key: cord-0018214-wlodjzoh authors: Popovic, Marko; Minceva, Mirjana title: Standard Thermodynamic Properties, Biosynthesis Rates, and the Driving Force of Growth of Five Agricultural Plants date: 2021-05-31 journal: Front Plant Sci DOI: 10.3389/fpls.2021.671868 sha: ff4d6da020b5bc0e132d0485d011f17db9fced74 doc_id: 18214 cord_uid: wlodjzoh Elemental composition of Gossypium hirsutum L. (cotton), Oryza sativa L. (Asian rice), Phaseolus vulgaris L. (common bean), Saccharum spp. L. (sugarcane), and Zea mays L. (corn) was used to calculate their empirical formulas (unit carbon formulas) and growth stoichiometry. The empirical formulas were used to find standard enthalpy of formation, standard molar entropy, standard Gibbs energy of formation, and standard molar heat capacity. A comparison was made between thermodynamic properties of live matter of the analyzed plants and other unicellular and multicellular organisms. Moreover, the growth process was analyzed through standard enthalpy, entropy, and Gibbs energy of biosynthesis. The average standard Gibbs energy of biosynthesis was found to be +463.0 kJ/C-mol. Thus, photosynthesis provides energy and carbon for plant growth. The average intercepted photosynthetic energy was found to be 15.5 MJ/C-mol for the analyzed plants. However, due to inefficiency, a great fraction of the intercepted photosynthetic energy cannot be used by plants. The average usable photosynthetic energy was found to be –2.3 MJ/C-mol. The average thermodynamic driving force for growth is –1.9 MJ/C-mol. Driving forces of growth of C3 and C4 plants were compared. It was found that C4 plants have a greater driving force of growth than C3 plants, which reflects the greater efficiency of C4 photosynthesis. The relationship between the driving force and growth rates was analyzed by determining phenomenological L coefficients. The determined phenomenological coefficients span two orders of magnitude, depending on plant species and environmental conditions. The L coefficient of P. vulgaris was found to be lower than that of other plants, due to additional energy requirements of nitrogen fixation. Thermodynamics has a great potential in life sciences (Von Stockar and Liu, 1999; Von Stockar, 2010 , 2013a Lucia, 2015) . Thermodynamics has been used in systems biology to analyze metabolic pathways and analyze metabolic strategies evolved by microorganisms (Von Stockar and Liu, 1999; Von Stockar, 2013b) . Thermodynamics is used more and more often in biology for analysis of microorganism cells and communities McInerney and Beaty, 1988; Del Giorgio and Cole, 1998; Kleerebezem and Van Loosdrecht, 2010; Soh and Hatzimanikatis, 2010) , as well as plants (Dragicevic and Sredojevic, 2011; Keller, 2013) . Morowitz (1968) was the first to roughly estimate thermodynamic properties of ubiquitously present microorganisms using statistical thermodynamics. Afterward, the interest in biothermodynamics increased, which motivated researchers to study the subject in more detail. Wang et al. (1976) and Duboc et al. (1999) reported experimental enthalpies of combustion of microorganisms. The next great advance in the field was made by Battley et al. (1997) by experimentally determining the entropy of Saccharomyces cerevisiae. Based on this data, Battley (1998) has made a very detailed analysis of S. cerevisiae growth, determining the change in specific enthalpy, entropy, and Gibbs free energy for the process in both aerobic and anaerobic modes and on several substrates. Barros and Feijóo (2003) made a quantitative analysis of soil microorganism growth, describing it with a growth reaction. Multicellular and unicellular organisms represent a highly organized amount of substance, containing nucleic acid, proteins, lipids, and carbohydrates, characterized by specific elemental composition (empirical formula) and thermodynamic properties (enthalpy, entropy, and Gibbs energy) (Von Stockar, 2013a; Ozilgen and Sorguven, 2016) . The conversion of energy by plants must be performed in full accordance with the general laws of thermodynamics (Ksenzhek and Volkov, 1998) . Plant growth rate is proportional to the product of the metabolic rate and the metabolic efficiency for production of anabolic products . Thermodynamics has been used to study plant growth and photosynthesis. Hansen et al. (1994) related metabolic heat and CO 2 rates with the plant growth rate and carbon-conversion efficiency, by developing a model based on physiological parameters. Criddle et al. (2005) found that plants have evolved to have maximum growth rates at temperatures most common in their environment. Moreover, analysis of the energetics of respiration shows that measurements of metabolic heat and CO 2 rates by calorespirometry, combined with estimates of substrate and biomass composition, are sufficient to calculate substrate carbon-conversion efficiencies, anabolic rates or rates of growth and development, and relative activities of metabolic paths (Macfarlane et al., 2002; Hansen et al., 2005) . Dragicevic and Sredojevic (2011) used thermodynamic principles to analyze plant physiological processes, including seed germination and plant growth. Keller (2013) analyzed the energetics of photosynthesis. Enquist et al. (1998 Enquist et al. ( , 1999 found that energetic constraints on metabolism are reflected in population density, and other ecological and evolutionary phenomena, including resource allocation among species in ecosystems. Attorre et al. (2019) developed a model of plant growth based on nonequilibrium thermodynamics, which was able to predict CO 2 fixation by plants, as well as their response to changes in their environment. Colombi et al. (2019) used isothermal calorimetry to study energy used for root growth and penetration through the soil. Herrmann and Colombi (2019) used calorespirometry to study energy use efficiency of root growth. Ksenzhek and Volkov (1998) analyzed the applications of thermodynamic principles to plant growth and metabolism. Growth of organisms requires a thermodynamic driving force, in the form of negative Gibbs energy (Von Stockar, 2013b) . While the value of the driving force of higher plants remains unknown, it has been determined for heterotrophic microorganisms (Von Stockar and Liu, 1999; Von Stockar, 2013b) and microalgae. Von Stockar et al. (2011) used calorimetry to measure thermodynamic properties of photosynthetic growth of the microalgae Chlorella vulgaris, strain CCAP 211/11B. They found that the culture absorbed -5,000 kJ of light per C-mol of biomass formed (Von Stockar et al., 2011) . This energy is partly used to cover the cost of biosynthesis of new biomass, which was measured to be +458 kJ/C-mol (Von Stockar et al., 2011) . The remaining -4,500 kJ/C-mol represents the driving force for growth (Von Stockar et al., 2011) . This excess driving force is required to keep the reaction sufficiently far from equilibrium for it to occur at a desired rate, since the reaction rate is proportional to distance from equilibrium Westerhoff et al., 1982; Balmer, 2011; Von Stockar, 2013b,c; Demirel, 2014; Popovic and Minceva, 2020a,b) . Solar radiation is the ultimate source of energy for most life, with the exceptions of hydrothermal vents in the deep ocean. It is harvested by autotrophs and then distributed through food webs in the form of chemical energy. The role of photosynthesis in plant growth has been extensively analyzed in the literature (Pearcy et al., 1981; Allen et al., 1988; Kirschbaum, 2011; Holding and Streich, 2013; Weraduwage et al., 2016) . However, the absorbed light energy goes through intermediate steps before it can be used to drive