Next Article in Journal
Feasibility Study of Robust Optimization to Reduce Dose Delivery Uncertainty by Potential Applicator Displacements for a Cervix Brachytherapy
Previous Article in Journal
Sensitivity and Uncertainty Analyses of Human and Organizational Risks in Fire Safety Systems for High-Rise Residential Buildings with Probabilistic T-H-O-Risk Methodology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biofuels from Micro-Organisms: Thermodynamic Considerations on the Role of Electrochemical Potential on Micro-Organisms Growth

Dipartimento Energia “Galileo Ferraris”, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2021, 11(6), 2591; https://doi.org/10.3390/app11062591
Submission received: 1 March 2021 / Revised: 11 March 2021 / Accepted: 12 March 2021 / Published: 14 March 2021
(This article belongs to the Section Energy Science and Technology)

Abstract

:
Biofuels from micro-organisms represents a possible response to the carbon dioxide mitigation. One open problem is to improve their productivity, in terms of biofuels production. To do so, an improvement of the present model of growth and production is required. However, this implies an understanding of the growth spontaneous conditions of the bacteria. In this paper, a thermodynamic approach is developed in order to highlight the fundamental role of the electrochemical potential in bacteria proliferation. Temperature effect on the biosystem behaviour has been pointed out. The results link together the electrochemical potential, the membrane electric potential, the pH gradient through the membrane, and the temperature, with the result of improving the thermodynamic approaches, usually introduced in this topic of research.

1. Introduction

In 2017, UN IPCC Conference in Marrakesh suggested containing the increase in temperature up to 2 C, in order to prevent the overcoming of the thermal threshold for a possible disastrous climate change [1] due to the use of fossil fuels as main source of energy [2]. In the last decade a growing interest in biofuels from bacteria has emerged, due to their sustainable characteristics [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17].
However, at present, the third generation of biofuels is not yet competitive, in comparison to fossil fuels, because they present high costs of production and extraction, in relation to the amount of fuel obtainable per unite volume of culture [18].
In order to improve their competitiveness, the control of bacteria growth is required. This topic has been studied in the second half of the XX century, starting with Monod [19,20], who developed a whole theory on bacteria growth, by introducing a hyperbolic relationship, which links the amount of limiting resources in the environment to the growth rates of microorganisms. This first result has been improved by coupling it to other models able to fit the experimental data [21,22,23,24]. In 1956, the concept of continuous culture has been introduced with the aim to obtain biological cultures able to operates continuously for a long time [24,25]. This approach showed, in 1958, that one of the fundamental requirement of life is the maintenance of concentration or osmotic gradients [26]. Then, the Black cox model was introduced in order to obtain the relationships among energy, biomass and yield [24], by pointing out that the growth of bacteria, on a single carbon compound as sole energy source, uses a constant amount of energy, per electron, to produce dry weight [27]. Mayberry et al. pointed out the role of the electrons (as electric charge) in the bacteria growth, in relation to the Oxygen fluxes [28].This thermochemical approach led to confirm the Thornton’s rule [29]. However, the thermodynamic approach has been highlighted to be too dependent by the approximation of enthalpy and Gibbs free energy, with the consequence to require a new development.
In this context, biofuel production by microorganisms must be analysed in order to optimise their production process [30,31,32], and it can be carried out from an improvement of the thermodynamic approach to biophysical processes. We have recently developed a thermodynamic approach to mutualism [4,33], in order to improve the biodiesel production from bacteria and algae [34,35,36,37,38,39,40], in the context of a new general approach to sustainability [41]. So, we suggest an explanation to the behaviour of the bacteria in relation to a new nonequilibrium thermodynamic approach to bacteria membrane. Indeed, bacterial electrical signalling has experimentally been proven to be able to regulate a great variety of physiological processes, mediated by membrane electric potential [42,43,44,45], such as proliferation [46].
In accordance with the previous biological and biochemical results, the inflow of nutrients and outflow of waste products occur through the membrane, with a related restoration of ionic gradients, changed by the transport systems. There is a great number of theoretical and computational studies on the membrane mechanics, but these models cannot obtain a full description of the membrane dynamical and chemical behaviour [47]. There is a continuous interaction between bacteria and their environment; indeed, all bacteria are able to modify their environment by secretion of enzymes, toxins, or pheromones and by macromolecules on their external membrane surfaces. This interaction is the key-point of our scientific interest: we wish to point out the conditions of bacteria environment control, useful to improve the production of biofuels by micro-organisms.
In this paper, we develop the analysis of the role of the membrane potential in the behaviour of the bacteria, in relation to their ability to produce biofuels, in order to suggest an approach to improve their productivity, starting from a natural behaviour of these living biosystems. Last, we suggest a possible improvement of this behaviour, by introducing the cooperative interaction between different species (mutualism).

2. Materials and Methods

The aim of this paper is to introduce, into the thermodynamic analysis of the biofuels production by bacteria, a nonequilibrium thermodynamic approach to membrane heat and mass transport, recently developed in relation to life [48] and applied to cancer [49,50,51,52] and glaucoma [53], with results in accordance with the experimental results in literature, both in relation to cancer [54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80] and to glaucoma [81,82,83,84,85,86,87,88,89].
The nonequilibrium thermodynamic approach has been obtained by introducing the Onsager general phenomenological relations, previously related only to membrane electric potential, and, here, generalised by introducing the electrochemical potential [48,49,90,91,92,93]:
J e = L 11 μ e T L 12 T T 2 J Q = L 21 μ e T L 22 T T 2
where J e is the net current density [A m 2 ], if the effect of some different species f ions are considered or simply the effect of one species if only one ion species is considered, J Q denotes the heat flux [W m 2 ], μ e = μ + Z e ϕ [92,93,94] is the electrochemical potential [J mol 1 ], with μ the chemical potential [J mol 1 ], z e the electric charge [A s mol 1 ], and ϕ the membrane potential [V], T is the living cell temperature and L i j represent the phenomenological coefficients, such that [94] L 12 ( B ) = L 21 ( B ) (Onsager-Casimir relation [95]), and L 11 0 and L 22 0 , and [94] L 11 L 22 L 12 L 21 > 0 .
The result consists in modelling the life cycle of a living cell by introducing two related processes [48,49]:
  • A continuous energy generation (metabolism), due to the ion fluxes: the ions and metabolites fluxes can be described by imposing J e 0 and J Q = 0 ;
  • A continuous heat fluxes from the cell to its environment: The heat exchange towards the environment can be described by imposing J e = 0 and J Q 0 .
In this way, we can split the life cycle into two thermodynamic processes, as usually done in thermodynamics for any complex process [91,96].
If the ions and metabolites fluxes occur, J e 0 and J Q = 0 , so, by taking into account the Equations (1), it is possible to obtain [48,90,91]
d μ e d T = L 21 L 11 1 T
with a related heat flux [90,91]:
d u d t = · J Q
where u is the internal energy density [J m 3 ].
Living bacteria exchange heat power towards their environment by convective transmission and thermal infrared emission, so, as a consequence of the First Law of Thermodynamics, it follows [51]
d u d t d V = δ Q ˙ = α ( T T 0 ) d A ϵ i r r σ S B T 4 T 0 4 d A
where ϵ i r r 0.97 is the emissivity factor [97,98,99], σ S B = 5.67 × 10 8 W m 2 K 4 is the Stefan–Boltzmann constant, α is the coefficient of convection, A area of the external surface of the cell membrane, V is the cell volume, T is the mean temperature of the external surface of the bacteria membrane, and T 0 is the temperature of their environment.
So, considering the Equations (3) and (4), together with the Divergence Theorem [100], the heat power exchanged:
Q ˙ = A J Q · n ^ d A α ( T T 0 ) A + ϵ i r r σ S B T 4 T 0 4 A
However, considering Equation (1), and that ϵ i r r σ S B T 4 T 0 4 < < α ( T T 0 ) , and the second hypothesis of our approach ( J e = 0 , J Q 0 ), we can obtain [48]:
d μ e d = T J Q L 22 L 11 L 12 L 21 = α T ( T T 0 ) L 22 L 11 L 12 L 21
where is the length of bacteria membrane and | μ e | d μ e / d . This last relation represents a link between the bacteria membrane electric potential and the temperature of the bacteria but to their heat exchange, too. Now, Equation (6) allow us to obtain:
α T ( T T 0 ) = L 22 L 11 L 12 L 21 μ e
Considering that [93]:
L 22 L 21 L 12 L 11 = K J T 2
where K J is the Thomson coefficient [93], it is possible to obtain that:
μ e = μ e T α K J T s u r f T 0
which, considering that μ e = μ + z e ϕ , allows us to obtain:
μ = z e d ϕ d + μ e T α K J T s u r f T 0
We can point out that the first fundamental quantity in the equation is the chemical potential, which is defined as [91]:
μ i = G n i T , p , n k i G n i = g
where G is the Gibbs energy, g is the Gibbs molar specific energy, n is the number of moles, and p is the pressure. Moreover, the Gibbs energy is related to the membrane electric potential by the Nernst equation [101]:
Δ g = F Δ ϕ 2.3 R T 0 Δ pH
where F = 96,485 A s mol 1 is the Faraday constant and R = 8.314 J mol 1 K 1 is the universal constant of ideal gas. Now, introducing the Equation (12) into the Equation (10), we can obtain:
μ e T = K J α F + z e T s u r f T 0 d ϕ d K J α 2.3 R T 0 T s u r f T 0 d pH d
As a consequence of the previous processes, a density entropy rate due to irreversibility (dissipation function [90]) is generated [102]:
σ = 1 T 0 i = 1 N J i · μ i = i = 1 N J i z i e T 0 d ϕ d + J u α K J 1 T s u r f T 0 i = 1 N μ e , i T
where T 0 is the environmental temperature, i = 1 N μ i J i is the contribution of the inflows and outflows, and μ is the chemical potential. Considering that σ 0 , we can find the condition of control for the previous relations:
i = 1 N J i z i e T 0 d ϕ i d J u α K J 1 T s u r f T 0 i = 1 N μ e , i T i = 1 N J i z i e ϕ i J u α K J ( T s u r f T 0 ) i = 1 N μ e , i T

