Sherbakov's Mechanothermodynamics [39–41]

      (coming soon)

       

      Profs. Michael Bryant and Frederick Ling's Friction and Wear Entropy [26]

      (coming soon)

       

    • The Theory Emerges

      “Any sufficiently advanced technology is indistinguishable from magic.” - Sir Arthur C. Clarke

      ONE METHODOLOGY FOR ALL ENGINEERING SYSTEMS AND PROCESSES [42-46].

      The Li-ion Battery's DEG Domain [43]

      \!/ REST OF THE CHAPTER IS coming SOON. \!/

    • REFERENCES

      [1] Osara J. A. The Thermodynamics of Degradation. Doctoral Thesis, UT Austin; 2017.

      [2] Doelling K. L., Ling F. F., Bryant M. D. and Heilman B. P. An Experimental Study of the Correlation Between Wear and Entropy Flow in Machinery Components. Journal of Applied Physics; 2000.

      [3] Bryant M. D., Khonsari M. M., Ling F. F. On the Thermodynamics of Degradation. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences; 2008.

      [4] Ling F. F., Bryant, M. D., Doelling K. L. On Irreversible Thermodynamics for Wear Prediction. Wear; 2002.

      [5] Bryant M. D. Entropy and Dissipative Processes of Friction and Wear. FME Transactions; 2009.

      [6] Bryant M. D. On Constitutive Relations for Friction from Thermodynamics and Dynamics. Journal of Tribology; 2016.

      [7] Burghardt M. D., Harbach J. A. Engineering Thermodynamics. 4th ed. HarperCollins College Publishers; 1993.

      [8] DeHoff R. T. Thermodynamics in Material Science. 2nd ed. CRC Press; 2006.

      [9] Callen H. B. Thermodynamics and an Introduction to Thermostatistics. John Wiley & Sons, Ltd; 1985.

      [10] Prigogine I. Introduction to Thermodynamics of Irreversible Processes. Charles C Thomas; 1955.

      [11] de Groot S. R. Thermodynamics of Irreversible Processes. North-Holland Publishing Company; 1951.

      [12] Kondepudi D., Prigogine I. Modern Thermodynamics: From Heat Engines to Dissipative Structures. John Wiley & Sons Ltd; 1998.

      [13] Moran, M. J., Shapiro, H. N. Fundamentals of Engineering Thermodynamics, 5th ed. Wiley, 2004.

      [14] Cengel Y. A., Boles M. A. Thermodynamics: An Engineering Approach, 8th ed. McGraw Hill; 2015.

      [15] Bejan A. The Method of Entropy Generation Minimization. Energy and the Environment; 1990.

      [16] Bejan, A. Advanced Engineering Thermodynamics. 3rd ed. John Wiley & Sons, Inc; 1997.

      [17] Bejan A. Method of entropy generation minimization, or modeling and optimization based on combined heat transfer and thermodynamics. Revue Générale de Thermique; 1996.

      [18] Gomez, J.; Basaran, C. A thermodynamics based damage mechanics constitutive model for low cycle fatigue analysis of microelectronics solder joints incorporating size effects. Int. J. Solids Struct. 2005.

      [19] Gomez, J.; Basaran, C. Damage mechanics constitutive model for Pb/Sn solder joints incorporating nonlinear kinematic hardening and rate dependent effects using a return mapping integration algorithm. Mech. Mater. 2006.

      [20] Basaran, C.; Lin, M.; Ye, H.Athermodynamic model for electrical current induced damage. Int. J. Solids Struct. 2003.

      [21] Basaran, C.; Nie, S. An Irreversible Thermodynamics Theory for Damage Mechanics of Solids. Int. J. Damage Mech. 2004.

      [22] Basaran, C.; Gomez, J.; Gunel, E.; Li, S. Thermodynamic Theory for Damage Evolution in Solids. In Handbook of Damage Mechanics; Voyiadjis, G., Ed.; Springer: New York, NY, USA, 2014.

      [23] Kuhn E. An energy interpretation of thixotropic effects. Wear; 1991.