metabolic processes. It is first converted into chemical energy in the form of photosynthate, which is then partly oxidized in catabolism to produce ATP for metabolic processes (e.g., maintaining homeostasis and growth). A detailed discussion of thermodynamics of plant catabolism and respiration is given in Hansen et al. (1994 Hansen et al. ( , 2004 Hansen et al. ( , 2021 , Gary et al. (1995) , and Ellingson et al. (2003) . The goal of this research is to make a thermodynamic analysis of growth of five widely cultivated plants: Gossypium hirsutum L. (cotton), Oryza sativa L. (Asian rice), Phaseolus vulgaris L. (common bean), Saccharum spp. L. (sugarcane), and Zea mays L. (corn). Elemental compositions of the analyzed plants will be used to calculate empirical formulas, thermodynamic properties of plant tissues, and their growth, as well as to formulate growth reactions. After that, the calculated thermodynamic properties will be compared with those of viruses, bacteria, fungi, archaea, and human tissues. This will show whether there are significant differences in the driving force of growth of phototrophic and chemotrophic organisms, as well as between prokaryotes and eukaryotes. Moreover, the driving force will be related to the growth rate using nonequilibrium thermodynamics. Like all other organisms, plants have characteristic empirical formulas and thermodynamic properties of live matter. Their determination, starting from elemental composition data, is discussed in the Properties of Live Matter section. The empirical formulas are then used to find stoichiometry of growth and thermodynamic properties of biosynthesis, as discussed in the Properties of Growth and Biosynthesis section. The determined properties have an uncertainty, as discussed in the Uncertainties section. The starting point for this research is elemental composition of plant dry matter reported in the literature. Elemental composition of live matter is measured as mass fractions, which are then converted into unit carbon formulas, also known as C-mol or empirical formulas (Battley, 2013) . The general unit carbon formula of the analyzed plants is C nC H nH O nO N nN P nP S nS Na nNa K nK Mg nMg Ca nCa Cl nCl , where n J is the atomic coefficient of element J, which can be calculated using the equation where w J and w C are the mass fractions of element J and carbon in the biomass, respectively, while M J and M C are the molar masses of element J and C, respectively (Duboc et al., 1999) . Empirical formulas are reported for organism dry matter (excluding water) (Duboc et al., 1999) . Molar masses of empirical formulas, M bio , can be found through the equation (Duboc et al., 1999 ) Thermodynamic properties of live matter were calculated as described in, Patel and Erickson (1981) ; Hurst and Harrison (1992) , Battley (1998 Battley ( , 1999 , and Popovic (2019) . Elemental composition of animate matter can be used to determine its enthalpy of formation through the Patel-Erickson equation and classical reaction thermochemistry. The Patel-Erickson equation relates heat released during combustion to number of electrons transferred to oxygen (Patel and Erickson, 1981; Battley, 1998) : where C H 0 is standard enthalpy of combustion of live matter and E is the number of electrons transferred to oxygen during combustion to CO 2 (g), H 2 O(l), N 2 (g), P 4 O 10 (s), and SO 3 (g) (for a discussion on other conventions concerning SO 3 please see Popovic, 2019) . During combustion, a C atom gives its four valence electrons to O, H gives one, N gives none since it is converted to N 2 , P gives five, and S gives six. Inorganic ions, like Na + and Mg 2+ , are not included, since they are already in their highest oxidation state and cannot transfer any electrons to oxygen (Battley, 1998) . Thus, E is calculated through the equation: where n C , n H , n O , n N , n P , and n S are, respectively, the number of C, H, O, N, P, and S atoms in the biomass empirical formula (Patel and Erickson, 1981; Battley, 1998) . If any of these atoms are not present in the empirical formula, they are just neglected during the calculation (Battley, 1998) . Standard enthalpy of combustion of live matter C H 0 (bio) is the standard enthalpy change of the reaction. C nC H nH O nO N nN P nP S nS K nK Mg nMg Ca nCa Al nAl Si nSi Mn nMn Fe nFe Cl nCl + (n C + 1 /4n H + 1 1 /4n P + 1 1 /2n S + 1 /4n K + 1 /2n Mg Thus, C H 0 can be used to find standard enthalpy of formation, f H 0 (bio), of live matter, through the Hess law (Battley, 1998; Popovic, 2019) . Elemental composition can also be used to determine standard molar entropy of live matter, S 0 m (bio), through the Battley equation (Battley, 1999) S o m (bio) = 0.187 where n J is the number of atoms of element J in the empirical formula of the biomass, S 0 m (J) is standard molar entropy of element J, and a J is the number of atoms per molecule of element J in its standard state elemental form. For example, the standard state elemental form of carbon is graphite, which is simply written as C, which makes a C = 1. On the other hand, hydrogen, oxygen, and nitrogen are in their standard state elemental forms of all diatomic gasses H 2 , O 2 , and N 2 , respectively, which implies that a H = a O = a N = 2. The summation is over all elements constituting the dry live matter. The Battley equation can also be used to determine standard entropy of formation of live matter f S 0 (bio). In this case, it takes the form (Battley, 1999) Finally, by combining standard enthalpy and entropy of formation, it is possible to calculate standard Gibbs energy of formation of live matter f G 0 (bio), as where T is the temperature. Standard molar heat capacities of the analyzed plants were determined using the Hurst-Harrison equation. The Hurst-Harrison equation is a group contribution model, describing standard molar heat capacity of a substance, C 0 p , m , as a sum of contributions from its constituent elements where n J is the number of atoms of element J in the substance formula, while c J is an empirical parameter that describes the contribution of element J to heat capacity (Hurst and Harrison, 1992; Ozilgen and Sorguven, 2016) . The values of c J can be found in Hurst and Harrison (1992) . Growth is a chemical process in which an organism converts nutrients into new live matter, a process that can be represented by chemical reactions, known as growth reactions (Raven, 1985; Battley, 1998 Battley, , 2013 Von Stockar, 2013b; Pineda-Insuasti et al., 2014; Popovic and Minceva, 2020a) . Growth reactions describe the stoichiometry with which nutrients are converted into live matter. The non-metals C, H, N, O, P, and S, also known as CHNOPS elements, are essential for all living systems, forming the majority of cell structures (Wackett et al., 2004) . Carbon enters photosynthetic tissues of plants as carbon dioxide. Hydrogen and oxygen enter plants mostly through the root as water. Nitrogen enters plants mainly through roots in the form of ammonium or nitrate salts. Phosphorus enters plants predominantly from soil in the form of phosphate salts. Sulfur is taken up by plants mostly from soil in the sulfate form (Linzon et al., 1979) . Except through soil, plants can take many elements through leaves from foliar fertilizers, which can reduce the total amount of fertilizer applied and achieve high fertilizer efficiency (Niu et al., 2021) . Except for the CHNOPS elements, plants use a great number of metallic ions, which are taken predominantly from the soil through the roots. Potassium greatly influences plant growth and is taken as K + (Ragel et al., 2019; Xu et al., 2020) . Magnesium is