3. Results

The fundamental result of this paper is Equation (10). Indeed, this equation represents a link among the chemical potential gradient and the electric potential through the membrane and the thermophysical properties of the membrane and the environmental fluid, related to the temperature of the micro-organisms membrane external surface. Indeed, the electrochemical potential variation with temperature in Equation (13) shows a dependence on the pH gradient. This result is in accordance with the behaviour of the analysis of the lipid membranes, related to chemical stimuli; indeed, lipid membranes are extremely responsive to chemical stimuli, with particular regards to pH gradients [47], as experimentally proven by using synthetic membranes [47,103]. In this context, it is possible to change the pH gradient by introducing the interaction between different species, as verified experimentally, for example, for mutualism between Chlorella vulgaris and Spirulina platensis [104].
Proteins play a fundamental role in ion transport. Proteins in the cytosolic can be modified in their functions by phosphorylation or dephosphorylation. An ion actively crosses the membrane against its electrochemical potential, whereby the necessary energy is derived either from the hydrolysis of ATP, or from the movement of a cotransported, or coupled ion along its electrochemical gradient. In this context, the role played by the H + -ATPase is fundamental, because it moves positive charges into the cell, while it generates membrane voltage and a pH gradient [105,106,107]. Protein phosphorylation is an important cellular regulatory mechanism, because many enzymes and receptors [51,108,109] are activated or deactivated by phosphorylation [110,111,112,113].
Moreover, the second fundamental quantity in Equation (10) is the coefficient of convection α [114], directly related to the characteristic length R = V / A . This is a geometrical quantity, and it is related to the shape of the micro-organism. The geometrical characteristic of a system plays a fundamental role in the heat exchange. Recently, it has been shown its importance in the bacteria biological behaviour [115,116], pointing out also a fundamental role in the biological behaviour of the bacteria.
The last fundamental quantity for the behaviour of the micro-organisms is related to the control of the environmental temperature. It is possible to point out that hyperpolarisation is generated in relation to the temperature difference between micro-organisms and their environment, showing the fundamental role of the environmental temperature, and in accordance with the experimental data [46,117,118,119]: the biomass concentrations results for ethanol 3.43 ± 0.08 g L 1 and acetate 0.93 ± 0.12 g L 1 Clostridium carboxidivorans at 37 C [120], higher than at 25 C, (ethanol 1.58 ± 0.03 g L 1 and acetate 0.61 ± 0.15 g L 1 [120]).
In order to obtain biofuels from micro-organisms, all the above mentioned quantities result fundamental and, one of the main issues is how to increase their production, by modifying the micro-organisms behaviour. In relation to our result (Relation (15)), we can point out that a possible way to improve the production of biofuels by micro-organisms can be obtained by increasing the ion and metabolite fluxes, in relation to the heat flux. Indeed, in nature, it is possible to observe that mutual interactions between different species can improve or modify the micro-organisms metabolism and their metabolites fluxes [121]. As a matter of fact, in nature, microbial consortia are widely diffuse, where living organisms create a community [122,123,124,125]. Thus, the coexistence of more species in the same environment can lead to a phenomenon called symbiosis [126,127]. The interaction between different species can bring to different effects for the two symbionts, that can be positive or negative [128]: mutualism, cooperation, commensalism, predation, parasitism, amensalism, and competition. In this context, a possible improvement in biofuels production from micro-organisms can be obtained by exploiting the capability of different species to positively interact among them. The main communication/interaction path, among the symbionts, is just the mutual exchange of metabolites [129,130,131,132,133,134,135,136].
In some biotechnological applications, the creation of artificial consortia among different species of micro-organisms through cocultures, has been positively applied, obtaining an improvement for their relative end-products production [137,138,139,140,141].