      [24] Kuhn E. Description of the Energy Level of Tribologically Stressed Greases. Wear; 1995.

      [25] Kuhn E. and Balan C. Experimental Procedure for the Evaluation of the Friction Energy of Lubricating Greases. Wear; 1997.

      [26] Kuhn E, Investigations into the Degradation of the Structure of Lubricating Greases. Tribology Transactions; 1998.

      [27] Kuhn E. Correlation between System Entropy and Structural Changes in Lubricating Grease. Lubricants; 2015.

      [28] Kuhn E. Friction and Wear of a Grease Lubricated Contact — An Energetic Approach in Tribology - Fundamentals and Advancements, J. Gegner, Ed. Intech; 2013.

      [29] Amiri M. and Khonsari M. M. Life Prediction of Metals Undergoing Fatigue Load Based on Temperature Evolution. Materials Science and Engineering A; 2010.

      [30] Naderi M. and Khonsari M. M. An Experimental Approach to Low-Cycle Fatigue Damage Based on Thermodynamic Entropy. International Journal of Solids and Structures; 2010.

      [31] Naderi M. and Khonsari M. M. A Thermodynamic Approach to Fatigue Damage Accumulation Under Variable Loading. Materials Science and Engineering A; 2010.

      [32] Naderi M. and Khonsari M. M. Real-Time Fatigue Life Monitoring Based on Thermodynamic Entropy. Structural Health Monitoring; 2011.

      [33] Amiri M., Naderi M. and Khonsari M. M. An Experimental Approach to Evaluate the Critical Damage. International Journal of Damage Mechanics; 2011.

      [34] Naderi and Khonsari M. M. A Comprehensive Fatigue Failure Criterion Based on Thermodynamic Approach. Journal of Composite Materials, vol. 46, no. 4; 2012.

      [35] Naderi M. and Khonsari M. M. Thermodynamic Analysis of Fatigue Failure in A Composite Laminate. Mechanics of Materials; 2012.

      [36] Naderi M. and Khonsari M. M. On the Role of Damage Energy in the Fatigue Degradation Characterization Of A Composite Laminate. Composites Part B: Engineering; 2013.

      [37] Amiri M. and Modarres M. An Entropy-Based Damage Characterization. Entropy; 2014.

      [38] Naderi M., Amiri M. and Khonsari M. M. On the Thermodynamic Entropy of Fatigue Fracture, Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences; 2010.

      [39] Sosnovskiy L. A. and Sherbakov S. S. Surprises of TriboFatigue; 2009.

      [40] Sosnovskiy L. A. and Sherbakov, S. S. Mechanothermodynamical System and Its Behavior. Contin. Mech. Thermodyn; 2012.

      [41] Sosnovskiy L. A. and Sherbakov S.S. Mechanothermodynamic Entropy and Analysis of Damage State of Complex Systems. Entropy; 2016.

      [42] Osara J. A. and Bryant M. D. Thermodynamics of Grease Degradation. Tribology International, 2019.

      [43] Osara J. A. and Bryant M. D. A Thermodynamic Model for Lithium-Ion Battery Degradation: Application of the Degradation-Entropy Generation Theorem. Inventions, 2019.

      [44] Osara J. A. and Bryant, M. D. Thermodynamics of Fatigue: Degradation-Entropy Generation Methodology for System and Process Characterization and Failure Analysis. Entropy, 2019.

      [45] Osara J. A. Thermodynamics of Manufacturing Processes: The Workpiece and the Machinery Inventions, 2019.

      [46] Osara J. A. and Bryant M. D. Thermodynamics of Lead-Acid Battery Degradation: Application of the Degradation-Entropy Generation Methodology. Journal of the Electrochemical Society, 2019.

  • \!/ THERMODYNAMICS OF MANUFACTURING PROCESSES – PRODUCT FORMATION

     

    Dr. Jude Osara, PhD PE

    Published on Jan 4, 2019

     

    “... if your theory is found to be against the Second Law of Thermodynamics, I can give you no hope; there is nothing for it but to collapse in deepest humiliation.” - Sir Arthur Eddington

     

    Considered the world’s largest industry, manufacturing transforms billions of raw materials into useful products [1]. Like all real processes and systems, manufacturing processes and equipment are subject to the first and second laws of Thermodynamics and can be modeled via thermodynamic formulations. This article presents a simple thermodynamic model of a manufacturing sub-process or task, assuming multiple tasks make up the entire process. For example, to manufacture a machined component such as an aluminium gear, tasks include cutting the original shaft into gear blanks of desired dimensions, machining the gear teeth, surfacing, etc. The formulations presented here also apply to hand crafting, albeit consistent isolation and measurement of human energy changes due to food intake and work output alone pose a significant challenge; hence this discussion focuses on standardized product-forming processes typically via machine fabrication.

     

    Industrial processes – manufacturing or servicing – involve one or more forms of electrical, mechanical, chemical (including nuclear) or thermal energy conversion processes. For a manufactured component, an interpretation of the first law of Thermodynamics indicates that the internal energy content of the component is the energy that formed the product. Cursorily, this sums all the work that went into the manufacturing process via the electrical, mechanical, chemical and/or thermal power consumption by the manufacturing equipment. However, in accordance with experience, not all of this energy goes into the material-to-product transformation. The heat generation in the machine (most machines run hot enough to require active cooling during operation) also comes from the input power. It is noteworthy that a significant fraction of the heat generation in electromechanical machines is from friction heating (mechanical) and Ohmic heating (electrical), energy dissipating processes.

     

    \!/ click on sub-sections below to read the rest of the article \!/

    • Product in Formation – The Workpiece

      Via the first law, the energy content in a manufactured product

       ∆Up = Qp + Wp         (1)

      where Qp is the net heat transfer between component and surroundings during manufacture and Wp is the net work that transformed the material. Via the second law, the entropy change during manufacture

       ∆Sp = (Q / T)p + S'p         (2)

      where S’p is entropy produced or generated in the component and Tp is component's temperature. Equations (1) and (2) combine to give

       ∆Up = Tp∆Sp – TpS'p + Wp          (3)

      which can be re-arranged to give entropy production or generation

       S'p = (Wp – ∆Ap) / Tp           (4)

      where ∆Ap = ∆Up – Tp∆Sp is the component’s Helmholtz free energy, the minimum work required for product manufacture or maximum work that can be extracted/obtained from the product in service (after manufacture). Actual work Wp includes dissipative phenomena at machine-workpiece interface that do not contribute to the product's desired final form, e.g. using the wrong cutting tool for a machining task will increase Wp for a given ∆Ap, and in turn entropy generation S'p is increased. Therefore, it is easily inferred that a low S'p is desirable. Note that the second law prohibits negative entropy generation in all real systems/processes, i.e. S'p ≥ 0 or Wp ≥ ∆Ap, establishing S'p = 0 or Wp = ∆Ap as the limit of possibility or an ideal case. In other words, in accordance with everyday experience, one canNOT obtain from the product more than one puts into its formation.

    • Manufacturing Equipment and Process

      In addition to characterizing and detecting unusual phenomena at the machine-workpiece interface, of significant interest are the efficiency and degradation of the manufacturing equipment and/or process. Energy balance on the machinery gives

       ∆Um/c = Qm/c + Wm/c         (5)

      where net machine work (work that transforms the machine)

      Wm/c = WinWp        (6)

      the difference between input work (e.g. power supplied to lathe or CNC machine during operation) and actual product formation work. Equation (6) indicates that some of the input power is lost between input point and the machine-workpiece interface, hence a low Wm/c is always desired. Wm/c includes all energy conversion losses, friction, Ohmic dissipation, corrosion, plasticity, shaft misalignment effects, etc, and should be resolved into appropriate constituents based on order of magnitude analysis of the specific system or manufacturing process. As done for the workpiece, a combined energy and entropy balance on the machinery (substituting entropy balance and equation (6) into equation (5)) yields

       ∆Um/c = Tm/c(∆S – S')m/c + Win – Wp .         (7)

      Rearranging with ∆Am/c = ∆Am/cTm/cSm/c, entropy generation in the machinery

       S'm/c = (Win – Wp – ∆Am/c) / Tm/c ≥ 0 .        (8)

      Here, ∆Am/c is the ideal machine/process capacity, a constant based on machine/process specifications, which could be dropped if unknown to give instantaneous entropy generation

      S'm/c = (WinWp) / Tm/c ≥ 0 ,         (9)

      the quotient of the difference between input work and minimum work required to manufacture component and representative machine temperature. Equation (9) includes all the dissipative phenomena taking place in the machinery during manufacture.

       

      Total entropy generation in both product and machinery

      S''total = S'p + S'm/c = (Wp – ∆Ap) / Tp + (WinWp) / Tm/c .        (10)

      If Tp ≈ Tm/c, equation (10) becomes

       S''total = (Win – ∆Ap) / Tp .        (11)

      S''total, equation (11), measures the efficiency of the entire system-process interaction and can be used as a first or basic analysis parameter, given its relative ease of evaluation: Win is usually known and ∆Ap is easily specified and standardized for a typical manufacturing process/task (e.g. drilling a hole in a thick steel plate or adding a thickener to grease in production) – a straight line joins Helmholtz energy states before and after task, an artifact of the thermodynamic state principle. Equations (4) and (9) give the individual contributions from workpiece and machinery respectively. Further sub-system analyses can be performed as necessary to determine the significant sources of irreversibilities in the process.

       

      Example: Battery Charging [1]

      For a simple energy-adding process like battery charging, equation (4) gives entropy generation in battery, in terms of voltage V, current I and temperature T,

       S' = (∫t VI dt – V0qrev) / T ≥ 0        (12)

      where ∫t VI dt can be determined via knowledge of the charging process, readily obtained from measurements, while V0qrev is determined via knowledge of the battery. Similar to the Helmholtz free energy change ∆A for non-reactive systems, the minimum recharge energy required for a battery or other electrochemical energy device is the change in its Gibbs free energy ∆G, i.e. ∆G = V0qrev. V0 is battery's standard potential and qrev can be evaluated using Faraday's first law.

    • Entropy – Generation OR Change?

      Manufacturing processes increase the workpiece’s Helmholtz energy ∆A ≥ 0, its availability (or available energy) to do work, to a maximum (finished state), while reducing its overall entropy ∆S ≤ 0. Note that while entropy generation S' is always non-negative, entropy change ∆S can be negative or positive: S is a state variable whereas S' is a "path" or "evolution" variable. From equation (3),

       ∆Sp = (∆Up + Tp S'pWp) / Tp ≤ 0       (13)

      indicating ∆Up + Tp S'p) ≤ Wp, the input product formation work at the machine-workpiece interface must equal or be greater than the required energy content of the finished product and the workpiece irreversibilities. However, according to the second law of Thermodynamics, the process described by equation (13) is not possible unless the total – component and machine – entropy is monotonically non-decreasing, i.e.

       ∆Stotal = ∆Sp + ∆Sm/c ≥ 0 .       (14)

      A comparison of equations (13) and (14) indicates that the overall entropy of the machine Sm/c (and that of the immediate surrounding considering entropy transfer out via heat) will increase by an amount equal to or greater than the reduction in the component’s entropy via the manufacturing process.

    • Product in Use

      After manufacture, an opposite process begins at the start of product use. Manufacturers and consumers are primarily concerned with a product’s usability and durability. The product’s free energy measures the usability of the product while entropy production, a measure of its degradation, determines its performance. Following similar procedure as the manufacturing process analysis,

       S'= (∆Ap – Wout) / Tp ≥ 0         (15)

      where ∆Ap and Wout are the product’s maximum and actual work outputs respectively.

    • Utility-Based Analysis vs Availability Analysis – Tp vs T0

      The result of a thermodynamic analysis and its interpretation depend considerably on the choice of boundary which in turn, depends on analyst’s preference or available/required information. Thermodynamics experts often estimate the maximum work obtainable from a system as it proceeds from its initial state to the thermodynamic dead state, a fixed state of equilibrium with the environment where T = T0, P = P0 [2–5]. Often referred to as the Guoy-Stodola theorem, exergetic destruction is given as

       T0 S' = Wrev – Wout ,        (16)

      similar to equation (15) and the converse versions in equations (4) and (11) (with Wrev = ∆Ap). However, it is worth noting that equations (15) and (16) will give different results under conditions where component temperature Tp is a variable on the “phenomenological” path, i.e. Tp ≠ T0, allowing for more accurate and consistent evaluation of internal irreversibilities within the system of interest only, while considering changes in the surroundings and the system’s thermal characteristic; whereas the use of constant T0 obtained far enough from the system where surroundings temperature is truly steady includes external irreversibilities in the portion of the surroundings between system and location of T0 establishment [2][5]. Moreover, if surroundings temperature is significantly different from T0 (typically 298K in most availability/exergy analyses), analysis results are less consistent with reality, possibly impacting the widespread usage of Thermodynamics in degradation/performance modeling of non-thermal systems.

       

      The formulations derived in this article present convenient and consistent ways of assessing a manufacturing process, sub-process and/or multiple simultaneously occurring processes, using “utility-based” entropy generation (equations (4), (9) and (11)): the Helmholtz and Gibbs potentials or free energies – while contributing, along with Enthalpy, the letters that form EnHeGi – conveniently encapsulate energy transfer via heat, and are especially useful when heat capacities and heat exchanges with the environment during manufacture are unknown or difficult to measure accurately. “Utility-based” entropy generation measures the wastage/losses in any system-process interaction, a knowledge of which is critical to improving system and process efficiencies, and reducing costs. As shown previously, an efficient process minimizes its entropy generation, i.e. a low S’ is preferred.

    • REFERENCES

      [1] Osara J. A. The Thermodynamics of Degradation. Doctoral Thesis, UT Austin; 2017.

      [2] Moran M. J. Shapiro H N, Fundamentals of Engineering Thermodynamics. 5th ed. Wiley; 2004.

      [3] Burghardt M. D. Harbach J A, Engineering Thermodynamics. 4th ed. HarperCollins College Publishers; 1993.

      [4] Bejan A. The Method of Entropy Generation Minimization. Energy and the Environment; 1990:11–22.

      [5] Bejan A. Advanced Engineering Thermodynamics. vol. 70. 3rd ed. John Wiley & Sons Inc; 1997.

    • FOR MORE INFO, CONTACT US AT [email protected]

© 2020 ENHeGi RESEARCH AND ENGINEERInG.

 ALL RIGHtS ReSERVED.

    • where Qp is the net heat transfer between component and surroundings during manufacture and Wp is the net work that transformed the material. Via the second law, the entropy change during manufacture

    • the difference between input work (e.g. power supplied to lathe or CNC machine during operation) and actual product formation work. Equation (6) indicates that some of the input power is lost between input point and the machine-workpiece interface, hence a low Wm/c is always desired. Wm/c includes all energy conversion losses, friction, Ohmic dissipation, corrosion, plasticity, shaft misalignment effects, etc, and should be resolved into appropriate constituents based on order of magnitude analysis of the specific system or manufacturing process. As done for the workpiece, a combined energy and entropy balance on the machinery (substituting entropy balance and equation (6) into equation (5)) yields

      where ∫t VI dt can be determined via knowledge of the charging process, readily obtained from measurements, while V0qrev is determined via knowledge of the battery. Similar to the Helmholtz free energy change ∆A for non-reactive systems, the minimum recharge energy required for a battery or other electrochemical energy device is the change in its Gibbs free energy ∆G, i.e. ∆G = V0qrev. V0 is battery's standard potential and qrev can be evaluated using Faraday's first law.

<<UPDATED 06/29/19 >>