important in protein synthesis and is imported by plants in the form of Mg 2+ (Verbruggen and Hermans, 2013; Guo et al., 2016; Hauer-Jákli and Tränkner, 2019) . Calcium is taken as Ca 2+ (White et al., 2002; White and Broadley, 2003) and has a dual function as a structural component of cell walls (Hepler, 2005; Thor, 2019) and membranes and as intracellular second messenger (Knight et al., 1997; Blume et al., 2000; Michard et al., 2011; Monshausen et al., 2011; Wilkins et al., 2016; Ortiz-Ramírez et al., 2017; Tang and Luan, 2017; Kudla et al., 2018; Thor, 2019) . Silicon is taken in the form of silicic acid Si(OH) 4 (Ma and Yamaji, 2006) and has beneficial effects on growth, development, yield, and disease resistance of many plant species (Ma, 2004; Ma and Yamaji, 2006; Luyckx et al., 2017) . Aluminum is present in plants as Al 3+ ion (Bojórquez-Quintal et al., 2017) . In small amounts, aluminum stimulates root growth, increases nutrient uptake, and improves enzyme activity but inhibits root growth at higher concentrations (Bojórquez-Quintal et al., 2017) . Manganese is taken as Mn 2+ and is an important cofactor for catalyzing the water-splitting reaction in photosystem II (Alejandro et al., 2020) . Iron is present in the soil mostly as Fe 3+ ions but can also be present as Fe 2+ in smaller amounts. Iron is imported by plants using various strategies (Morrissey and Guerinot, 2009; Jacoby et al., 2017; Mhlongo et al., 2018) and is used in many processes, including electron transfer chains and photosynthesis (Schmidt et al., 2020) . Chlorine is taken as Cl − and is important in osmoregulation, stomata opening, and photosynthesis (Critchley, 1985; Coleman et al., 1987; Chen et al., 2010; Marschner, 2011; Colmenero-Flores et al., 2019) . Other elements, such as Cu and Zn, are also used by plants. However, the literature did not contain sufficient data to include these elements into the calculations. Since their amount is relatively small compared with the CHNOPS elements (Wackett et al., 2004) , they do not influence thermodynamic properties of plants. However, it would be interesting to consider their influence on plant nutrition in the future from the perspective of biothermodynamics and growth reactions. Based on the discussion above, the general growth reaction proposed for the analyzed plants is where "Bio" denotes live matter (excluding water). Stoichiometric coefficients of nutrients and metabolic products are specific for Frontiers in Plant Science | www.frontiersin.org each plant species and are discussed in the Live Matter Elemental Composition and Stoichiometry of Growth section. Growth reaction thermodynamic parameters were calculated using classical thermochemistry where ν's are stoichiometric coefficients of species participating in the reaction, while bs H 0 , bs S 0 , and bs G 0 are standard enthalpy, entropy, and Gibbs energy of biosynthesis, respectively (Atkins and de Paula, 2011; Atkins et al., 2017) . Thermodynamic properties of the studied plants were determined from elemental composition using empirical relations and thus have some uncertainty. C H 0 was found using the Patel-Erickson equation, the uncertainty of which is 5.36% (Popovic, 2019) . The determined C H 0 values were then subtracted from standard enthalpies of formation of oxides (Equation 6 ) to find f H 0 (bio). Since standard enthalpies of formation of oxides were precisely determined by experiment (more details in Chase, 1998) , they have a negligible error compared with that in C H 0 . Thus, the uncertainty in standard enthalpy of formation of live matter, δ( f H 0 (bio)), is equal to the error in C H 0 . δ f H 0 bio = 0.0536 · 111.14 kJ mol (4 n C + n H − 2 n O − 0 n N + 5 n P + 6 n S ) (15) The uncertainty in estimation of entropy using the Battley equation is 2% for dry matter and 19.7% for hydrated matter (Battley, 1999) . The greater of the two values was taken, and hence the uncertainty in standard molar entropy of live matter, δ(S 0 m (bio)), is f S 0 (bio) is defined as the standard entropy change of the reaction of formation of live matter from its constituent elements in their standard forms. + n P P(s, white) + n S S(s, rhombic) + n Na Na(s) + n K K(s) + n Mg Mg(s) + n Ca Ca(s) + 1 /2n Cl Cl 2 (g) → C nC H nH O nO N nN P nP S nS Na nNa K nK Mg nMg Ca nCa Cl nCl (17) and is defined as the difference in S 0 m (bio) and standard molar entropies of the elements, which have been determined with great accuracy by experiment (Chase, 1998) . Thus, the uncertainty in f S 0 (bio) is equal to that in S 0 m (bio) (Popovic, 2019) . f H 0 (bio) and f S 0 (bio) are used to find f G 0 (bio). Therefore, the uncertainty in the standard Gibbs energy of formation of live matter, δ( f G 0 (bio)), is (18) Finally, the uncertainty in f G 0 (bio) is equal to that in bs G 0 , since it is the greatest source of uncertainty in its determination. bs G 0 is determined using Equation (14), as the difference of f G 0 values of reactants and products of reaction (Equation 11 ). The f G 0 values of all reaction participants, except for live matter, have been determined with great accuracy by experiment (Chase, 1998) . Thus, uncertainty in Gibbs energy of biosynthesis, δ( bs G 0 ), is equal to δ( f G 0 (bio)). Similarly, δ( bs H 0 ) and δ( bs S 0 ) are equal to δ( f H 0 (bio)) and S 0 m (bio), respectively. Elemental composition and thermodynamic properties of plants and their growth are discussed in the following sections. Elemental compositions of the five plant species were taken from the literature (Latshaw and Miller, 1924; Medic et al., 2012; Titiloye et al., 2013; Pineda-Insuasti et al., 2014; Nam and Capareda, 2015; Fialho et al., 2019) . Information about the plant samples is summarized in Table 1 , including geographic origin, plant parts, and whether experimental enthalpy of combustion was reported. Calorimetrically determined enthalpies of combustion were reported for Gossypium hirsutum L., Oryza sativa L., Saccharum spp. L., and Zea mays L. stalk and cobs from Ghana (Titiloye et al., 2013; Nam and Capareda, 2015) . Since for these samples C H 0 (bio) was determined directly by experiment instead of Equation (4), their standard enthalpies of formation, f H 0 (bio), are more accurate. Elemental compositions of all the analyzed plants have been reported for dry matter. Based on the elemental composition of plants, unit carbon formulas were calculated, as described in the Properties of Live Matter section. The results are shown in Table 2 . It is interesting to compare the empirical formulas in Table 2 with average formulas of some of the major classes of organisms: bacteria (CH 1.7 O 0.4 N 0.2 ), fungi (CH 1.7 O 0.5 N 0.1 ), and algae (CH 1.7 O 0.5 N 0.1 ) (Popovic, 2019 ). The C:H ratio of the analyzed plants varies from 1.54 to 1.85, with an average of 1.63, which is close to that of the other classes of organisms. However, the analyzed plants have a C:O ratio from 0.71 to 0.90, with an average of 0.80, which is a significantly higher than the C:O ratio of the other classes of organisms. On the other hand, the C:N ratio of the analyzed plants varies from 7 × 10 −4 to 0.15, with an average of 0.02, which is lower than that of the other classes of organisms. The great span of C:N ratios (2014) and Fialho et al. (2019) and will be labeled "Waste 1" and "Waste 2," respectively, in further discussion. Waste 1 refers to "harvest wastes, the entire plant" (Pineda-Insuasti et al., 2014) . Waste 2 refers to "Straw bean: the residue of production of the bean" (Fialho et al., 2019) . Similarly, data for Zea mays L. cobs from Latshaw and Miller (1924) and Titiloye et al. (2013) will be labeled "Cobs 1" and "Cobs 2," respectively. The row "Zea mays L.-average" denotes average composition of the entire plant, based on organ weights reported by Latshaw and Miller (1924) . in plants originates from its dependence on growth conditions, including soil, meteorological factors, agricultural practices, and application of nitrogen fertilizers, as well as availability of other nutrients involved in N accumulation and cycling in crops. Thus, the average oxidation state of carbon in the analyzed plants is 0.00 (from -0.24 to 0.33), which is slightly higher than the averages for bacteria (-0.3), fungi (-0.4), and algae (-0.4). Moreover, the oxidation state of carbon in carbohydrates is also 0.00, which is in accordance with the fact that carbohydrates are a major constituent of plants. Based on the elemental composition, growth reactions were made and stoichiometric coefficients calculated, as described in the Properties of Growth and Biosynthesis section, which are summarized in Table 3 and Supplementary Material 3. Growth of plants is autotrophic, using CO 2 as a source of carbon, which is reduced using energy from light. Moreover, data in Supplementary Material 3 show that the greatest amount of O 2 is produced by growth of Z. mays grain (1.06 O 2 -mol/C-mol), while the lowest is that of Phaseolus vulgaris waste (0.92 O 2mol/C-mol), with the average being 1.00 O 2 -mol/C-mol. This is in good agreement with the average oxidation state of carbon of 0.00, since while going from +4 in CO 2 to 0.00 in biomass, 1 mol of carbon takes four electrons, which is exactly the amount released when 1 mol of O 2 is produced from H 2 O (half reaction: Based on elemental composition, standard thermodynamic properties of the analyzed plants were determined, as described in the Properties of Live Matter section. Table 4 gives standard thermodynamic properties of the analyzed plants and plant organs, including standard enthalpy of formation, standard Stoichiometric coefficient The table gives stoichiometric coefficients for growth reactions (Equation 11 ), of the form CO 2 (g) + H 2 O (l) + NH 4 + (aq) + H 2 PO 4 − (aq) + SO 4 2− (aq) + K + (aq) + Mg 2+ (aq) + Ca 2+ (aq) + Al 3+ (aq) + Si(OH) 4 (s) + Mn 2+ (aq) + Fe 3+ (aq) + Cl − (aq) → (Bio) + O 2 (g) + H + (aq). The product "(Bio)" denotes live matter (C nC H nH O nO N nN P nP S nS K nK Mg nMg Ca nCa Al nAl Si nSi Mn nMn Fe nFe Cl nCl ). The calculated stoichiometric coefficient values are negative for the reactants and positive for the products. molar entropy, standard Gibbs energy of formation, and standard molar heat capacity at constant pressure. All the standard enthalpies of formation are negative, implying that the biomass has a lower energy content than its constituent elements, due to partial oxidation of all other elements by oxygen and nitrogen (Erickson et al., 1978; Popovic, 2019) . All the standard molar entropies are positive, in accordance with the third law of thermodynamics (Atkins and de Paula, 2011; Atkins et al., 2017) . Standard molar entropies, S 0 m , per C-mol of live matter are around 37 J/C-mol K, laying between that of graphite, 5.740 J/mol K, and carbon in gaseous state, 158.10 J/mol K (Atkins et al., 2017) . This indicates that the mobility of C atoms in live matter is greater than in graphite but lower than in the gaseous state. All the standard Gibbs energies of formation are negative, with the average value being -126 kJ/C-mol. The standard molar heat capacities at constant pressure are around 35 J/C-mol K. Thermodynamic properties of the analyzed plants are compared with those of other organisms in Figure 1 . The data for other classes of organisms are summarized in Supplementary Material 2. Data in Figure 1 represent the average of values for 14 bacteria, 11 fungi, 11 algae, and 18 virus species, as well as 30 human tissues. Figure 1 shows that standard enthalpy and Gibbs energy of formation of the analyzed plants is among the most negative measured, but within the span of algae, which are also photosynthetic organisms. Plants have a more negative standard enthalpy and Gibbs energy of formation, indicating a greater energy content and lower oxidation state of carbon in the biomass, compared with other classes of organisms. The reason for this might be that plants gain energy directly from the sun, while heterotrophic organisms have to obtain energy from food. This is supported by the fact that algae, which are also photosynthetic organisms, also have highly negative enthalpy and Gibbs energy of formation. Standard molar entropies of all the organisms in Figure 1 are very similar. Furthermore, plants use a part of the stored energy in catabolic processes to maintain homeostasis, as well as for growth. Thermodynamic properties of biosynthesis were calculated for the analyzed plants and plant organs, as described in the Properties of Growth and Biosynthesis section. The results are given in Table 5 . The average standard enthalpy and Gibbs energy of growth are +454.4 and +463.0 kJ/C-mol, respectively, while the average standard entropy of growth is -29.1 kJ/Cmol. To compare, Gibbs energies of biosynthesis of several organisms have been determined for aerobic growth on a mixture of amino acids and glucose: Escherichia coli -45.25 kJ/C-mol, Pseudomonas (strain C12B) -18.67 kJ/C-mol, Bacillus subtilis -31.75 kJ/C-mol, Cyanobacteria synechocystis (strain PCC 6803) -13.74 kJ/C-mol, and Saccharomyces cerevisiae -15.90 kJ/C-mol (Popovic and Minceva, 2020a) . For S. cerevisiae, Gibbs energy of biosynthesis was determined for several environments: anaerobic on glucose -11.05 kJ/C-mol, aerobic on glucose -35.76 kJ/C-mol, aerobic on ethanol +28.86 kJ/C-mol, and aerobic on acetic acid +24.10 kJ/C-mol (Battley, 1998) . Chlorella vulgaris was found to have a Gibbs energy of biosynthesis of +458 kJ/C-mol, during phototrophic growth (Von Stockar et al., 2011) . Therefore, Gibbs energy of biosynthesis of heterotrophic organisms is slightly negative or sometimes slightly positive. The reason for this is that Plant part f H 0 (kJ/C-mol) S m 0 (J/C-mol K) f G 0 (kJ/C-mol) C p,m 0 (J/C-mol K) heterotrophic organisms use substrates of biological origin, with a Gibbs energy similar to their own live matter. On the other hand, the analyzed plants and C. vulgaris are phototrophic organisms and have highly positive Gibbs energies of biosynthesis of around +500 kJ/C-mol. This is due to the high energy requirement for reduction of carbon from CO 2 . However, phototrophic growth is still thermodynamically feasible, due to photosynthesis (Von Stockar, 2013b) . Standard Gibbs energy of formation is by definition the Gibbs energy change of reaction (17), the hypothetic formation of live matter from elements in their standard states (Atkins and de Paula, 2011) . The elements represent merely a reference state, relative to which Gibbs energy is measured (Atkins and de Paula, 2011) . This reference state was chosen because it is not possible to determine absolute Gibbs energy. On the other hand, Gibbs energy of biosynthesis is the Gibbs energy change of reaction (11), representing formation of live matter from nutrients in its environment (Von Stockar, 2013b) . Depending on the nutrients and environment, Gibbs energy of biosynthesis can vary even for a single organism (Battley, 1998) . Energetics of photosynthesis is summarized in Figure 2 . The relationship between the biomass accumulation and light energy intercepted by plants is known in agriculture as radiation use efficiency, RUE, an important quantifier of crop production, combining photosynthesis and the efficiency of a plant in producing live matter (Monteith and Moss, 1977; Hatfield and Dold, 2019) . RUE is defined as the slope of the line expressing the relationship between plant biomass, m bio , and intercepted solar radiation where f i is the fraction of intercepted photosynthetically active radiation at stage i of plant's development and R sp is the incoming photosynthetically active radiation (Monteith and Moss, 1977; Sadras et al., 2016; Hatfield and Dold, 2019) . Photosynthetically active radiation denotes light in the 400-700 nm range (Murchie et al., 2019) . The value of f i can be found through the equation (Sadras et al., 2016) . Crop coefficients are calculated, according to the Food and Agriculture Organization (FAO) method (Doorenbos and Pruitt, 1977) , by dividing the crop cycle into four stages: initial, rapid growth, maximum, and declining (Allen et al., 1998; Sadras et al., 2016) . For each stage, K C,i values have been tabulated for many plants (Allen et al., 1998) . Similarly, RUE has been determined for many plants and depends on environmental conditions, such as availability of nitrogen and other nutrients, temperature, soil water, atmospheric humidity, and wind (Sadras et al., 2016; Hatfield and Dold, 2019; Murchie et al., 2019) . More information about RUE can be found in Murchie et al. (2019) . RUE, defined as the biomass grown per unit light energy, can be used to calculate the intercepted photosynthetic energy, q, needed by plants to incorporate 1 mol of carbon into new biomass where M bio is the molar mass of live matter empirical formula [mass per 1 C-mol, calculated using Equation (2)]. Equation (20) is Equation (19) written for 1 C-mol of plant biomass, with q representing the sum term. The values of q for the analyzed (C) standard Gibbs energy of formation f G 0 . The columns represent the average thermodynamic parameters for the organism groups, while the error bars show the spread in thermodynamic properties within the groups. Data for bacteria, fungi, and algae taken from Popovic (2019) . Data for human tissues taken from Popovic and Minceva (2020c) . Data for viruses are taken from (Popovic and Minceva, 2020a,b) . plants were calculated using literature RUE values and are given in Table 6 . It would be interesting to compare the q-values from Table 6 with the result of an alternative calculation. Thus, we will estimate q in another way. Approximately 5.7 × 10 24 J of solar energy is irradiated to the Earth's surface on an annual basis (Miyamoto, 1997) . Of this energy, 98-99% is reflected or absorbed by objects, leaving (1.5 ± 0.5)% to be captured by photosynthesis (Thompson et al., 2018) . Thus, the total energy absorbed by photosynthetic organisms is (8.6 ± 2.5) × 10 22 J/year. Photosynthetic organisms utilize this solar energy to fix 2 × 10 11 tons of CO 2 per year (Miyamoto, 5 | Thermodynamic properties of biosynthesis of the analyzed plants: standard enthalpy ( bs H 0 ), entropy ( bs S 0 ), and Gibbs energy ( bs G 0 ) of biosynthesis. Plant part bs H 0 (kJ/C-mol) bs S 0 (J/C-mol K) bs G 0 (kJ/C-mol) Gossypium hirsutum L. Cob shell 456 ± 24 -24.1 ± 6.8 463 ± 25 Whole stover 457 ± 25 -23.2 ± 6.6 464 ± 25 Biosynthesis represents a chemical process in which CO 2 , H 2 O, and nutrients are converted into new live matter. Thermodynamic properties of biosynthesis do not include the contribution of photosynthetic energy. The reported thermodynamic properties are for biosynthesis reaction (Equation 11) with stoichiometric coefficients calculated using Tables 2, 3. 1997). Since 1 mol of CO 2 contains 1 mol of carbon and the molar mass of CO 2 is 44 g/mol, the total amount of carbon fixed by plants is found to be 4.5 × 10 15 C-mol/year. Therefore, by dividing the total energy (8.6 ± 2.5) × 10 22 J/year by the amount of carbon fixed 4.5 × 10 15 C-mol/year, the intercepted photosynthetic energy is found to be q = (18.8 ± 6.3) MJ/C-mol. This value is in good agreement with the q-values of the analyzed plants from Table 6 . The intercepted photosynthetic energy represents all light energy received by a plant and seems to be very large. However, photosynthesis has several sources of inefficiency: (1) 47% is lost since it is not in the 400-700 nm region used for photosynthesis; (2) 30% is lost in reflection, absorption, and transmission by leaves; (3) 24% for quantum efficiency requirements; (4) 68% in conversion to glucose; and (5) 35-40% on the processes of dark and photorespiration (Sudhakar and Mamat, 2019) . Inefficiency sources (1) and (2) are taken into account in Equation (19) by R sp and f i , respectively. The first source of inefficiency is taken into account when calculating RUE, since Equation (19) contains only R sp which represents only photosynthetically active 400-700 nm light. Moreover, f i in Equation (19) takes into account the correction for reflection, absorption, and transmission by leaves. Thus, these two sources of inefficiency are taken into account when calculating q using Equation (20), which is derived from Equation (19). However, the remaining three sources of inefficiency have not been taken into account. Thus, the photosynthetic efficiency, µ, in converting the intercepted photosynthetic energy q into energy usable to plants is µ = 100% − 24%quantum efficiency loss · 100% − 68%conversion to glucose loss · 100% − 40% photorespiration loss = 15% (21) Thus, the usable photosynthetic energy, ps G, is the photosynthetic energy actually available for plants to drive their photosynthetic metabolism and is lower than the intercepted photosynthetic energy, q The negative sign was added because plants receive energy from the sun as q and then spend it on metabolic processes as ps G. A more formal derivation of Equation (22) is given in Supplementary Material 1. The sum of the usable photosynthetic energy and the Gibbs energy required for biosynthesis, bs G 0 , represents the driving force of growth, DF G (Von Stockar et al., 2011; Von Stockar, 2013b) . The values of q, ps G, and DF G for the analyzed plants are given in Table 6 . The DF G values greatly vary, even within a species. For example, the driving force for growth of Saccharum spp. ranges between -1.0 and -3.8 MJ/C-mol. The reason for this is great variation in reported RUE, which is very different for plants grown with irrigation and in dry farming (rain-fed) (De Silva and De Costa, 2012) . Moreover, RUE of Z. mays depends on its growth location (Hatfield and Dold, 2019) . However, all the driving force values, despite their differences, have a general trend: they are all negative, making growth of plants a thermodynamically favorable process. The results for vascular plants obtained in this work can be compared with those for unicellular algae C. vulgaris. C. vulgaris was found to have an intercepted photosynthetic energy of q = 5,000 kJ/C-mol and a Gibbs energy of biosynthesis of bs G = +458 kJ/C-mol (Von Stockar et al., 2011) . The average values for the plants analyzed in this research are q = 15,537 kJ/Cmol and bs G = +469 kJ/C-mol. To compare the q-values, the photosynthetic efficiency, µ, has to be taken into account through Equation (22). Thus, usable photosynthetic energy, ps G, is -750 kJ/C-mol for C. vulgaris and -2,330 kJ/C-mol for the plants analyzed in this research. These values are then added to the corresponding bs G (Equation 23) to obtain the driving forces of growth, DF G, which are -292 kJ/C-mol for C. vulgaris and -1,861 kJ/C-mol for the plants analyzed in this research. The difference between these two values most likely originates from additional processes present in multicellular organisms, such as nutrient uptake by the root (Sondergaard et al., 2004) and great energy used by the root to penetrate through the soil during growth Herrmann and Colombi, 2019) . In nutrient limited conditions, plants will spend energy to acquire FIGURE 2 | Distribution of energy in photosynthesis. (A) Total energy of the sun is partly not in the photosynthetic 400-700 nm region and is partly reflected, absorbed, or transmitted by leaves. The remaining energy is used in photosynthesis and is called intercepted photosynthetic energy (q, shown in orange). (B) Intercepted photosynthetic energy is not the actual energy used by the plant. A part is lost due to quantum efficiency requirements, in conversion to glucose, and in dark and photorespiration. The remaining energy is usable photosynthetic energy ( ps G, shown in orange). (C) Usable photosynthetic energy is used partly to provide energy for biosynthesis ( bs G) and partly dissipated to make growth occur at a desired rate ( DF G). the limiting nutrient, in accordance with the law of the minimum. Moreover, photosynthesis does not directly drive growth but generates sugars that are oxidized in aerobic respiration to make ATP, which is hydrolyzed to release energy required for plant growth (Hansen et al., 1994 (Hansen et al., , 2004 (Hansen et al., , 2021 Gary et al., 1995; Ellingson et al., 2003) . Each step in this process dissipates energy gained during photosynthesis. The comparison can be carried further by including heterotrophic organisms. Heterotrophic organisms have a driving force of growth of approximately DF G = -500 kJ/Cmol (Heijnen and Van Dijken, 1992; Von Stockar and Liu, 1999; Liu et al., 2007; Von Stockar, 2013b) . The corresponding DF G values for photosynthetic growth are -292 kJ/C-mol for C. vulgaris and -1,861 kJ/C-mol for the plants analyzed in this research. The value for C. vulgaris is on the same order of magnitude as that for heterotrophic organisms. However, the plants analyzed in this research have a significantly greater driving force, which is most likely due to greater energy demands required by multicellular organisms, as discussed above. Data in Table 6 can be used to compare energetics of C3 and C4 plants. Saccharum spp. L. have on average the greatest usable photosynthetic energy ps G and driving force DF G. The reason is that Saccharum spp. L. is a C4 plant and is thus more efficient than C3 plants. Representatives of C3 plants, G. hirsutum 2.0 ± 0.4 13.5 ± 3.0 -2.0 ± 0.4 +0.515 ± 0.011 -1.5 ± 0.4 Phaseolus vulgaris L. 1.9 ± 0.4 16.0 ± 3.0 -2.4 ± 0.5 +0.428 ± 0.025 -2.0 ± 0.5 Saccharum spp. L. 1.4 ± 0.7 19.4 ± 9.5 -2.9 ± 1.4 +0.464 ± 0.005 -2.4 ± 1.4 Zea mays L. 2.3 ± 1.9 11.7 ± 9.7 -1.8 ± 1.5 +0.463 ± 0.026 -1.3 ± 1.5 The table contains radiation use efficiency (RUE), intercepted photosynthetic energy (q), usable photosynthetic energy ( ps G), Gibbs energy of biosynthesis ( bs G), and driving force of growth ( DF G). The bs G column gives average values for each species from Table 5 . RUE data were taken from Kiniry et al. (1989) ; Sinclair et al. (1992) , Kiniry (1994) ; Muchow et al. (1994) , Milroy and Bange (2003) ; Lindquist et al. (2005) , L., and Phaseolus vulgaris L. have lower photosynthetic energy ps G and driving force DF G. The lowest values are those of O. sativa L., which is also a C3 plant. Z. mays L. has the greatest span of ps G and DF G, due to many reported values. Many values have been reported because Z. mays is grown throughout the world, under a wide range of conditions. Under optimal conditions, Z. mays has a driving force of DF G = -(1.3 + 1.5) = -2.8 MJ/C-mol, which is the one of the highest values for the analyzed plant species. This is in good agreement with it being a C4 plant. The driving force of growth, in the form of negative Gibbs energy, exists to insure that growth occurs at a desired rate. If a process occurs near equilibrium, its Gibbs energy is very close to zero, making it thermodynamically very efficient (Atkins and de Paula, 2011; Atkins et al., 2017) . However, processes near equilibrium occur at an infinitely slow rate (Atkins and de Paula, 2011; Demirel, 2014; Atkins et al., 2017) , which would not be practical for growth of organisms. Thus, organisms have to make a compromise, between energetic efficiency and growth rate (Von Stockar and Liu, 1999; Von Stockar, 2013c) . In other words, they have to waste a certain amount of energy to make growth occur at a desired rate (Von Stockar and Liu, 1999; Von Stockar, 2013c) . This wasted energy is the driving force for growth (Von Stockar and Liu, 1999; Von Stockar, 2013c) . The driving force, DF G, is hence related to the growth rate of the plant, r, through the equation where L is a phenomenological coefficient and T is temperature Westerhoff et al., 1982; Von Stockar, 2013c; Demirel, 2014; Popovic and Minceva, 2020a,b) . The phenomenological coefficient is a proportionality constant that takes into account the kinetic factors (Demirel, 2014) and can be calculated from Equation (24) if the driving force and growth rate are known. Plant growth data were taken from the literature for G. hirsutum L. (Shi et al., 2013) , P. vulgaris L. (Shi et al., 2013) , and Z. mays L. (Shi et al., 2013; Koca and Erekul, 2016) . The growth data were analyzed using the three-phase linear model , which distinguishes three phases in growth: lag, exponential, and stationary Pérez-Rodríguez and Valero, 2013; Pla et al., 2015) . These three phases are analogous to the initial, rapid growth, maximum crop cycle stages, from the FAO method for calculating crop coefficients (Doorenbos and Pruitt, 1977) . The original threephase linear model was modified to describe plant growth where m is plant mass at time t, m 0 is plant mass before the intensive growth phase, t lag is the duration of the lag phase (time from planting to the intensive growth phase), r is the growth rate, and t max is the moment the steady state is reached and growth ceases. The model has four adjustable parameters: m 0 , m max , t lag , and t max . The three-phase linear model is simple, making it a good primary model for obtaining input data for studying more complicated questions, such as the environmental influence on the growth rate (Garthright and Buchanan, 1997) . In our case, the Growth rates in g/day were obtained as the slopes of fitted lines shown in Figure 3 . The reported growth rates are for the rapid growth stage. The growth rates were converted into C-mol/day, using the equation r(C-mol/day) = r(g/day)/M bio , where M bio is the average molar mass for the plant species (Table 2) . three-phase linear model was chosen since it provides a simple and impartial way of determining plant growth rates from growth data. Fitting was made using least squares regression. The fitting results are presented in Figure 3 . Growth rate r was calculated as the slope of the fitted line using Equation (25d). The growth rates obtained from fitted functions were combined with DF G ( Table 6 ) to find the phenomenological coefficients using Equation (24). The results are given in Table 7 . The obtained phenomenological coefficients vary greatly, depending on plant species and growth conditions. Phenomenological coefficients are specific for every plant species and environmental conditions, including soil, insolation, and water. Thus, the values of the phenomenological coefficients can vary even within one species, if it is cultivated under various conditions. For example, Z. mays cultivated in Turkey and China have different phenomenological coefficients, which are influenced by use of different seeds (hybrids) and growth on different soil, as well as temperature, water, and nutrients (Niu and Masabni, 2018) . Data in Table 7 can be interpreted in terms of metabolic processes performed by the analyzed plants. P. vulgaris L. has the lowest phenomenological L coefficient. The reason is that it is a legume. Legumes can fix nitrogen and thus have a high nitrogen content (P. vulgaris L. Waste 1 has the greatest nitrogen content in among all the samples from Table 2 ). However, nitrogen fixation requires a lot of energy, making less energy available to drive growth (Sinclair and Muchow, 1999) . Moreover, by comparing Z. mays L. grown in China and Turkey, it can be seen that various sorts of the same species grown under different conditions can have very different phenomenological L coefficients and growth rates. Empirical formulas with both macro and micro elements were calculated for Gossypium hirsutum L. (cotton), Oryza sativa L. (Asian rice), Phaseolus vulgaris L. (common bean), Saccharum spp. L. (sugarcane), and Zea mays L. (corn). Moreover, stoichiometry of growth was determined and expressed in the form of growth reactions. Plant live matter and growth process were analyzed by determining their standard thermodynamic properties of formation and biosynthesis, respectively. The average standard Gibbs energy of formation of the analyzed plants was found to be -126 kJ/C-mol. The analyzed plants were found to have among the lowest enthalpies and Gibbs energies of formation relative to the elements, along with algae, compared with bacteria, fungi, viruses, and human tissues. The highly negative enthalpies and Gibbs energies of formation of plants and algae indicate a high energy content, originating from their ability to obtain energy directly from the sun. Standard Gibbs energies of biosynthesis are positive for all the analyzed plants, since photosynthetic growth is based on reducing carbon in CO 2 . The average standard Gibbs energy of biosynthesis of the analyzed plants is +463.0 kJ/C-mol. Photosynthesis was analyzed through three parameters: intercepted photosynthetic energy, usable photosynthetic energy, and the driving force of growth. Energetics of photosynthesis was analyzed starting from RUE, biomass grown per unit light energy, reported in the literature. The average RUE for the analyzed plants was found to be 1.8 g/MJ. RUE was used to find the intercepted photosynthetic energy, light energy in the photosynthetic 400-700 nm range received by plants. The average intercepted photosynthetic energy was found to be 15.5 MJ/C-mol for the analyzed plants. However, intercepted photosynthetic energy is not completely usable to plants. Thus, it was corrected for photosynthetic efficiency of 15%, to find the usable photosynthetic energy, photosynthetic energy actually available for plants to drive their metabolism. The average usable photosynthetic energy was found to be -2.3 MJ/C-mol for the analyzed plants. The usable photosynthetic energy was added to the standard Gibbs energy of biosynthesis, to find the driving force of growth-energy spent on keeping growth sufficiently far from equilibrium for it to occur at a desired rate. The average driving force of growth of the analyzed plants was found to be -1.9 MJ/C-mol. This value is greater than that of heterotrophic and phototrophic unicellular organisms reported in the literature, most likely due to additional energy requirements by multicellular organisms. Driving forces of growth of C3 and C4 plants were compared. It was found that C4 plants have a greater driving force of growth than C3 plants, which reflects the greater efficiency of C4 photosynthesis. The driving force of growth is related to the growth rate, the proportionality constant being phenomenological L coefficients. Phenomenological L coefficients were determined for the first time for the analyzed plants. The phenomenological L coefficients of the analyzed plants span two orders of magnitude, depending on plant species and growth conditions. The L coefficient of P. vulgaris was found to be lower than that of other plants, due to additional energy requirements of nitrogen fixation. The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author/s. MP conceived and designed the research, collected the data, analyzed and interpreted the data, and wrote the manuscript. MM analyzed and interpreted the data, and wrote the manuscript. Both authors contributed to the article and approved the submitted version. The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2021. 671868/full#supplementary-material Manganese in plants: from acquisition to subcellular allocation ETc-single crop coefficient (Kc) Relationship between growth rate and net photosynthesis of Azolla in ambient and elevated CO2 concentrations Physical Chemistry for the Life Sciences Atkins' Physical Chemistry, 11th Edn A thermodynamic model for plant growth, validated with Pinus sylvestris data A combined mass and energy balance to provide bioindicators of soil microbiological quality The development of direct and indirect methods for the study of the thermodynamics of microbial growth An empirical method for estimating the entropy of formation and the absolute entropy of dried microbial biomass for use in studies on the thermodynamics of microbial growth A theoretical study of the thermodynamics of microbial growth using saccharomyces cerevisiae and a different free energy equation Heat capacity measurements from 10 to 300 K and derived thermodynamic functions of lyophilized cells of Saccharomyces cerevisiae including the absolute entropy and the entropy of formation at 298.15 K Receptor-mediated increase in cytoplasmic free calcium required for activation of pathogen defense in parsley Aluminum, a friend or foe of higher plants in acid soils When is simple good enough: a comparison of the Gompertz, Baranyi, and three-phase linear models for fitting bacterial growth curves NIST-JANEF Thermochemical Tables, 4th Edn Chlorine nutrition of higher plants: progress and perspectives The location of the chloride binding sites in the oxygen-evolving complex of spinach Photosystem II Chloride as a beneficial macronutrient in higher plants: new roles and regulation Cortical cell diameter is key to energy costs of root growth in wheat Thermodynamic law for adaptation of plants to environmental temperatures The role of chloride in Photosystem II Growth and radiation use efficiency of sugarcane under irrigated and rain-fed conditions in Sri Lanka Bacterial growth efficiency in natural aquatic systems Nonequilibrium Thermodynamics: Transport and Rate Processes in Physical, Chemical and Biological Systems Crop Water Requirements. FAO Irrigation and Drainage Paper 24 Thermodynamics of seed and plant growth Chapter 6 quantitative calorimetry and biochemical engineering Determination of the enthalpy change for anabolism by four methods Allometric scaling of plant energetics and population density Plant energetics and population density Application of mass and energy balance regularities in fermentation Bio-coal production with agroforestry biomasses in Brazil The three-phase linear model of bacterial growth: a response Heat of combustion, degree of reduction and carbon content: 3 interrelated methods of estimating the construction cost of plant tissues Evaluation of radiation use efficiency of common bean (Phaseolus vulgaris L .) cultivars as affected by plant density under Mashhad climatic conditions Magnesium deficiency in plants: an urgent problem Calorespirometry in plant biology The relation between plant growth and respiration: a thermodynamic model Use of calorespirometric ratios, heat per CO 2 and heat per O 2, to quantify metabolic paths and energetics of growing cells Transformation of matter in living organisms during growth and evolution Radiation use efficiency: evaluation of cropping and management systems Chapter 1photosynthesis in the solar corridor system Critical leaf magnesium thresholds and the impact of magnesium on plant growth and photo-oxidative defense: a systematic review and meta-analysis from 70 years of research In search of a thermodynamic description of biomass yields for the chemotrophic growth of microorganisms Energetics of microbial growth: an analysis of the relationship between growth and its mechanistic basis by mosaic nonequilibrium thermodynamics Calcium: a central regulator of plant growth and development Energy use efficiency of root growtha theoretical bioenergetics framework Plant Growth Processes: Transpiration, Photosynthesis, and Respiration Leaf photosynthetic performance related to higher radiation use efficiency and grain yield in hybrid rice Estimation of liquid and solid heat capacities using a modified kopp's rule The role of soil microorganisms in plant mineral nutrition-current knowledge and future directions Effects of planting patterns and sowing densities on grain-filling, radiation use efficiency and yield of maize (Zea mays L.) in semi-arid regions Chapter 20: Thermodynamic analysis of photosynthesis Radiation-use efficiency and grain yield of maize competing with johnsongrass Radiation-use efficiency in biomass accumulation prior to grainfilling for five grain-crop species Does enhanced photosynthesis enhance growth? Lessons learned from CO2 enrichment studies A generalized method for thermodynamic state analysis of environmental systems Calcium signalling in Arabidopsis thaliana responding to drought and salinity Changes of dry matter, biomass and relative growth rate with different phenological stages of corn Plant Energetics Advances and current challenges in calcium signaling Elemental composition of the corn plant Maize radiation use efficiency under optimal growth conditions Sulfur concentrations in plant foliage and related effects A comparison of various Gibbs energy dissipation correlations for predicting microbial growth yields Bioengineering thermodynamics of biological cells Silicon and plants: current knowledge and technological perspectives Role of silicon in enhancing the resistance of plants to biotic and abiotic stresses Silicon uptake and accumulation in higher plants Application of an enthalpy balance model of the relation between growth and respiration to temperature acclimation of Eucalyptus globulus seedlings Marschner's Mineral Nutrition of Higher Plants Anaerobic community structure from a nonequilibrium thermodynamic perspective The effects of particle size, different corn stover components, and gas residence time on torrefaction of corn stover The chemistry of plant-microbe interactions in the rhizosphere and the potential for metabolomics to reveal signaling related to defense priming and induced systemic resistance Glutamate receptor-like genes form Ca2+ channels in pollen ubes and are regulated by pistil D-serine Nitrogen and light responses of cotton photosynthesis and implications for crop growth Renewable Biological Systems for Alternative Sustainable Energy Production (Issue 128). Food & Agriculture Org Dynamics of auxin-dependent Ca 2+ and pH signaling in root growth revealed by integrating high-resolution imaging with automated computer vision-based analysis Climate efficiency and crop production in britain Efficiency on the Use of radiation and corn yield under three densities of sowing Energy Flow in Biology: Biological Organization as a Problem in Thermal Physics Iron uptake and transport in plants: The good, the bad, and the ionome Radiation interception and biomass accumulation in a sugarcane crop grown under irrigated tropical conditions Crop radiation capture and use efficiency Experimental investigation of torrefaction of two agricultural wastes of different composition using RSM (response surface methodology) Plant production in controlled environments Effects of foliar fertilization: a review of current status and future perspectives Water and radiation use efficiency of sugarcane for bioethanol production in South Africa, benchmarked against other selected crops. South Afr Glutamate receptor-like channels are essential for chemotaxis and reproduction in mosses Biothermodynamics: Principles and Applications Estimation of heats of combustion of biomass from elemental analysis using available electron concepts Relationships between growth, photosynthesis and competitive interactions for a C3 and C4 plant Predictive Microbiology in Foods Stoichiometry equation to describe the growth of the Pleurotus ostreatus ceba-gliie-po-010606 strain Comparison of primary models to predict microbial growth by the plate count and absorbance methods Thermodynamic properties of microorganisms: determination and analysis of enthalpy, entropy, and Gibbs free energy of biomass, cells and colonies of 32 microorganism species A thermodynamic insight into viral infections: do viruses in a lytic cycle hijack cell metabolism due to their low Gibbs energy? Heliyon Thermodynamic insight into viral infections 2: empirical formulas, molecular compositions and thermodynamic properties of SARS, MERS and SARS-CoV-2 (COVID-19) viruses. Heliyon 6:e04943 Thermodynamic properties of human tissues Regulation of K + nutrition in plants Ph regulation in plants Radiation interception, radiation use efficiency and crop productivity Editorial: iron nutrition and interactions in plants The "general" ontogenetic growth model is inapplicable to crop growth Radiation use efficiency Variation in crop radiationuse efficiency with increased diffuse radiation Network thermodynamics in the postgenomic era Energization of transport processes in plants. roles of the plasma membrane H+-ATPase Artificial leaves: towards bio-inspired solar energy converters Regulation of calcium and magnesium homeostasis in plants: from transporters to signaling network Efficiency of solar energy utilization Calciumnutrient and messenger Thermochemical characterisation of agricultural wastes from West Africa. Industrial Crops Prod Seed-oriented planting improves light interception, radiation use efficiency and grain yield of maize Physiological and molecular responses to magnesium nutritional imbalance in plants Biothermodynamics of live cells: a tool for biotechnology and biochemical engineering Biothermodynamics: The Role of Thermodynamics in Biochemical Engineering Biothermodynamics of live cells: energy disipation and heat generation in cellular cultures Live cells as open non-equilibrium systems Does microbial life always feed on negative entropy? Thermodynamic analysis of microbial growth Calorimetry and thermodynamic aspects of heterotrophic, mixotrophic, and phototrophic growth Microbial genomics and the periodic table Thermodynamic evaluation of microbial growth Pectin methylesterification impacts the relationship between photosynthesis and plant growth Thermodynamics of growth non-equilibrium thermodynamics of bacterial growth the phenomenological and the Mosaic approach Genes for calcium-permeable channels in the plasma membrane of plant root cells Calcium in plants Calciummediated abiotic stress signaling in roots Effects of potassium levels on plant growth, accumulation and distribution of carbon, and nitrate metabolism in apple dwarf rootstock seedlings Radiation interception and use efficiency contributes to higher yields of newer maize hybrids in Northeast China The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.Copyright © 2021 Popovic and Minceva. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.