4. Discussion and Conclusions

Thermodynamics can improve the comprehension of the microorganisms growth and, consequently, it can allow us to improve their productivity of chemical compounds useful to biofuels production. The thermodynamic black box model has been proved to be a good tool for the evaluation of the microbial growth. In particular, it is useful to define the conditions of optimal growth, in relation to the interaction between the microorganisms and their environment. However, this interaction is based on fluxes through the microorgnasms membrane. Our results improve this approach, by analysing both the thermal and the electric fluxes. Indeed, the comprehension of the effects of the exogenous stimuli represents a fundamental improvement in the understanding of the bacterial electrophysiology [46,142,143].
So, biofuel and bioplastic production can be related to the proliferation, which is controlled just by membrane electric potential, driven by ATP synthesis [118,144,145,146]. The fundamental role of the membrane electric potential has been experimentally pointed out, by the analysis of the energy used by Escherichia coli to maintain its membrane electric potential; indeed, it results in around half of its total energy consumption [147].
Our results suggest explanation to some experimental evidence; indeed, an exogenous electrical stimulus has been shown to shape the proliferative capacity of bacteria, by inducing hyperpolarization in the cells [46].
Moreover, within the temperature range of life for the micro-organisms, higher temperature shortens the lag period and stimulates cell growth, in accordance with the microbial metabolism [148,149,150,151,152]. In our relation, the metabolism is taken into account by the surface temperature, while the environmental conditioning is expressed by the environmental temperature.
We highlight that a variation of the micro-organisms behaviour can be induced by the symbiosis between two different species. Indeed, in order to improve the formation of the useful high-value biomolecules, for biofuels production from micro-organisms, an interesting approach is to exploit the natural positive mutual interactions between different species. This could lead to a variation in their metabolites exchanges across their membranes and to a related change of the pH gradient [104]. For example, the cocultivation of two different species can bring to an enhancement of the lipid biomolecules inside the cells [153], which are the useful molecules employed for the biodiesel production [154]. Several studies have investigated the cocultivation of different micro-organisms, with the aim of improving their lipid concentration:
  • In Ref. [155], the species Chlorella pyrenoidosa and Rhodospiridium toruloides have been cocultivated, obtaining 4.60 g lipid L 1 (compared respectively to 3.00 g lipid L 1 and 3.40 g lipid L 1 , respectively for each single species);
  • In Ref. [156], the species Spirulina platensis and Rhodotorula glutinis have been cocultivated, obtaining 0.467 g lipid L 1 (compared respectively to 0.013 g lipid L 1 and 0.135 g lipid L 1 for the single species);
  • In Ref. [157], the species Chlorella sp. and Toluraspore have been cocultivated, obtaining 2.42 g lipid L 1 (compared respectively to 0.052 g lipid L 1 and 1.141 g lipid L 1 for the single species);
  • In Ref. [158], the species Chlorella sp. and Monoraphidium sp. have been cocultivated, obtaining an improved lipid productivity of 29.52 ± 1.13 mg lipid L 1 d 1 (compared respectively to 17.99 ± 3.39 mg lipid L 1 d 1 and 17.70 ± 1.19 mg lipid L 1 d 1 for the single species).
These examples represent some application of the mutualism to produce biodiesel; other promising consortia involve the production of biohydrogen [159], biomethane [160], and bioethanol [161,162]. Thus, the establishment of consortia, such as the microalgae-bacteria ones, can be a powerful tool to improve microalgal biomass production and to enhance the production of the high-value compounds useful to the biofuels production [163]. These processes involve a large amount of exchanged metabolites, molecular signals, and transporters, which induce a variation in the fluxes exchanged by the micro-organisms themselves, on their membrane potential and on the existing pH gradient.
Our results are in agreement with the recent open problems highlighted in the new frontiers in microbiological researches for industrial use of bacteria [142]. Indeed, Bacillus subtilis is an example of industrialised bacterium to hydrolyze polypeptides through its secreted proteases and to convert amino acids into advanced biofuels and ammonia fertilizer [164]. Just in relation to Bacillus subtilis, some experiments of stimulation of the membrane have been developed by using an electric stimulus of 60 mV μ m 1 AC 0.1 kHz for 2.5 s [46]: a hyperpolarization response has been shown, concluding that electrical stimulation causes the efflux of K + cations. The authors pointed out that Bacillus subtilis reacts to external stimuli by maintaining the resting-state membrane potential, but to do so it consumes a constant amount of ATP in order to keep the intracellular K + level. Moreover, the opening of voltage-gated K + channels, with the related hyperpolarization due to K + efflux, has an effect on the proliferative capacity of the cell [46]. These experimental results represent a further proof of our thermophysical results and considerations. Indeed, cell membranes contain enzymes complexes which further the oxidative phosphorylation along which electrical potential of H + ions, and chemical potential of reduced transporters (i.e., NADH, etc.), play a fundamental role in ATP production. Last, we can highlight that mutual interactions among different species can represent an interesting approach to improving the production of biofuels and must be considered an important topic of investigation for future development in optimisation of biofuels production.

Author Contributions

Conceptualization, U.L. and G.G.; methodology, U.L. and G.G.; software, G.G.; validation, U.L. and G.G.; formal analysis, U.L.; investigation, G.G.; resources, U.L.; data curation, G.G.; writing—original draft preparation, U.L. and G.G.; writing—review and editing, U.L. and G.G.; visualization, G.G.; supervision, U.L.; project administration, U.L.; funding acquisition, U.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Trifirò, F. Fuels from Biomass. Tec. Ital. Ital. J. Eng. Sci. 2019, 63, 86. [Google Scholar] [CrossRef]
  2. Rittmann, B. Opportunities for Renewable Bioenergy using Microorganisms. Biotechnol. Bioeng. 2008, 100, 203–212. [Google Scholar] [CrossRef]
  3. Gambelli, D.; Alberti, F.; Solfanelli, F.; Vairo, D.; Zanoli, R. Third generation algae biofuels in Italy by 2030: A scenario analysis using Bayesian networks. Energy Policy 2017, 103, 165–178. [Google Scholar] [CrossRef]
  4. Lucia, U.; Grisolia, G. Cyanobacteria and microalgae: Thermoeconomic considerations in biofuel production. Energies 2018, 11, 156. [Google Scholar] [CrossRef] [Green Version]
  5. Chisti, Y. Biodiesel from Microalgae. Biotechnol. Adv. 2007, 25, 294–306. [Google Scholar] [CrossRef]
  6. Saladini, F.; Patrizi, N.; Pulselli, F.M.; Marchettini, N.; Bastianoni, S. Guidelines for emergy evaluation of first, second and third generation biofuels. Renew. Sustain. Energy Rev. 2016, 66, 221–227. [Google Scholar] [CrossRef]
  7. Battista, F.; Mancini, G.; Ruggeri, B.; Fino, D. Selection of the best pretreatment for hydrogen and bioethanol production from olive oil waste products. Renew. Energy 2016, 88, 401–407. [Google Scholar] [CrossRef]
  8. Bensaid, S.; Conti, R.; Fino, D. Direct liquefaction of ligno-cellulosic residues for liquid fuel production. Fuel 2012, 94, 324–332. [Google Scholar] [CrossRef]
  9. Mata, T.; Martins, A.; Caetano, N. Microalgae for biodiesel production and other applications: A review. Renew. Sustain. Energy Rev. 2010, 14, 217–232. [Google Scholar] [CrossRef] [Green Version]
  10. Pols, A.; Spahn, A. Biofuels: Ethical aspects. In Encyclopedia of Food and Agricultural Ethics; Thompson, P.B., Kaplan, D.M., Eds.; Springer: Dordrecht, The Netherlands, 2014; pp. 211–217. [Google Scholar]
  11. Ziolkowska, J.R.; Simon, L. Recent developments and prospects for algae based fuels in the US. Renew. Sustain. Energy Rev. 2014, 29, 847–853. [Google Scholar] [CrossRef]
  12. Leong, W.H.; Zaine, S.N.A.; Ho, Y.C.; Uemura, Y.; Lam, M.K.; Khoo, K.S.; Kiatkittipong, W.; Cheng, C.K.; Show, P.L.; Lim, J.W. Impact of various microalgal-bacterial populations on municipal wastewater bioremediation and its energy feasibility for lipid-based biofuel production. J. Environ. Manag. 2019, 249, 109384. [Google Scholar] [CrossRef] [PubMed]
  13. Rodolfi, L.; Zittelli, G.C.; Bassi, N.; Padovani, G.; Biondi, N.; Bonini, G.; Tredici, M.R. Microalgae for oil: Strain selection, induction of lipid synthesis and outdoor mass cultivation in a low-cost photobioreactor. Biotechnol. Bioeng. 2009, 102, 100–112. [Google Scholar] [CrossRef]
  14. Correa, D.F.; Beyer, H.L.; Fargione, J.E.; Hill, J.D.; Possingham, H.P.; Thomas-Hall, S.R.; Schenka, P.M. Towards the implementation of sustainable biofuel production systems. Renew. Sustain. Energy Rev. 2019, 107, 250–263. [Google Scholar] [CrossRef]
  15. Chowdhury, H.; Loganathan, B. Third-generation biofuels from microalgae: A review. Curr. Opin. Green Sustain. Chem. 2019, 20, 39–44. [Google Scholar] [CrossRef]
  16. Alaswad, A.; Dassisti, M.; Prescott, T.; Olabia, A.G. Technologies and developments of third generation biofuel production. Renew. Sustain. Energy Rev. 2015, 51, 1446–1460. [Google Scholar] [CrossRef]
  17. Carriquiry, M.A.; Dub, X.; Timilsina, G.R. Second generation biofuels: Economics and policies. Energy Policy 2011, 39, 4222–4234. [Google Scholar] [CrossRef] [Green Version]
  18. Enamala, M.K.; Enamala, S.; Chavali, M.; Donepudi, J.; Yadavalli, R.; Kolapalli, B.; Aradhyula, T.V.; Velpuri, J.; Kuppam, C. Production of biofuels from microalgae - A review on cultivation, harvesting, lipid extraction, and numerous applications of microalgae. Renew. Sustain. Energy Rev. 2018, 94, 49–68. [Google Scholar] [CrossRef]
  19. Monod, J. The growth of bacterial cultures. Annu. Rev. Microbiol. 1949, 3, 371–394. [Google Scholar] [CrossRef] [Green Version]
  20. Monod, J. La Technique De Culture Continue: Theorie Et Applications; Masson: Paris, France, 1950. [Google Scholar]
  21. Blackman, F. Optima and Limiting Factors. Ann. Bot. 1905, 19, 281–296. [Google Scholar] [CrossRef]
  22. Tessier, G. Croissance des populations bactériennes et quantité d’aliment disponible. Revue Sci. Tech. 1942, 3208, 209–214. [Google Scholar]
  23. Esener, A.A.; Roels, J.A.; Kossen, N.W.F. Theory and applications of unstructured growth models: Kinetic and energetic aspects. Biotechnol. Bioeng. 1983, 25, 2803–2841. [Google Scholar] [CrossRef]
  24. Saadat, N.P.; Nies, T.; Rousset, Y.; Ebenhöh, O. Thermodynamic Limits and Optimality of Microbial Growth. Entropy 2020, 22, 277. [Google Scholar] [CrossRef] [Green Version]
  25. Herbert, D.; Elsworth, R.; Telling, R.C. The Continuous Culture of Bacteria: A Theoretical and Experimental Study. J. Gen. Microbiol. 1956, 14, 601–622. [Google Scholar] [CrossRef] [Green Version]
  26. Pirt, S. The maintenance energy of bacteria in growing cultures. Proc. R. Soc. Lond. B Biol. Sci. 1965, 163, 224–231. [Google Scholar] [CrossRef]
  27. van Bodegom, P. Microbial Maintenance: A Critical Review on Its Quantification. Microb. Ecol. 2007, 53, 513–523. [Google Scholar] [CrossRef] [Green Version]
  28. Mayberry, W.R.; Prochazka, G.J.; Payne, W.J. Factors derived from studies of aerobic growth in minimal media. J. Bacteriol. 1968, 96, 1424–1426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Thornton, W. XV. The relation of oxygen to the heat of combustion of organic compounds. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1917, 66, 196–203. [Google Scholar] [CrossRef]
  30. Musa, M.; Ayoko, G.A.; Ward, A.; Rösch, C.; Brown, R.J.; Rainey, T.J. Factors Affecting Microalgae Production for Biofuels and the Potentials of Chemometric Methods in Assessing and Optimizing Productivity. Cells 2019, 8, 851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Saravanan, A.P.; Mathimani, T.; Deviram, G.; Rajendran, K.; Pugazhendhi, A. Biofuel policy in India: A review of policy barriers in sustainable marketing of biofuel. J. Clean. Prod. 2018, 193, 734–747. [Google Scholar] [CrossRef]
  32. Su, Y.; Zhang, P.; Su, Y. An overview of biofuels policies and industrialization in the major biofuel producing countries. Renew. Sustain. Energy Rev. 2015, 50, 991–1003. [Google Scholar] [CrossRef]
  33. Grisolia, G.; Fino, D.; Lucia, U. Thermodynamic optimisation of the biofuel production based on mutualism. Energy Rep. 2020, 6, 1561–1571. [Google Scholar] [CrossRef]
  34. Santos, C.; Reis, A. Microalgal symbiosis in biotechnology. Appl. Microbiol. Biotechnol. 2014, 98, 5839–5846. [Google Scholar] [CrossRef] [Green Version]
  35. Shurin, J.B.; Abbott, R.L.; Deal, M.S.; Kwan, G.T.; Litchman, E.; McBride, R.C.; Mandal, S.; Smith, V.H. Industrial-strength ecology: Trade-offs and opportunities in algal biofuel production. Ecol. Lett. 2013, 16, 1393–1404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Kazamia, E.; Riseley, A.S.; Howe, C.J.; Smith, A.G. An Engineered Community Approach for Industrial Cultivation of Microalgae. Ind. Biotechnol. 2014, 10, 184–190. [Google Scholar] [CrossRef] [Green Version]
  37. Cooper, M.B.; Smith, A.G. Exploring mutualistic interactions between microalgae and bacteria in the omics age. Curr. Opin. Plant Biol. 2015, 26, 147–153. [Google Scholar] [CrossRef] [Green Version]
  38. Heimann, K. Novel approaches to microalgal and cyanobacterial cultivation for bioenergy and biofuel production. Curr. Opin. Biotechnol. 2016, 38, 183–189. [Google Scholar] [CrossRef]
  39. Cho, D.; Ramanan, R.; Heo, J.; Lee, J.; Kim, B.; Oh, H. Enhancing microalgal biomass productivity by engineeringa microalgal–bacterial community. Bioresour. Technol. 2015, 175, 578–585. [Google Scholar] [CrossRef]
  40. Ramanan, R.; Kim, B.; Cho, D.; Kim, H. Algae-bacteria interactions: Evolution, ecology and emerging applications. Biotechnol. Adv. 2016, 34, 14–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Lucia, U.; Grisolia, G. Unavailability percentage as energy planning and economic choice parameter. Renew. Sustain. Energy Rev. 2017, 75, 197–204. [Google Scholar] [CrossRef]
  42. Prindle, A.; Liu, J.; Asally, M.; Ly, S.; Garcia-Ojalvo, J.; Süel, G.M. Ion channels enable electrical communication in bacterial communities. Nature 2015, 527, 59–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Sirec, T.; Buffard, P.; Garcia-Ojalvo, J.; Asally, M. Electrical-charge accumulation enables integrative quality control during B. subtilis sporulation. iScience 2019, 16, 378–389. [Google Scholar] [CrossRef] [Green Version]
  44. McCaig, C.D.; Rajnicek, A.M.; Song, B.; Zhao, M. Controlling cell behavior electrically: Current views and future potential. Physiol. Rev. 2005, 85, 943–978. [Google Scholar] [CrossRef] [Green Version]
  45. Levin, M. Molecular bioelectricity: How endogenous voltage potentials control cell behavior and instruct pattern regulation in vivo. Mol. Biol. Cell. 2014, 25, 3835–3850. [Google Scholar] [CrossRef]
  46. Stratford, J.P.; Edwards, C.L.A.; Ghanshyam, M.J.; Malyshev, D.; Delise, M.A.; Hayashi, Y.; Asally, M. Electrically induced bacterial membrane-potential dynamics correspond to cellular proliferation capacity. Proc. Natl. Acad. Sci. USA 2019, 116, 9552–9557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Arroyo, M.; Walani, N.; Torres-Sánchez, A.; Kaurin, D. Onsager’s Variational Principle in Soft Matter: Introduction and Application to the Dynamics of Adsorption of Proteins onto Fluid Membranes. In The Role of Mechanics in the Study of Lipid Bilayers; Steigmann, D.J., Ed.; Springer International Publishing: Berlin, Germany, 2018. [Google Scholar]
  48. Lucia, U.; Grisolia, G. How Life Works—A Continuous Seebeck-Peltier Transition in Cell Membrane? Entropy 2020, 22, 960. [Google Scholar] [CrossRef] [PubMed]
  49. Lucia, U.; Grisolia, G. Non-equilibrium thermodynamic approach to Ca2+-fluxes in cancer. Appl. Sci. 2020, 10, 6737. [Google Scholar] [CrossRef]
  50. Lucia, U. Bioengineering thermodynamics: An engineering science for thermodynamics of biosystems. Int. J. Thermodyn. 2015, 18, 254–265. [Google Scholar] [CrossRef] [Green Version]
  51. Lucia, U.; Grisolia, G. Thermal Resonance and Cell Behavior. Entropy 2020, 22, 774. [Google Scholar] [CrossRef]
  52. Lucia, U.; Grisolia, G. Resonance in Thermal Fluxes Through Cancer Membrane. Atti Dell’Accad. Pelorit. Pericol. 2020, 98, SC1–SC6. [Google Scholar] [CrossRef]
  53. Lucia, U.; Grisolia, G. Thermal Physics and Glaucoma: From Thermodynamic to Biophysical Considerations to Designing Future Therapies. Appl. Sci. 2020, 10, 7071. [Google Scholar] [CrossRef]
  54. Yang, M.; Brackenbury, W.J. Membrane potential and cancer progression. Front. Physiol. 2013, 4, 185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Sundelacruz, S.; Levin, M.; Kaplan, D.L. Role of the membrane potential in the regulation of cell proliferation and differentiation. Stem Cell Rev. 2009, 5, 231–246. [Google Scholar] [CrossRef]
  56. Lobikin, M.; Chernet, B.; Lobo, D.; Levin, M. Resting potential, oncogene-induced tumorigenesis, and metastasis: The bioelectric basis of cancer in vivo. Phys. Biol. 2012, 9, 065002. [Google Scholar] [CrossRef] [Green Version]
  57. Schwab, A.; Fabian, A.; Hanley, P.J.; Stock, C. Role of the ion channels and transporters in cell migration. Physiol. Rev. 2012, 92, 1865–1913. [Google Scholar] [CrossRef]
  58. Ambrose, E.J.; James, A.M.; Lowick, J.H. Differences between the electrical charge carried by normal and homologous tumour cells. Nature 1956, 177, 576–577. [Google Scholar] [CrossRef]
  59. Cone, C.D. Electroosmotic interactions accompanying mitosis initiation in sarcoma cells in vitro. Trans. N. Y. Acad. Sci. 1969, 31, 404–427. [Google Scholar] [CrossRef] [PubMed]
  60. Cone, C.D. Variation of the transmembrane potential level as a basic mechanism of mitosis control. Oncology 1970, 24, 438–470. [Google Scholar] [CrossRef] [PubMed]
  61. Cone, C.D. Unified theory on the basic mechanism of normal mitotic control and oncogenesis. J. Theor. Biol. 1971, 30, 151–181. [Google Scholar] [CrossRef]
  62. Tokuoka, S.; Marioka, H. The membrane potential of the human cancer and related cells (I). Gann 1957, 48, 353–354. [Google Scholar] [PubMed]
  63. Altman, P.L.; Katz, D. Biological Handbook Vol. 1: Cell Biology; Federation of American Society for Experimental Biology: Bethesda, MD, USA, 1976. [Google Scholar]
  64. Balitsky, K.P.; Shuba, E.P. Resting potential of malignant tumour cells. Acta Unio. Int. Contra Cancrum 1964, 20, 1391–1393. [Google Scholar] [PubMed]
  65. Jamakosmanovic, A.; Loewenstein, W. Intracellular communication and tissue growth. III. Thyroid cancer. J. Cell Biol. 1968, 38, 556–561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Rizzuto, R.; Marchi, S.; Bonora, M.; Aguiari, P.; Bononi, A.; Stefani, D.D.; Giorgi, C.; Leo, S.; Rimessi, A.; Siviero, R.; et al. Ca(2+) transfer from the ER to mitochondria: When, how and why. Biochim. Biophys. Acta 2009, 1787, 1342–1351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Giorgi, C.; Missiroli, S.; Patergnani, S.; Duszynski, J.; Wieckowski, M.R.; Pinton, P. Mitochondria-associated membranes: Composition, molecular mechanisms, and physiopathological implications. Antioxid. Redox Signal. 2015, 22, 995–1019. [Google Scholar] [CrossRef]
  69. Pinton, P.; Ferrari, D.; Rapizzi, E.; Virgilio, F.D.; Pozzan, T.; Rizzuto, R. The Ca2+ concentration of the endoplasmic reticulum is a key determinant of ceramide-induced apoptosis: Significance for the molecular mechanism of Bcl-2 action. EMBO 2001, 20, 2690–2701. [Google Scholar] [CrossRef] [PubMed]
  70. Pinton, P.; Ferrari, D.; Magalhaes, P.; Schulze-Osthoff, K.; Virgilio, F.D.; Pozzan, T.; Rizzuto, R. Reduced loading of intracellular Ca(2+) stores and downregulationof capacitative Ca(2+) influx in Bcl-2-overexpressing cells. J. Cell Biol. 2000, 148, 857–862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Foyouzi-Youssefi, R.; Arnaudeau, S.; Borner, C.; Kelley, W.L.; Tschopp, J.; Lew, D.P.; Demaurex, N.; Krause, K.H. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 2000, 97, 5723–5728. [Google Scholar] [CrossRef] [Green Version]
  72. Akl, H.; Vervloessem, T.; Kiviluoto, S.; Bittremieux, M.; Parys, J.B.; Smedt, H.D.; Bultynck, G. A dual role for the anti-apoptotic Bcl-2 protein in cancer: Mitochondria versus endoplasmic reticulum. Biochim. Biophys. Acta 2014, 1843, 2240–2252. [Google Scholar] [CrossRef] [Green Version]
  73. Akl, H.; Bultynck, G. Altered Ca(2+) signaling in cancer cells: Proto-oncogenes and tumor suppressors targeting IP3 receptors. Biochim. Biophys. Acta 2013, 1835, 180–193. [Google Scholar]
  74. Marchi, S.; Marinello, M.; Bononi, A.; Bonora, M.; Giorgi, C.; Rimessi, A.; Pinton, P. Selective modulation of subtype III IP(3)R by Akt regulates ER Ca2+ releaseand apoptosis. Cell Death Dis. 2012, 3, e304. [Google Scholar] [CrossRef] [Green Version]
  75. Giorgi, C.; Ito, K.; Lin, H.K.; Santangelo, C.; Wieckowski, M.R.; Lebiedzinska, M.; Bononi, A.; Bonora, M.; Duszynski, J.; Bernardi, R.; et al. PML regulates apoptosis at endoplasmic reticulum by modulating calcium release. Science 2019, 330, 1247–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Stewart, T.A.; Yapa, K.T.; Monteith, G.R. Altered calcium signaling in cancer cells. Biochim. Biophys. Acta 2015, 1848, 2502–2511. [Google Scholar] [CrossRef] [Green Version]
  77. Bononi, A.; Bonora, M.; Marchi, S.; Missiroli, S.; Poletti, F.; Giorgi, C.; Pandolfi, P.P.; Pinton, P. Identification of PTEN at the ER and MAMs and its regulation of Ca2+ signaling and apoptosis in a protein phosphatase-dependent manner. Cell Death Differ. 2013, 20, 1631–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Giorgi, C.; Bonora, M.; Sorrentino, G.; Missiroli, S.; Poletti, F.; Suski, J.M.; Galindo Ramirez, F.; Rizzuto, R.; Di Virgilio, F.; Zito, E.; et al. p53 at the endoplasmic reticulum regulates apoptosis in a Ca2+-dependent manner. Proc. Natl. Acad. Sci. USA 2015, 112, 1779–1784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Giorgi, C.; Bonora, M.; Missiroli, S.; Poletti, F.; Ramirez, F.G.; Morciano, G.; Morganti, C.; Pandolfi, P.P.; Mammano, F.; Pinton, P. Intravital imaging reveals p53-dependent cancer cell death induced by phototherapy via calcium signaling. Oncotarget 2015, 6, 1435–1445. [Google Scholar] [CrossRef] [Green Version]
  80. Rimessi, A.; Marchi, S.; Patergnani, S.; Pinton, P. H-Ras-driven tumoral main-tenance is sustained through caveolin-1-dependent alterations in calcium signaling. Oncogene 2014, 33, 2329–2340. [Google Scholar] [CrossRef] [PubMed]
  81. Parsadaniantz, S.M.; le Goazigo, A.R.; Sapienza, A.; Habas, C.; Baudouin, C. Glaucoma: A Degenerative Optic Neuropathy Related to Neuroinflammation? Cells 2020, 9, 535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Soto, I.; Howell, G.R. The complex role of neuroinflammation in glaucoma. Cold Spring Harb. Perspect. Med. 2014, 4, a017269. [Google Scholar] [CrossRef]
  83. Tezel, G. Immune regulation toward immunomodulation for neuroprotection in glaucoma. Curr. Opin. Pharmacol. 2013, 13, 23–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Wax, M.B.; Tezel, G.; Yang, J.; Peng, G.; Patil, R.V.; Agarwal, N.; Sappington, R.M.; Calkins, D.J. Induced autoimmunity to heat shock proteins elicits glaucomatous loss of retinal ganglion cell neurons via activated T-cell-derived fas-ligand. J. Neurosci. 2008, 28, 12085–12096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gupta, N.; Yucel, Y.H. Glaucoma as a neurodegenerative disease. Curr. Opin. Ophthalmol. 2007, 18, 110–114. [Google Scholar] [CrossRef] [PubMed]
  86. Yucel, Y.H.; Gupta, N. Glaucoma of the brain: A disease model for the study of transsynaptic neural degeneration. Prog. Brain Res. 2008, 173, 465–478. [Google Scholar] [CrossRef]
  87. Imamura, K.; Onoe, H.; Shimazawa, M.; Wada, S.N.Y.; Kato, K.; Nakajima, H.; Mizuma, H.; Onoe, K.; Taniguchi, T.; Sasaoka, M.; et al. Molecular imaging reveals unique degenerative changes in experimental glaucoma. Neuroreport 2009, 20, 139–144. [Google Scholar] [CrossRef] [PubMed]
  88. Shum, J.W.H.; Liu, K.; So, K.F. The progress in optic nerve regeneration, where are we? Neural Regen. Res. 2016, 11, 32–36. [Google Scholar] [CrossRef]
  89. Bejan, A. The golden ratio predicted: Vision, cognition and locomotion as a single designin nature. Int. J. Des. Nat. Ecodyn. 2009, 4, 97–104. [Google Scholar] [CrossRef]
  90. Yourgrau, W.; van der Merwe, A.; Raw, G. Treatise on Irreversible and Statistical Thermophysics; Dover: New York, NY, USA, 1982. [Google Scholar]
  91. Callen, H.B. Thermodynamics; Wiley: New York, NY, USA, 1960. [Google Scholar]
  92. Goupil, C. Thermodynamics of Thermoelectricity. In Thermodynamics; Mizutani, T., Ed.; IntechOpen: Shanghai, China, 2011. [Google Scholar]
  93. Goupil, C.; Seifert, W.; Zabrocki, K.; Müller, E.; Snyder, G.J. Thermodynamics of Thermoelectric Phenomena and Applications. Entropy 2011, 13, 1481–1517. [Google Scholar] [CrossRef] [Green Version]
  94. Katchalsky, A.; Currant, P.F. Nonequilibrium Thermodynamics in Biophysics; Harvard University Press: Boston, MA, USA, 1965. [Google Scholar]
  95. Degroot, S.R.; Mazur, P. Non-Equilibrium Thermodynamics; North-Holland Publishing Company: Amsterdam, The Netherlands, 1962. [Google Scholar]
  96. Lucia, U.; Grazzini, G. Global analysis of dissipations due to irreversibility. Revue Gen. Therm. 1997, 36, 605–609. [Google Scholar] [CrossRef]
  97. López, A.; Molina-Aiz, F.D.; Valera, D.L.; Peña, A. Determining the emissivity of the leaves of nine horticultural crops by means of infrared thermography. Sci. Hortic. 2012, 137, 49–58. [Google Scholar] [CrossRef]
  98. Lam, O.; Wheeler, J.; Tang, C.M. Thermal control of virulence factors in bacteria: A hot topic. Virulence 2014, 5, 852–862. [Google Scholar] [CrossRef] [Green Version]
  99. Couradeau, E.; Karaoz, U.; Lim, H.C.; da Rocha, U.N.; Northen, T.; Brodie, E.; Garcia-Pichel, F. Bacteria increase arid-land soil surface temperature through the production of sunscreens. Nat. Commun. 2016, 7, 10373. [Google Scholar] [CrossRef] [Green Version]
  100. Apostol, T.S. Calculus. Volume 2: Multi-Variable Calculus and Linear Algebra with Applications to Differential Equations and Probability; Wiley: Hoboken, NJ, USA, 1969. [Google Scholar]
  101. Grabe, M.; Wang, H.; Oster, G. The mechanochemistry of V-ATPase proton pumps. Biophs. J. 2000, 78, 2798–2813. [Google Scholar] [CrossRef] [Green Version]
  102. Lucia, U.; Grisolia, G. Second law efficiency for living cells. Front. Biosci. 2017, 9, 270–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Sorre, B.; Callan-Jones, A.; Manneville, J.B.; Nassoy, P.; Joanny, J.F.; Prost, J.; Goud, B.; Bassereau, P. Curvature-driven lipid sorting needs proximity to a demixing point and is aided by proteins. Proc. Natl. Acad. Sci. USA 2009, 106, 5622–5626. [Google Scholar] [CrossRef] [Green Version]
  104. Dianursanti, D.; Santoso, A. Increasing Lipid Accumulation of Chlorella vulgaris using Spirulina platensis in Flat Plate Reactor for Synthesizing Biodiesel. Energy Procedia 2015, 65, 58–66. [Google Scholar] [CrossRef] [Green Version]
  105. Nakanishi-Matsui, M.; Sekiya, M.; Futai, R.K.N.M. The mechanism of rotating proton pumping ATPases. BBA Bioenerg. 2010, 1797, 1343–1352. [Google Scholar] [CrossRef] [Green Version]
  106. Stevens, T.H.; Forgac, M. Structure, function and regulation of the vacuolar (H+)-ATPase. Annu. Rev. Cell. Dev. Biol. 1997, 13, 779–808. [Google Scholar] [CrossRef]
  107. Tuszynski, J.A.; Kurzynski, M. Introduction to Molecular Biophysics; CRC Press: Boca Raton, FL, USA, 2003; pp. 383–392. [Google Scholar]
  108. Rudolph, M.G.; Stanfield, R.L.; Wilson, I.A. How TCRs bind MHCs, peptides, and coreceptors. Annu. Rev. Immunol. 2006, 24, 419–466. [Google Scholar] [CrossRef]
  109. Strong, R.K. Asymmetric ligand recognition by the activating natural killer cell receptor NKG2D, a symmetric homodimer. Mol. Immunol. 2002, 38, 1029–1037. [Google Scholar] [CrossRef]
  110. Ardito, F.; Giuliani, M.; Perrone, D.; Troiano, G.; Muzio, L. The crucial role of protein phosphorylation in cell signaling and its use as targeted therapy. Int. J. Mol. Med. 2017, 40, 271–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Lucia, U.; Grisolia, G.; Dolcino, D.; Astori, M.R.; Massa, E.; Ponzetto, A. Constructal approach to bio-engineering: The ocular anterior chamber temperature. Sci. Rep. 2016, 6, 31099. [Google Scholar] [CrossRef]
  112. Lucia, U.; Grisolia, G.; Astori, M.R. Constructal law analysis of Cl transport in eyes aqueous humor. Sci. Rep. 2017, 7, 6856. [Google Scholar] [CrossRef] [Green Version]
  113. Lucia, U.; Grisolia, G.; Francia, S.; Astori, M.R. Theoretical biophysical approach to cross-linking effects on eyes pressure. Physica A 2019, 534, 122163. [Google Scholar] [CrossRef]
  114. Nag, P.K. Heat and Mass Transfer; Mcgraw Hill Education: Noida, India, 2011. [Google Scholar]
  115. Shi, Z.; Baumgart, T. Membrane tension and peripheral protein density mediate membrane shape transitions. Nat. Commun. 2015, 6, 5974. [Google Scholar] [CrossRef] [PubMed]
  116. Morris, C.E.; Homann, U. Cell surface area regulation and membrane tension. J. Memb. Biol. 2001, 179, 79–102. [Google Scholar] [CrossRef]
  117. Felle, H.; Porter, J.S.; Slayman, C.L.; Kaback, H.R. Quantitative measurements of membrane potential in Escherichia coli. Biochemistry 1980, 19, 3585–3590. [Google Scholar] [CrossRef]
  118. Ramos, S.; Schuldiner, S.; Kaback, H.R. The electrochemical gradient of protons and its relationship to active transport in Escherichia coli membrane vesicles. Proc. Natl. Acad. Sci. USA 1976, 73, 1892–1896. [Google Scholar] [CrossRef] [Green Version]
  119. Zheng, J.; Trudeau, M.C. Handbook of Ion Channels; CRC Press: Boca Raton, FL, USA, 2015. [Google Scholar]
  120. Shen, S.; Wang, G.; Zhang, M.; Tang, Y.; Gu, Y.; Jiang, W.; Wang, Y.; Zhuang, Y. Effect of temperature and surfactant on biomass growth and higher-alcohol production during syngas fermentation by Clostridium carboxidivorans P7. Bioresour. Bioprocess. 2020, 7, 56. [Google Scholar] [CrossRef]
  121. Padmaperuma, G.; Kapoore, R.V.; Gilmour, D.J.; Vaidyanathan, S. Microbial consortia: A critical look at microalgae co-cultures for enhanced biomanufacturing. Crit. Rev. Biotechnol. 2018, 35, 690–703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Stanley, G.D. Photosymbiosis and the Evolution of Modern Coral Reefs. Science 2006, 312, 857–858. [Google Scholar] [CrossRef] [Green Version]
  123. Kent, A.D.; Triplett, E.W. Microbial communities and their interactions in soil and rhizosphere ecosystems. Annu. Rev. Microbiol. 2002, 56, 211–236. [Google Scholar] [CrossRef] [PubMed]
  124. Johansson, J.F.; Paul, L.R.; Finlay, R.D. Microbial interactions in the mycorrhizosphere and their significance for sustainable agriculture. FEMS Microbiol. Ecol. 2004, 48, 1–13. [Google Scholar] [CrossRef] [PubMed]
  125. Singh, B.K.; Millard, P.; Whiteley, A.S.; Murrel, J.C. Unravelling rhizosphere-microbial interactions: Opportunities and limitations. Trends Microbiol. 2004, 12, 386–393. [Google Scholar] [CrossRef]
  126. Oulhen, N.; Schulz, B.J.; Carrier, T.J. English translation of Heinrich Anton de Bary’s 1878 speech, Die Erscheinung der Symbiose. Symbiosis 2016, 69, 131–139. [Google Scholar] [CrossRef]
  127. Paracer, S.; Ahmadjian, V. Symbiosis: An Introduction to Biological Associations, 2nd ed.; Oxford University Press: New York, NY, USA, 2000. [Google Scholar]
  128. Willey, J.; Sherwood, L.; Woolverton, C. Prescott’s Microbiology, 9th ed.; McGraw-Hill Education: New York, NY, USA, 2013. [Google Scholar]
  129. Delaux, P.M.; Radhakrishnan, G.V.; Jayaraman, D.; Cheema, J.; Malbreil, M.; Volkening, J.D.; Sekimoto, H.; Nishiyama, T.; Melkonian, M.; Pokorny, L.; et al. Algal ancestor of land plants was preadapted for symbiosis. Proc. Natl. Acad. Sci. USA 2015, 112, 13390–13395. [Google Scholar] [CrossRef] [Green Version]
  130. Volk, R.-B.; Furkert, F.H. Antialgal, antibacterial and antifungal activity of two metabolites produced and excreted by cyanobacteria during growth. Microbiol. Res. 2006, 61, 180–186. [Google Scholar] [CrossRef] [PubMed]
  131. Dashti, Y.; Grkovic, T.; Abdelmohsen, U.R.; Hentschel, U.; Quinn, R.J. Production of Induced Secondary Metabolites by a Co-Culture of Sponge-Associated Actinomycetes, Actinokineospora sp. EG49 and Nocardiopsis sp. RV163. Mar. Drugs 2014, 12, 3046–3059. [Google Scholar] [CrossRef]
  132. González, J.E.; Keshavan, N.D. Messing with bacterial quorum sensing. Microbiol. Mol. Biol. Rev. 2006, 70, 559–875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Hooshangi, S.; Bentley, W.E. From unicellular properties to multicellular behavior: Bacteria quorum sensing circuitry and applications. Curr. Opin. Biotechnol. 2008, 19, 550–555. [Google Scholar] [CrossRef]
  134. March, J.C.; Bentley, W.E. Quorum sensing and bacterial cross-talk in biotechnology. Curr. Opin. Biotechnol. 2004, 15, 495–502. [Google Scholar] [CrossRef]
  135. Shank, E.A.; Kolter, R. New developments in microbial interspecies signaling. Curr. Opin. Biotechnol. 2009, 12, 205–214. [Google Scholar] [CrossRef] [Green Version]
  136. Rateb, M.E.; Hallyburton, I.; Houssen, W.E.; Bull, A.T.; Goodfellow, M.; Santhanam, R.; Jaspars, M.; Ebel, R. Induction of diverse secondary metabolites in Aspergillus fumigatus by microbial co-culture. R. Soc. Chem. 2013, 3, 14444–14450. [Google Scholar] [CrossRef] [Green Version]
  137. Bernstein, H.C.; Carlson, R.P. Microbial Consortia Engineering for Cellular Factories: In vitro to in silico systems. Comput. Struct. Biotechnol. J. 2012, 3, e201210017. [Google Scholar] [CrossRef] [Green Version]
  138. Gebreslassie, B.H.; Waymire, R.; You, F. Sustainable design and synthesis of algae-based biorefinery for simultaneous hydrocarbon biofuel production and carbon sequestration. Am. Inst. Chem. Eng. J. 2013, 59, 1599–1621. [Google Scholar] [CrossRef]
  139. Trzcinski, A.-P.; Hernandez, E.; Webb, C. A novel process for enhancing oil production in algae biorefineries through bioconversion of solid by-products. Bioresour. Technol. 2012, 116, 295–301. [Google Scholar] [CrossRef] [PubMed]
  140. Cheirsilp, B.; Suwannarat, W.; Niyomdecha, R. Mixed culture of oleaginous yeast Rhodotorula glutinis and microalga Chlorella vulgaris for lipid production from industrial wastes and its use as biodiesel feedstock. New Biotechnol. 2011, 28, 362–368. [Google Scholar] [CrossRef] [PubMed]
  141. Kitcha, S.; Cheirslip, B. Enhanced lipid production by co-cultivation and co-encapsulation of oleaginous yeast Trichosporonoides spathulata with microalgae in alginate gel beads. Appl. Biochem. Biotechnol. 2014, 173, 522–534. [Google Scholar] [CrossRef] [PubMed]
  142. Benarroch, J.M.; Asally, M. The Microbiologist’s Guide to Membrane Potential Dynamics. Trends Microbiol. 2020, 28, 304–314. [Google Scholar] [CrossRef] [Green Version]
  143. Schloss, A.C.; Liu, W.; Williams, D.M.; Kaufman, G.; Hendrickson, H.P.; Rudshteyn, B.; Fu, L.; Wang, H.; Batista, V.S.; Osuji, C.; et al. Fabrication of modularly functionalizable microcapsules using protein-based technologies. ACS Biomater. Sci. Eng. 2016, 2, 1856–1861. [Google Scholar] [CrossRef] [Green Version]
  144. Kotnik, T.; Miklavcic, D. Analytical description of transmembrane voltage induced by electric fields on spheroidal cells. Biophys. J. 2000, 79, 670–679. [Google Scholar] [CrossRef] [Green Version]
  145. Kaim, G.; Dimroth, P. ATP synthesis by F-type ATP synthase is obligatorily dependent on the transmembrane voltage. EMBO J. 1999, 18, 4118–4127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Strahl, H.; Hamoen, L.W. Membrane potential is important for bacterial cell division. Proc. Natl. Acad. Sci. USA 2010, 107, 12281–12286. [Google Scholar] [CrossRef] [Green Version]
  147. Milo, R.; Phillips, R. Cell Biology by the Numbers; Garland Science: New York, NY, USA, 2015. [Google Scholar]
  148. von Stockar, U. Biothermodynamics of live cells: A tool for biotechnology and biochemical engineering. J. Non-Equilib. Thermodyn. 2010, 35, 415–475. [Google Scholar] [CrossRef] [Green Version]
  149. Krabe, K.; van Weze, J. Improved derivation of phosphate potentials at different temperatures. Biochim. Biophys. Acta 1992, 1098, 172–176. [Google Scholar] [CrossRef]
  150. Harold, F.M. Ion currents and physiological functions in microorganisms. Annu. Rev. Microbiol. 1977, 31, 181–203. [Google Scholar] [CrossRef]
  151. Angelova, M.I.; Bitbol, A.F.; Seigneuret, M.; Staneva, G.; Kodama, A.; Sakuma, Y.; Kawakatsu, T.; Imai, M.; Puff, N. pH sensing by lipids in membranes: The fundamentals of pH-driven migration, polarization and deformations of lipid bilayer assemblies. BBA Biomemb. 2018, 1860, 2042–2063. [Google Scholar] [CrossRef]
  152. Price, P.B.; Sowers, T. Temperature dependence of metabolic rates for microbial growth, maintenance, and survival. Proc. Natl. Acad. Sci. USA 2004, 101, 4631–4636. [Google Scholar] [CrossRef] [Green Version]
  153. Rashid, N.; Ryu, A.J.; Jeong, K.J.; Lee, B.; Chang, Y.K. Co-cultivation of two freshwater microalgae species to improve biomass productivity and biodiesel production. Energy Convers. Manag. 2019, 196, 640–648. [Google Scholar] [CrossRef]
  154. Scott, S.A.; Davey, M.P.; Dennis, J.S.; Horst, I.; Howe, C.J.; Lea-Smith, D.J.; Smith, A.G. Biodiesel from algae: Challenges and prospects. Curr. Opin. Biotechnol. 2010, 21, 277–286. [Google Scholar] [CrossRef] [PubMed]
  155. Ling, J.; Nip, S.; Cheok, W.L.; de Toledo, R.A.; Shim, H. Lipid production by a mixed culture of oleaginous yeast and microalga from distillery and domestic mixed wastewater. Bioresour. Technol. 2014, 173, 132–139. [Google Scholar] [CrossRef]
  156. Xue, F.; Miao, J.; Zhang, X.; Tan, T. A new strategy for lipid production by mix cultivation of Spirulina platensis and Rhodotorula glutinis. Appl. Biochem. Biotechnol. 2010, 160, 498–503. [Google Scholar] [CrossRef] [PubMed]
  157. Papone, T.; Kookkhunthod, S.; Leesing, R. Microbial Oil Production by Monoculture and Mixed Cultures of Microalgae and Oleaginous Yeasts using Sugarcane Juice as Substrate. World Acad. Sci. Eng. Technol. Int. J. Nutr. Food Eng. 2012, 6, 195–199. [Google Scholar] [CrossRef]
  158. Zhao, P.; Yu, X.; Li, J.; Tang, X.; Huang, Z. Enhancing lipid productivity by co-cultivation of Chlorella sp. U4341 and Monoraphidium sp. FXY-10. J. Biosci. Bioeng. 2014, 118, 72–77. [Google Scholar] [CrossRef]
  159. Lakatos, G.; Deák, Z.; Vass, I.; Rétflvi, T.; Rozgonyi, S.; Rákhely, G.; Ördög, V.; Kondorosi, E.; Maróti, G. Bacterial symbionts enhance photo-fermentative hydrogen evolution of Chlamydomonas algae. Green Chem. 2014, 16, 4716–4727. [Google Scholar] [CrossRef]
  160. Muñoz, C.M.; Hidalgo, C.; Zapata, M.; Jeison, D.; Riquelme, C.; Rivas, M. Use of cellulolytic marine bacteria for enzymatic pretreatment in microalgal biogas production. Appl. Environ. Microbiol. 2014, 80, 4199–4206. [Google Scholar] [CrossRef] [Green Version]
  161. Matsumoto, M.; Yokouchi, H.; Zuzuki, N.; Ohata, H.; Matsunaga, T. Saccharification of marine microalgae using marine bacteria for ethanol production. Appl. Biochem. Biotechnol. 2003, 105, 247–254. [Google Scholar] [CrossRef]
  162. Silva, C.E.F.; Bertucco, S. Bioethanol from microalgae and cyanobacteria: A review and technological outlook. Process Biochem. 2016, 51, 1833–1842. [Google Scholar] [CrossRef]
  163. Yao, S.; Liu, S.; An, Y.; Lu, J.; Gjermansen, C.; Schramm, A. Microalgae–bacteria symbiosis in microalgal growth and biofuel production: A review. J. Appl. Microbiol. 2018, 126, 359–368. [Google Scholar] [CrossRef] [PubMed]
  164. Choi, K.W.; Wernick, D.G.; Tat, C.A.; Liao, J.C. Consolidated conversion of protein waste into biofuels and ammonia using Bacillus subtilis. Metab. Eng. 2014, 23, 53–61. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lucia, U.; Grisolia, G. Biofuels from Micro-Organisms: Thermodynamic Considerations on the Role of Electrochemical Potential on Micro-Organisms Growth. Appl. Sci. 2021, 11, 2591. https://doi.org/10.3390/app11062591

AMA Style

Lucia U, Grisolia G. Biofuels from Micro-Organisms: Thermodynamic Considerations on the Role of Electrochemical Potential on Micro-Organisms Growth. Applied Sciences. 2021; 11(6):2591. https://doi.org/10.3390/app11062591

Chicago/Turabian Style

Lucia, Umberto, and Giulia Grisolia. 2021. "Biofuels from Micro-Organisms: Thermodynamic Considerations on the Role of Electrochemical Potential on Micro-Organisms Growth" Applied Sciences 11, no. 6: 2591. https://doi.org/10.3390/app11062591

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop