Skip to main content

Optimization of Daniel Cam Engines

  • Chapter
  • First Online:
Optimal Control in Thermal Engineering

Part of the book series: Studies in Systems, Decision and Control ((SSDC,volume 93))

  • 1190 Accesses

Abstract

In this chapter the piston motion of a Daniel cam engine is optimized. Piston acceleration is taken as a control. The objective is to maximize the net output work during the compression and power strokes. A thermally insulated cylinder is considered and a realistic model taking into account the cooling system is developed. The sinusoidal approximation of piston motion in the classical rod-crank system overestimates the engine efficiency. The exact description of the piston motion in rod-crank system is used here as a reference. The net output work is much larger (by 12–13%) for the optimized system than for the classical rod-crank system, for similar thickness of cylinder walls and thermal insulation. The optimized cam is smaller for a cylinder without thermal insulation than for an insulated cylinder (by up to 8%, depending on the local polar radius).

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Annand, W.J.D.: Heat transfer in the cylinder of reciprocating internal combustion engines. Proc. Inst. Mech. Eng. 177, 973–990 (1963)

    Google Scholar 

  • Azadi, M., Baloo, M., Farrahi, G.H., Mirsalim, S.M.: A review of thermal barrier coating effects on diesel engine performance and components lifetime. Int. J. Automot. Eng. 3, 305–317 (2013)

    Google Scholar 

  • Badescu, V.: Optimal piston motion for maximum net output work of Daniel cam engines with low heat rejection. Energy Convers. Manag. 101, 181–202 (2015)

    Article  Google Scholar 

  • Betts, J.T.: Practical methods for optimal control using nonlinear programming. Society for Industrial and Applied Mathematics (SIAM), Philadelphia (2001)

    Google Scholar 

  • Burzler, J.M., Blaudeck, P., Hoffmann, K.H.: Optimal piston paths for diesel engines. In: Sieniutycz, S., De Vos, A. (eds.) Thermodynamics of energy conversion and transport, pp. 173–198. Springer, New York (2000)

    Chapter  Google Scholar 

  • Cam engine: http://en.wikipedia.org/wiki/Cam_engine (2016)

  • Chen, L., Song, H., Sun, F.: Endoreversible radiative heat engine configuration for maximum efficiency. Appl. Math. Model. 34, 1710–1720 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Chen, L., Xia, S., Sun, F.: Optimizing piston velocity profile for maximum work output from a generalized radiative law Diesel engine. Math. Comput. Model. 54, 2051–2063 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Ciniviz, M., Salman, M.S., Canli, E., Kose, H., Solmaz, O.: Ceramic coating applications and research fields for internal combustion engines, INTECH Open Access Publisher; http://cdn.intechopen.com/pdfs-wm/29756.pdf (2012). Accessed 5 Mar 2015

  • Das, P., Subbarao, P.M.V., Subrahmanyam, J.P.: Effect of main injection timing for controlling the combustion phasing of a homogeneous charge compression ignition engine using a new dual injection strategy. Energy Convers. Manag. 95, 248–258 (2015)

    Article  Google Scholar 

  • Dawson, S.: Compacted graphite iron—a material solution for modern diesel engine cylinder blocks and heads. China Foundry 6, 241–246 (2009)

    Google Scholar 

  • Diviš, M., Tichánek, R., Španiel, M.: Heat transfer analysis of a diesel engine head. Acta Polytechnica 43, 34–39 (2003)

    Google Scholar 

  • Doric, J.Z., Klinar, I.J.: Efficiency characteristics of a new quasi-constant volume combustion spark ignition engine. Therm. Sci. 17, 119–133 (2012)

    Article  Google Scholar 

  • Douglas: Internal-combustion cam engines. http://www.douglas-self.com/MUSEUM/POWER/unusualICeng/cam-IC/cam-IC.htm (2016)

  • Ge, Y., Chen, L., Sun, F.: Optimal paths of piston motion of irreversible diesel cycle for minimum entropy generation. Therm. Sci. 15, 975–993 (2011)

    Article  Google Scholar 

  • Gerdts, M.: OCPID-DAE1, User’s guide, Optimal control and parameter identification with differential-algebraic equations of index 1, Version 1.3, March 20, 2013, Institut fur Mathematik und Rechneranwendung, Universitat der Bundeswehr Munchen, Germany; http: www.unibw.de/lrt1/gerdts (2013)

  • German Engineers: Zeitschrift des vereines Deutscher Ingenieure (The magazine of the Association of German Engineers), p. 1405 (1925)

    Google Scholar 

  • Ghojel, H., Honnery, D.: Heat release model for the combustion of diesel oil emulsions in DI diesel engines. Appl. Therm. Eng. 25, 2072–2085 (2005)

    Article  Google Scholar 

  • Hoffman, K.H., Berry, R.S.: Optimal paths for thermodynamic systems: the ideal diesel cycle. J. Appl. Phys. 58, 2125–2134 (1985)

    Article  Google Scholar 

  • Jafarmadar, S., Tasoujiazar, R., Jalilpour, B.: Exergy analysis in a low heat rejection IDI diesel engine by three dimensional modeling. Int. J. Energy Res. 38, 791–803 (2014)

    Article  Google Scholar 

  • Jia, B., Zuo, X., Tian, G., Feng, H., Roskilly, A.P.: Development and validation of a free-piston engine generator numerical model. Energy Convers. Manag. 91, 333–341 (2015)

    Article  Google Scholar 

  • Junk, H., Lenz, W.: Iron cylinder blocks for high-volume production. MTZ Worldwide 67, 27–29 (2006)

    Article  Google Scholar 

  • Kamo, R, Bryzik, W.: Adiabatic turbo-compound engine performance prediction. SAE International, Paper No. 780068 (1978)

    Google Scholar 

  • Karabulut, H.: Dynamic analysis of a free piston stirling engine working with closed and open thermodynamic cycles. Renew. Energy 36, 1704–1709 (2011)

    Article  Google Scholar 

  • Kleinschmidt, W.: Die Warmeubertragung in aufgeladenen Dieselmotoren aus neuerer Sicht. 5. Aufladetechnische Konferenz, Augsburg (1993)

    Google Scholar 

  • Lara-Curzio, E., Readey, M.J. (eds.): 28th International Conference on Advanced Ceramics and Composites B: Ceramic Engineering and Science Proceedings 25, pp. 2–653 (2008)

    Google Scholar 

  • Li, J., Chen, L., Sun, F.: Optimal configuration of a class of endoreversible heat-engines for maximum power-output with linear phenomenological heat-transfer law. Appl. Energy 84, 944–957 (2007)

    Article  Google Scholar 

  • Ma, K., Chen, L., Sun, F.: Optimizations of a model external combustion engine for maximum work output with generalized radiative heat transfer law. Int. J. Energy Environ. 2, 723–738 (2011)

    Google Scholar 

  • Marr, M.A.: An investigation of metal and ceramic thermal barrier coatings in a spark ignition engine. MS Thesis, Department of Mechanical and Industrial Engineering, University of Toronto. https://tspace.library.utoronto.ca/bitstream/1807/18906/1/Marr_Michael_A_200911_MASc_thesis.pdf (2009). Accessed 5 Mar 2015

  • Murthy, P.V.K., Murali Krishna, M.V.S., Sitarama Raju, A., Vara Prasad, C.M., Srinivasulu, N.V.: Performance evaluation of low heat rejection diesel engine with pure diesel. Int. J. Appl. Eng. Res. Dindigul 1, 428–451 (2010)

    Google Scholar 

  • NACA Memorandum: Comments on crankless engine types. NACA Technical Memorandum (462). Washington DC: NACA, May 1928

    Google Scholar 

  • Ngayihi Abbe, C.V., Nzengwa, R., Danwe, R., Ayissi, Z.M., Obonou, M.: A study on the 0D phenomenological model for diesel engine simulation: application to combustion of neem methyl esther biodiesel. Energy Convers. Manag. 89, 568–576 (2015)

    Article  Google Scholar 

  • Nocedal, J., Wright, S.J.: Numerical optimization. Springer, New York (1999)

    Book  MATH  Google Scholar 

  • Rakopoulos, C.D., Mavropoulos, G.C.: Study of the steady and transient temperature field and heat flow in the combustion chamber components of a medium speed diesel engine using finite element analyses. Int. J. Energy Res. 20, 437–464 (1996)

    Article  Google Scholar 

  • Rakopoulos, C.D., Antonopoulos, K.A., Rakopoulos, D.C., Giakoumis, E.G.: Investigation of the temperature oscillations in the cylinder walls of a diesel engine with special reference to the limited cooled case. Int. J. Energy Res. 28, 977–1002 (2004)

    Article  Google Scholar 

  • Rakopoulos, C.D., Giakoumis, E.G., Rakopoulos, D.C.: Study of the short-term cylinder wall temperature oscillations during transient operation of a turbo-charged diesel engine with various insulation schemes. Int. J. Engine Res. 9, 177–193 (2008)

    Article  Google Scholar 

  • Ratna Reddy, T., Murali Krishna, M.V.S., Kesava Reddy, C., Murthy, P.V.K.: Performance evaluation of a low heat rejection diesel engine with Mohr oil based biodiesel. Brit. J. Appl. Sci. Technol. 2, 179–198 (2012)

    Article  Google Scholar 

  • Shrets, I., Tolubinsky, V., Kirakovsky, N., Neduzhy, I., Sheludho, I.: Heat engineering, 2nd edn. Mir Publishers, Moscow (1980)

    Google Scholar 

  • Sivakumar, G., Shankar, V., Hemath Kumar, G., Renganathan, N.G., Garud, V.U.: Is thermal barrier coating for low heat rejection in SI engines or diesel engines? Int. J. Emerg. Technol. Adv. Eng. 2, 460–466 (2012)

    Google Scholar 

  • Song, H., Chen, L., Sun, F.: Endoreversible heat-engines for maximum power-output with fixed duration and radiative heat-transfer law. Appl. Energy 84, 374–388 (2007)

    Article  Google Scholar 

  • Taylor, C.F.: The internal-combustion engine in theory and practice, 2nd edn., p. 579. MIT Press, Cambridge (1985)

    Google Scholar 

  • Teh, K.-Y., Miller, S.L., Edwards, C.F.: Thermodynamic requirements for maximum internal combustion engine cycle efficiency. Part 1: optimal combustion strategy. Int. J. Eng. Res. 9, 449–465 (2008)

    Article  Google Scholar 

  • US Patent: US Patent 817,905, 17 April 1906

    Google Scholar 

  • US Patent: US Patent no. 1654378; Marchetti Motor Patents Inc., Mills Field, San Bruno, California, Dec 1927

    Google Scholar 

  • Wang, Q.: Numerical analysis of cooling effects of a cylinder head water jacket, MsD thesis, University of Minnesota, Duluth (2009)

    Google Scholar 

  • Xia, S., Chen, L., Sun, F.: The optimal path of piston motion for Otto cycle with linear phenomenological heat transfer law. Sci. China Ser. G-Phys. Mech. Astron. 52, 708–719 (2009)

    Article  Google Scholar 

  • Xia, S., Chen, L., Sun, F.: Engine performance improved by controlling piston motion: Linear phenomenological law system Diesel cycle. Int. J. Therm. Sci. 51, 163–174 (2012)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Viorel Badescu .

Appendices

Appendix 21A

1.1 21.A.1 Combustion

The combustion process in CIE is described in details in many papers. Reasonably accurate results may be obtained by using simpler fixed (frozen) composition analysis in the range of pressure 1–100 bar, temperature 800–2500 K and relative air-fuel ratio 0.7–3.35 (Ghojel and Honnery 2005). Also, single-zone models are much simpler and easier to run as compared with multi-zone models.

The simple single-zone combustion model proposed by Burzler et al. (2000) has been used by other authors (Xia et al. 2012) and is used here. The fuel is usually injected during the final stage of the compression stroke. It evaporates in the hot air and after a short delay ignites and starts to burn rapidly. The moment when the fuel ignites, measured from the start of the compression stroke is called auto-ignition moment and denoted \( t_{z} \). The reaction coordinate \( \xi (t) \) describes the extent of the combustion. Burzler et al. (2000) used the heat production curves of real CIEs investigated by Kleinschmidt (1993) and the following approximation has been obtained:

$$ \xi (t) = \left\{ {\begin{array}{*{20}l} {0\quad for\;0 \le t \le t_{z} } \hfill \\ {1 + \left( {\frac{{t - t_{z} }}{{t_{b} }} - 1} \right)\exp \left( {\frac{{t_{z} - t}}{{t_{b} }}} \right)\quad for\;t_{z} < t \le t_{z} + t_{b} } \hfill \\ {1\quad for\;t_{z} + t_{b} < t \le t_{tot} } \hfill \\ \end{array} } \right. $$
(21.A.1)

with a characteristic combustion duration \( t_{b} \). Before the auto-ignition moment \( t_{z} \) the reaction coordinate is \( \xi (t \le t_{z} ) = 0 \) while \( \xi (t) = 1 \) means total combustion. Since ttot denotes the time for compression and power stroke, the position of the crank angle A associated with the auto-ignition moment \( t_{z} \) is simply approximated by \( A = (t_{z} /t_{tot} ) \cdot 360^{ \circ } \). Note that A = 180° when the piston is closest to the fire deck.

The charge consists of combustion gases and is treated as an ideal gas whose time variation of the mole number N and heat capacity C depend on the reaction coordinate :

$$ \begin{aligned} N(t) & = N_{i} + (N_{f} - N_{i} )\xi (t) \\ C(t) & = C_{i} + (C_{f} - C_{i} )\xi (t) \\ \end{aligned} $$
(21.A.2,A.3)

where the subscript i and f denote the initial and final combustion moments, respectively. The heating function \( h(t) \) describes the rate of heat generated during combustion:

$$ h(t) = Q_{c} N_{i} \dot{\xi }(t) $$
(21.A.4)

Here \( Q_{c} \) is the molar heat of the air-fuel mixture.

1.2 21.A.2 Heat Losses

During the compression and power strokes, heat losses occur through the cylinder walls. One denotes by T and \( T_{wg} \) the working fluid temperature and the temperature at the inner surface of the cylinder wall, respectively. Both quantities are space averages. Two different approaches are considered to calculate the heat flux \( q_{c} \) transferred from the charge towards the cylinder walls. The first approach is called Newton-type heat transfer and is based on the relationship:

$$ q_{c}^{N} = Ak_{N} \left( {T - T_{wg} } \right) $$
(21.A.5)

where \( k_{N} \) is an appropriate constant quasi-static heat transfer coefficient. Equation (21.A.5) is widely used in literature primarily due to its simplicity (Ge et al. 2011) and is used here as a reference.

In the absence of an universally applicable heat transfer model for CIEs most authors prefer simple methods. For instance, the Wiebe metho d was adopted by Ghojel and Honnery (2005) while the Whitehouse and Way method was adopted by Rakopoulos et al. (2008). However, many authors are using the model of Annand (1963) [see e.g. Burzler et al. (2000); Murthy et al. (2010)]. The Annand model takes account on both conduction and convection and is based on the relationship:

$$ q_{c}^{A} = Ak_{A} (T,x)(T - T_{wg} ) $$
(21.A.6)

where \( k_{A} \) is a time-variable heat transfer coefficient given by:

$$ k_{A} (T,x) = a\frac{\kappa (T)}{d}\text{Re}^{b} (T,x) $$
(21.A.7)

where \( \kappa (T) \) is charge’s thermal conductivity, \( \text{Re} \) is Reynolds number while a and b are empirical coefficients. Relationships used to compute \( \kappa (T) \) and \( \text{Re} (T,x) \) are shown in Table 21.8.

Table 21.8 Relationships used to compute \( k(T) \) and \( \text{Re} (T,x) \) in Eq. (21.A.7)

During the power stroke solid incandescent carbon particles appear as intermediate combustion products and radiate heat towards the cylinder’s walls. The following relationship is used to estimate the heat flux \( q_{r} \) lost by radiative transfer (Annand 1963):

$$ q_{r} = Ac\sigma (T^{4} - T_{wg}^{4} ) $$
(21.A.8)

Here \( \sigma \) is Stefan-Boltzmann constant while c is an empirical coefficient. During the compression stroke \( c = 0 \) while during the power stroke c ranges between 0.04 and 0.32 for CIEs (Annand 1963).

The total heat flux \( q_{lost} \) lost by the charge may be obtained simply by summing the losses by conduction and convection and by radiation, respectively:

$$ q_{lost} = q_{c} + q_{r} $$
(21.A.9)

where \( q_{c} \) is a shorthand for \( q_{c}^{N} \) or \( q_{c}^{A} \), depending on the approximation adopted.

1.3 21.A.3 Frictional Losses

The frictional loss rate of mechanical energy \( w_{f} \) is proportional to the square of piston velocity v (Burzler et al. 2000):

$$ w_{f} = \alpha v^{2} $$
(21.A.10)

where \( \alpha \) is a coefficient of constant value during the compression and power stroke. The assumption that the heat generated by friction does not contribute to the engine heat production as a low-grade heat source is adopted.

Appendix 21B

1.1 21.B.1 Classical Rod-Crank System

In case of the classical rod-crank system the piston movement is obtained by using Eqs. (21.2.3, 21.2.4). Equation (21.2.5) is not needed. In the following, the specific form of Eqs. (21.2.3, 21.2.4) is derived. The movement of the rod-crank system is ideally described in Fig. 21.19 where r and l are the length of the crank and rod, respectively. O and P denote crank and piston position, respectively, while M is the fire deck position. Also, A is crank angle while \( x_{0} \) and \( x_{f} \) denote the minimum and maximum distance between piston and fire deck. Note that A = 180° when the piston is closest to the fire deck .

Fig. 21.19
figure 19

Movement of the rod-crank system. P denotes the piston while r and l are the lengths of the crank and rod, respectively. A is crank angle (=180° at the top-dead-center, i.e. the position when the piston is closest to the fire deck)

Piston position is described in terms of r and l by the following equation:

$$ y = - r\cos A + \sqrt {l^{2} - r^{2} \sin^{2} A} $$
(21.B.1)

where:

$$ A \equiv \frac{2\pi t}{{t_{tot} }} $$
(21.B.2)

Here t denotes the time while \( t_{tot} \) denotes the duration of the compression and power stroke together. Equation (21.B.1) is now used to derive the piston motion in x coordinate. One easily sees that:

$$ \begin{aligned} y_{\hbox{max} } & \equiv y\left( {\cos A = - 1} \right) = l + r \\ y_{\hbox{min} } & \equiv y(\cos A = 1) = l - r \\ \end{aligned} $$
(21.B.3a,b)

Geometry constraints yield:

$$ x + y = x_{0} + y_{\hbox{max} } = x_{f} + y_{\hbox{min} } \equiv D $$
(21.B.4a–d)

where the constant D is still to be prescribed. Usage of Eqs. (21.B.3a, b) and (21.B.4a, b) yields

$$ x_{0} + r + l = x_{f} + l - r = D $$
(21.B.5a,b)

From Eq. (21.B.5a) one finds:

$$ r = \frac{{x_{f} - x_{0} }}{2} $$
(21.B.6)

Equations (21.B.5b) and (21.B.6) yield:

$$ l = D - \frac{{x_{f} + x_{0} }}{2} $$
(21.B.7)

Equations (21.B.4ac) shows that \( x + y = D \) and usage of Eq. (21.B.1) gives:

$$ x = D + r\cos A - \sqrt {l^{2} - r^{2} \sin^{2} A} $$
(21.B.8)

From Eqs. (21.B.6), (21.B.7) and (21.B.8) one finds after some algebra:

$$ x(t) = \frac{{x_{f} - x_{0} }}{2}\cos A + D - \left( {D - \frac{{x_{f} + x_{0} }}{2}} \right)\sqrt {1 - \left( {\frac{{x_{f} - x_{0} }}{{2D - (x_{f} + x_{0} )}}} \right)^{2} \sin^{2} A} $$
(21.B.9)

This is the piston movement law in \( x(t) \) coordinates.

In case the distance between the crank shaft and the cylinder is large, the approximation often used is \( D \to \infty \). Then, Eq. (21.B.9) yields:

$$ x = \frac{{x_{f} - x_{0} }}{2}\cos A + \frac{{x_{f} + x_{0} }}{2} $$
(21.B.10)

The sinusoidal piston law Eq. (21.B.1) is an approximation often used in practice. Here the more realistic Eq. (21.B.9) is implemented for the particular case:

$$ D = x_{f} + x_{0} $$
(21.B.11)

Equations (21.B.9) and (21.B.11) give:

$$ x(t) = \frac{{x_{f} - x_{0} }}{2}\cos A + \frac{{x_{f} + x_{0} }}{2}\sqrt {1 - \left( {\frac{{x_{f} - x_{0} }}{{x_{f} + x_{0} }}} \right)^{2} \sin^{2} A} $$
(21.B.12)

The speed and acceleration laws \( \dot{x}(t) \) and \( \ddot{x}(t) \), respectively, are given by:

$$ \dot{x}(t) = \frac{2\pi }{{t_{tot} }}\left[ { - r\sin A - \frac{{r^{2} \sin A\cos A}}{{\sqrt {l^{2} - r^{2} \sin^{2} A} }}} \right] $$
(21.B.13)
$$ \ddot{x}(t) = \left( {\frac{2\pi }{{t_{tot} }}} \right)^{2} \left[ { - r\cos A - \frac{{r^{2} \left( {\cos^{2} A - \sin^{2} A} \right)}}{{\sqrt {l^{2} - r^{2} \sin^{2} A} }} - \frac{{r^{4} \sin^{2} A\cos^{2} A}}{{\left( {\sqrt {l^{2} - r^{2} \sin^{2} A} } \right)^{3} }}} \right] $$
(21.B.14)

where r and l are given by Eqs. (21.B.6) and (21.B.11), respectively.

Note that in some works the classical rod-crank system under the approximation of a sinusoidal piston movement has been used as a reference (Hoffmann and Berry 1985; Burzler et al. 2000). In this chapter the “exact” solution Eq. (21.B.12) is used as a reference. The approximate sinusoidal solution Eq. (21.B.10) is considered only in Sect. 21.3.1.1.

Appendix 21C

See Figures  21.20, 21.21, 21.22 and 21.23; Tables 21.9, 21.10, 21.11, 21.12 and 21.13.

Fig. 21.20
figure 20

Pressure–volume diagram for optimally controlled cam-lever system and classical rod-crank system. Newton and Annand heat transfer models are shown

Fig. 21.21
figure 21

Pressure–volume diagram for optimally controlled cam-lever system. Two cases have been considered: no thermal insulation and a PSZ layer of 1.5 mm thickness, respectively

Fig. 21.22
figure 22

Pressure–volume diagram for optimally controlled cam-lever system, for two values of the auto-ignition moment \( t_{z} \)

Fig. 21.23
figure 23

Pressure–volume diagram for optimally controlled cam-lever system. Two values of the heat convection coefficient h c have been considered

Table 21.9 Results obtained by using the optimized cam-lever system and the classical rod-crank system
Table 21.10 The dependence of the net output work W on the module of the maximum acceleration
Table 21.11 The same as Table 21.5 but the heat transfer model from cylinder wall to the cooling fluid is not taken into consideration
Table 21.12 The same as Table 21.6 but results for the classical rod-crank system are shown
Table 21.13 The same as Table 21.7 but results for the classical rod-crank system are shown

Rights and permissions

Reprints and permissions

Copyright information

© 2017 Springer International Publishing AG

About this chapter

Cite this chapter

Badescu, V. (2017). Optimization of Daniel Cam Engines. In: Optimal Control in Thermal Engineering. Studies in Systems, Decision and Control, vol 93. Springer, Cham. https://doi.org/10.1007/978-3-319-52968-4_21

Download citation

  • DOI: https://doi.org/10.1007/978-3-319-52968-4_21

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-319-52967-7

  • Online ISBN: 978-3-319-52968-4

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics