Next Article in Journal
Changing Dynamics of SARS-CoV-2: A Global Challenge
Next Article in Special Issue
Hyperspectral Estimation of Nitrogen Content in Different Leaf Positions of Wheat Using Machine Learning Models
Previous Article in Journal
Empirical Analysis of Parallel Corpora and In-Depth Analysis Using LIWC
Previous Article in Special Issue
Mathematical Description of Changes of Dried Apple Characteristics during Their Rehydration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mathematical Modeling to Estimate Photosynthesis: A State of the Art

by
Luz del Carmen García-Rodríguez
1,
Juan Prado-Olivarez
1,
Rosario Guzmán-Cruz
2,
Martín Antonio Rodríguez-Licea
3,
Alejandro Israel Barranco-Gutiérrez
1,
Francisco Javier Perez-Pinal
1 and
Alejandro Espinosa-Calderon
4,*
1
Department of Electrical and Electronic Engineering, Tecnológico Nacional de México, Celaya 38010, Mexico
2
Engineering School, Universidad Autónoma de Querétaro, Querétaro 76010, Mexico
3
CONACYT-Tecnológico Nacional de México, Celaya 38010, Mexico
4
Regional Center for Optimization and Device Development, Tecnológico Nacional de México, Celaya 38020, Mexico
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(11), 5537; https://doi.org/10.3390/app12115537
Submission received: 7 April 2022 / Revised: 18 May 2022 / Accepted: 23 May 2022 / Published: 30 May 2022
(This article belongs to the Special Issue Applications of Computer Science in Agricultural Engineering)

Abstract

:
Photosynthesis is a process that indicates the productivity of crops. The estimation of this variable can be achieved through methods based on mathematical models. Mathematical models are usually classified as empirical, mechanistic, and hybrid. To mathematically model photosynthesis, it is essential to know: the input/output variables and their units; the modeling to be used based on its classification (empirical, mechanistic, or hybrid); existing measurement methods and their invasiveness; the validation shapes and the plant species required for experimentation. Until now, a collection of such information in a single reference has not been found in the literature, so the objective of this manuscript is to analyze the most relevant mathematical models for the photosynthesis estimation and discuss their formulation, complexity, validation, number of samples, units of the input/output variables, and invasiveness in the estimation method. According to the state of the art reviewed here, 67% of the photosynthesis measurement models are mechanistic, 13% are empirical and 20% hybrid. These models estimate gross photosynthesis, net photosynthesis, photosynthesis rate, biomass, or carbon assimilation. Therefore, this review provides an update on the state of research and mathematical modeling of photosynthesis.

1. Introduction

1.1. Photosynthesis Process

Plants perform physiological functions that allow them to grow, develop, and reproduce. For their development and growth, plants require energy from sunlight, assimilated carbon dioxide (CO2), hydrogen, mineral nutrients, oxygen (O2), and a suitable temperature both in air and roots [1,2]. A fundamental physiological process for the vegetable kingdom is photosynthesis. The photosynthesis process occurs in terrestrial and aquatic plants, algae, and some types of bacteria that are essential for living species [3]. Photosynthesis is a process by which plants convert light energy into chemical energy to obtain sugar as a final product. On the other hand, another photosynthetic reaction is oxygenic, which is released into the atmosphere as a waste, being useful for the biological process of respiration [4].
Chloroplasts (plant organelles located in the leaf) are active metabolic centers that capture solar energy through chlorophyll (green pigment) and manufacture carbohydrates (glucose molecules) through the process of photosynthesis [3]. The energy from light absorbed by chlorophyll molecules in a leaf can take one of three ways: it can be used to drive photosynthesis; excess energy can be dissipated as heat, or it can be re-emitted as light-chlorophyll fluorescence [5]. Glucose is one of the main molecules that serve as an energy source for plants and animals. It is found in plant sap and in the human bloodstream, where it is known as “blood sugar” [6,7]. Then, through photosynthesis, plants produce glucose for their food and dispose of valuable oxygen for the respiration of other living beings. Plants can capture carbon dioxide and release oxygen during the day, but they undergo another change at night: they absorb oxygen and release carbon dioxide [8].
Photosynthesis is divided into two phases: In the first phase, absorption and conversion of energy happen; while in the second, the intake and assimilation of carbon occur. Light energy is absorbed by photosensitive biomolecules and transformed into a stable biochemical energy form. The constituent elements are taken from inorganic mineral sources (water, H2O; carbon dioxide, CO2; nitrates, NO 3 ; sulfates, SO 4 2 , among others.) and are incorporated into metabolizable organic biomolecules. Both phases are perfectly coordinated and interrelated. Traditionally, these phases have been called the light phase and the light-independent phase. The first phase is produced only by utilizing sunlight, which will provide us with oxygen, ATP (adenosine triphosphate), and NADPH (nicotinamide adenine dinucleotide phosphate). In the second phase (it can occur during the day and also at night), the Calvin cycle is carried out where glucose is produced by using the ATP, NADPH, and CO2 generated in the light phase [2,8].
The essential and predominant element in organic material is carbon. In photosynthesis, carbon is taken from the air’s carbon dioxide (CO2). Generally, in terrestrial plants, CO2 is incorporated from the atmosphere through stomata. Algae and aquatic plants take it from the CO2 dissolved in the surrounding water. At night, when photosynthesis is not active and, therefore, there is no demand for CO2 inside the leaf, the stomatal openings are reduced preventing unnecessary water loss. In the morning, when the water supply is abundant and solar radiation favors photosynthetic activity, the demand for CO2 inside the leaf is big and the stomatal pores are wide open, reducing the stomatal resistance to the diffusion of the CO2 [8].
Plants have developed three photosynthetic systems, C3, C4, and CAM (Crassulacean acid metabolism), with different anatomical and chemical characteristics [9], the C3 type being the most common [10,11]. The main difference between the types of photosynthesis lies in the way CO2 is synthesized. The first step in the Calvin cycle is carbon dioxide fixation by rubisco. Plants that use only this “standard” carbon fixation mechanism are called C3 plants because of the three-carbon compound (3-PGA) that they produce during this part of the photosynthetic process. In C4 plants, the light-dependent reactions and the Calvin cycle are physically separated: the light-dependent reactions occur in the cells of the mesophyll (spongy tissue in the center of the leaf) and the Calvin cycle occurs in special cells around the veins of the leaf. These cells are called vascular bundle cells. Some plants adapted to dry environments use the crassulaceae acid metabolism (CAM) pathway to minimize photorespiration. Instead of separating the light-dependent reactions and the use of CO2 in the Calvin cycle, the temporary separation of photosynthetic processes in CAM plants causes carbon assimilation to take place at night. At night, they open their stomata so that the CO2 diffuses into the leaves. This CO2 is fixed to oxaloacetate by PEP carboxylase (the same step that C4 plants use), which is then converted into malate or another organic acid [8]. The organic acid is stored inside vacuoles until the next day. During the day, CAM plants do not open their stomata, but they can still carry out photosynthesis. Because organic acids are transported out of vacuoles and broken down to release CO2, which enters the Calvin cycle. This controlled cycle maintains a high concentration of CO2 around rubisco [4]. The ability to represent the three photosynthetic types (C3, C4, and CAM) has important implications for studying natural ecosystems and agroecosystems [12].
The described complex process of photosynthesis must work, integrally and efficiently, in an environment where there is an enormous natural variability of factors that affect the rate of photosynthesis. For instance, light, environment temperature, air humidity, availability of water, and mineral nutrients in the soil, among others. Carbon dioxide (CO2) can also be considered as part of this list of relevant factors due to global climate change [13]. The photosynthesis rate of a leaf is conditioned by one of more than 50 individual reactions, each presenting its response to each environmental variable. This photosynthetic rate can widely vary between days and also throughout seasons, due to environmental factors such as light and temperature. It can also vary in the longer term during the coming decades as a response to increasing atmospheric CO2 levels. The increase in CO2 and other greenhouse gases in the atmosphere can cause global climate change. As can be understood, each of the aforementioned environmental factors affects the photosynthesis rate in a different way, depending on the time scale [8].
Photosynthesis is important for many reasons. From humanity’s point of view, it produces food and oxygen; therefore, it is often studied in its end products. However, these are secondary aspects of an integral process. The most important aspects are capturing and transforming light [14]. A critical component of crop production is the ability to produce more biomass [15].
Photosynthesis at the ecosystem scale, also known as the gross primary productivity (GPP), is the first step of CO2 entering the biosphere from the atmosphere. Over the past century, with the increasing carbon release from landcover change and fossil fuel burning, the CO2 accumulation rate on land, in the ocean and in the atmosphere has continuously increased. The increment of CO2 in Earth’s atmosphere is the major cause of global climate change. A major contribution to this high variability comes from GPP, as the photosynthesis process is vulnerable to droughts, heatwaves, floods, frost, and other types of disturbances. An accurate estimation of GPP will not only provide information about the ecosystem response to these extreme events but also help to predict the future carbon cycle dynamics [16].
For the agronomic sector the research about photosynthesis is useful in many aspects, such as: determining the state of crops or of certain plants; indicating crop production; helping in crop-health control; helping with the optimization of natural resources; detecting genetic alterations in plants to know how plants react to stressful situations; and serving as a predictive indicator of biomass accumulation in green plants [17,18,19,20,21,22].

Methods for Inferring Photosynthesis

Many of the methods for inferring photosynthesis are invasive; they physically or chemically interfere with the plant, altering its natural process during measurement. Non-invasive methods do not alter the plant’s natural process since there is no contact with the specimen [23,24].
Millan et al. [25] presented a review of the advantages and disadvantages of the methods used for photosynthesis estimation. Table 1 shows the classification of these methods. CO2 exchange is the most used method for constructing commercial and experimental equipment [26,27]. It is possible to measure photosynthesis at the level of individual leaves, whole plants, plant canopies, and even forests [28,29].
As can be noticed from previous Table 1 Photosynthesis is a complex process that involves several and distinct variables; hence, it cannot be directly measured, if not through a mathematical model. Then, photosynthesis can be inferred by integrating and measuring these variables of interest. Hence, it is possible to generate a mathematical model if the contributions of the variables of interest of a physical, chemical, or biological process, among others, are known [36,37]. It should be noted that all invasive or non-invasive methods that are used to infer photosynthesis apply some type of mathematical model. Therefore, the invasiveness or non-invasiveness when developing any method to estimate photosynthesis depends on the technique used to measure the variables of interest; for this reason, the present document focuses on studying and comparing the most relevant mathematical models reported in the literature. Authors like Zufferey [38], Polerecky [39], Millan [25], Espinosa [40], Magney [41], Aziz & Dianursanti [42], and Kitić [43], demonstrate the use of some photosynthesis measurement methods.

1.2. Mathematical Modeling

Any branch of science, as it progresses from qualitative to quantitative, is likely to reach the point where the use of mathematics to connect experiment and theory is essential.
Mathematical modeling consists of the following steps [44]:
  • Definitions;
  • Systems analysis;
  • Modeling;
  • Simulation;
  • Validation.
Mathematical models can be classified into mechanistic (white box), empirical (black box), and hybrid (gray box). These, in turn, have sub-classifications, as shown in Figure 1 [36].
Empirical models, also called black box models, mainly described a system’s responses by using mathematical or statistical equations without any scientific content, restrictions, or scientific principle. Depending on particular goals, this may be the best type of model to build [44]. Its construction is based only on experimental data and does not explain dynamic mechanisms; this refers to the fact that the system’s process is unknown [40]. Estimating an unknown function from the observations of its values is a problem. The basic advice in this aspect is to estimate models of different complexity and evaluate them using validation data. A good way to restrict certain classes of models’ flexibility is to use a regularized fit criterion. A key issue is finding a sufficiently flexible parameterization model. Another key is to find a suitable “close approach “to the model structure [45]. Researchers usually employ methods for predicting physiological parameters by using intelligent algorithms, such as Support Vector Machines (SVM), Back-Propagation Neural Network (BPNN), Artificial Neural Network (ANN), Deep Neural Network (DNN), and the combination of Wide and Deep Neural Network (WDNN) [46].
Mechanistic models, also called white box models, provide a degree of understanding or explanation of the modeled phenomena. The term “understanding” implies a causal relationship between quantities and mechanisms (processes). A well-built mechanistic model is transparent and open to modifications and extensions, more or less without limits. A mechanistic model is based on our ideas about how the system works, the important elements, and how they are related [44]. These models allow knowing the input or output variables and the variables involved during the modeling process [40,47]. Mechanistic models are more research-oriented than application-oriented, although this is changing as our mechanistic models become more reliable. Evaluation of such models is essential, although it is often, and inevitably, rather subjective. Conventional mechanistic models are complex, and unfriendly [44].
Figure 1 shows that both mechanistic and empirical models can be deterministic or stochastic. Determinists make definite quantitative predictions (plant dry-matter or animal intake) without any associated probability distribution. This can be acceptable in many cases; however, it may not be satisfactory for quite changeful quantities or processes (e.g., rain or the migration of diseases, pests, or predators). On the other hand, stochastic models include a random element as a part of the model so that the predictions have a distribution. One problem with stochastic models is that they can be technically difficult to build and complex to test or falsify [44].
In turn, the deterministic and stochastic models can be continuous or discrete. A mathematical model that describes the relationship between continuous signals in time is called time-continuous. Differential equations are frequently used to describe such relationships. A model that directly relates the values of the signals at the sampling times is called a discrete or sampled time model. Such a model is typically described by differential equations [48].
The continuous models are classified as dynamic since they predict how quantities vary with time, so a dynamic model is generally presented as a set of ordinary differential equations with time (t), the independent variable. On the other hand, the continuous models can also be static; they do not contain time as a variable and do not make time-dependent predictions [44].
Finally, dynamic models can be grouped or distributed. Partial differential equations mathematically describe many physical phenomena. The events in the system are, so to speak, scattered over the spatial variables. This description is called the distributed parameter model. If a finite number of changing variables describes the events, we speak of grouped models. These models are usually expressed by ordinary differential equations [48].
An intermediate model is classified as the semi-empirical or semi-mechanistic model between the black box and white box models. These models are also called gray box or hybrid models; they consist of a combination of empirical and mechanistic models [40].
The practical use of a mathematical model classification lies in understanding “where you are” in the mathematical model space and what types of models might apply to the problem. To understand the nature of mathematical models, they can be defined by the chronological order in which the model’s constituents usually appear. Usually, a system is given first, then there is a question regarding that system, and only then is a mathematical model developed. This process is denoted as SQM, where S is a system, Q is a question relative to S, and M is a set of mathematical states M = (σ1, σ2, …, Σn) which can be used to answer Q. Based on this definition, it is natural to classify mathematical models in an SQM space [36]. Figure 2 shows an approach to visualize this SQM space of mathematical models based on the white box and black box models classification. At the black box, at the beginning of the spectrum, models can perform reliable predictions based on data. At the white box end of the spectrum, mathematical models can be applied to the design, testing, and optimization of computer processes before they are physically carried out. On each of the S, Q, and M axes in Figure 2b, the mathematical models are classified based on a series of criteria compiled from various classification attempts in the literature [36].
There are different mathematical models related to biochemical, physical, and agroecological variables that estimate photosynthesis at the leaf, plant, or group of plant levels. Therefore, the study of mathematical modeling focused on the photosynthetic process becomes important in the agricultural sector. Since it is a direct indicator of a plant’s health. It also makes it possible to assess the consequences of global climate change on crop growth, since the high concentration of CO2, the increase in temperature and altered rainfall patterns can have serious effects on crop production in the near future [44].
However, to the best of the authors’ knowledge, a study on the diversity of mathematical modeling in the field of scientific research has not been approached nor focused on: the mathematical formulation, the complexity of the model, the validation, the type of crop (at the leaf, plant or canopy level), the analysis of the diversity of variables used with their respective units, as well as the invasiveness in their measurements. Hence, this manuscript presents a selective review of mathematical modeling to estimate photosynthesis.
In the literature, there is a review of the mathematical modeling of photosynthesis developed by Susanne Von Caemmerer. However, here only several models derived from the C3 model by Farquhar, von Caemmerer, and Berry are discussed and compared. The models described and reviewed here describe the assimilation rates of CO2 in a steady state and provide a set of hypotheses collected in a quantitative way that can be used as research tools to interpret experiments both in the field and in the laboratory. Additionally, it also provides tools for reflective experiments [49]. Conversely, the present paper provides a new vision of the state of the art in mathematical models with certain specifications. This information can be used to develop new mathematical models to estimate photosynthesis with new variables related to the plant’s habitat and with greater relevance to be implemented in electronic systems during the development of photosynthesis estimation equipment.
The objective of this manuscript is to present mathematical models with different characteristics to estimate photosynthesis; discuss its formulation, complexity, validation, applications, and invasiveness in the estimation method; show that the units corresponding to photosynthesis depend on the measurable biochemical, physical, and agroecological variables used for the estimation; analyze the relevance of the behavior trend in the photosynthetic process and prioritize over the specific magnitude itself. Therefore, in the authors’ opinion, this study provides a great opportunity to contribute to the mathematical-modeling knowledge to estimate photosynthesis.
The content of this work is organized in the following sections. Section 2 describes the criteria for the selection of mathematical models according to their classification and characteristics. In Section 3, a critical review of the white box, black box, and gray box models is given. Finally, conclusions are reported in the last section of this manuscript.

2. Techniques

Over the years, various mathematical models have been developed and introduced for photosynthesis estimation. The review of these models is based on an exhaustive search and analysis of the reported literature, it focuses on their mathematical formulation, validation, input variables, as well as the type of cultivation at the leaf, plant, or canopy level. As a result, an updated list of the most relevant mathematical evaluation models was obtained, and possible areas of opportunity was detected. The research was conducted in the main databases (Science Direct, Web of Science, IEEE, Google Scholar, Scielo, ResearchGate), scientific organizations such as the American Society of Agricultural and Biological Engineers ASABE, and specialized journals such as Bioresource Technology Reports, Ecological Modeling, Metabolic Engineering, Biotechnology and Bioengineering, Environmental and Experimental Botany, Bioresource Technology, Journal of American Society Hortaliza Science, Aquatic Botany, European Journal of Agronomy, Agricultural Systems, Wageningen Journal of Life, Journal of Plankton Research, Limnology and Oceanography, Biological Theory, Agricultural and Forest Meteorology, Computers and electronics in agriculture, Journal of Theoretical Biology, Annals of Botany, Photosynthesis Research, Experimental Marine Biology and Ecology, Marine Biology, Progress in Oceanography, and Plant Cell and Environment, among others. The research approach adopted for this study was comparing mathematical models to estimate photosynthesis with a variety of input variables developed over time. This approach allows us to contribute to the knowledge of the different existing methodologies for estimating photosynthesis.
The selection criteria were:
(a)
The models had to estimate gross photosynthesis, net photosynthesis, photosynthetic rate, biomass, or carbon assimilation. Wohlfahrt & Gu [50] reviewed the photosynthesis background and the associated terminology; they showed that different definitions and names are used for photosynthesis in literature. Then, as the understanding of photosynthesis has deepened, the terminologies and definitions of photosynthesis have also evolved [51]. In this work, Gross photosynthesis is defined as the total energy that plants fix during the photosynthesis process, while net photosynthesis is the total fixation of CO2 [52]. The photosynthetic rate refers to the amount of CO2 absorbed by a plant per unit of leaf area and unit of time. Biomass results from the net accumulation of assimilated CO2 throughout the growth cycle [53]. Lastly, carbon assimilation is the amount of CO2 stored in the plant. These terms are related to photosynthesis since plants can capture CO2 from the atmosphere and metabolize it through the photosynthetic process to obtain sugars and other compounds that are required for the normal development of their life cycle [26,54]. Then, the series of definitions described above refers to the estimation of photosynthesis.
(b)
The articles had to have a validation method to check if the different models used in the existing literature yielded similar results among other accredited models or with a photosynthesis measurement device on the market. This is important because one of the steps for modeling development is validation [55,56]. Validation is defined as the comparison of the model predictions with the observed values of the actual process to determine if the model fits its purpose; the validation results are a necessary step in the acceptance of the model [57,58,59]. Validation can also be defined as a demonstration that a model has acceptable predictive accuracy in that domain within a specific application. The quantitative precision of the predictions is often referred to as the “validity” of the model. Generally, it refers to the ability of a model to accurately predict the results of a particular experimental or observational scenario [44].
In this review, a series of papers related to the topic of mathematical modeling were analyzed for inclusion. During data extraction, some studies were excluded because they did not present the information required in the criteria established for selection. These excluded models have the following issues:
-
They did not estimate gross photosynthesis, net photosynthesis, photosynthetic rate, biomass, or carbon assimilation; instead, they modeled chlorophyll, fluorescence, or stomatal conductance, among others.
-
The information was related to the measurement methodology [40,41] and not to the mathematical modeling.
-
Some articles did not present the mathematical formulation for photosynthesis [60,61].
-
Some models estimated crop growth [62,63,64,65], which is not the focus of this review.
-
Some articles use previously developed models to carry out experiments different from those originally proposed [66,67,68,69], but they do not propose any new model.
-
Some authors present a mathematical formulation of the model; however, they do not present validation or results obtained [70].
-
Some authors expose models to estimate photosynthesis only for the Calvin cycle [71,72], which is only one section of all the photosynthetic processes.
-
Some authors used repeated or similar models to those presented in this manuscript [73,74,75,76].
-
Some models are developed based on soft computing [77,78]. These models were excluded because they are not analytical models, that is, they are not models that present explicit algebraic expressions.
Finally, a total of 39 reported papers were selected and reviewed, being essential for conducting this review. In addition to these criteria, the type of cultivation, the mathematical formulation, as well as the validation of the modeling were considered.
The information retrieved was analyzed to determine their corresponding classification, according to the previously described types of mathematical modeling: white box, black box, or gray box.

3. Discussion

Models in the literature reported the use of input variables such as light (photosynthetically active radiation (PAR) µmol m−2 s−1, average amount of energy incident per unit, area per unit time on a surface (Irradiance) W/m2), CO2 in the atmosphere or the leaf, oxygen in the leaf or the atmosphere, chlorophyll concentration, chlorophyll fluorescence, energy composition, source concentration energy, stomatal conductance, ambient temperature, leaf’s temperature, time in hours, the vapor pressure in the leaf, nitrogen, and phosphorus; however, others focus on the physical characteristics of the plant, such as leaf’s area index, stem diameter, plant height, leaf’s mass per area, leaf’s age, or population density.
Regarding the models’ validation, several researchers use the Farquhar model as a basis [79,80], since they were the pioneers in estimating photosynthesis. Unfortunately, the parameters in the Farquhar model are difficult to estimate [81], since they use several biochemical reactions and thus use invasive techniques for plants. Another disadvantage is that it involves a long and complex mathematical calculation, which is not favorable for implementation in measurement systems. Therefore, some other models are validated with commercial devices or even with previously approved models developed by other authors.
Another difference between the models is the terminology and the definition in the output variable, since some authors present the result not as photosynthesis, but rather as carbon assimilation, net photosynthetic rate, net photosynthesis, photosynthetic rate, gross photosynthetic rate, or as biomass; because all these variables represent or are related to the estimation of photosynthesis.
The equations of the models to estimate photosynthesis can be seen in Table 2. The equation number for each model is referenced with its corresponding author. Based on this, it can be determined that mechanistic models are complex, require a lot of construction time, are difficult to parameterize, are very unfriendly to use, and tend to be more research-oriented than application-oriented. On the other hand, deterministic models are simpler to implement, but nevertheless must be adjusted to obtain accurate results. The latter is difficult to achieve because some black box-based models are designed to apply various transformations to the input data, and debugging these models, at any stage, is not an easy task. As for the gray box models, they have an interesting approach, as new models are needed that merge the white box and black box approaches to make it easier to interpret models. The abbreviations lists used to express the models’ equations.
Complex functions are not detailed and are written as f (main variables).
Table 3 describes the main characteristics of the selected models. Models are classified according to their type: mechanistic, empirical, and gray box. Table 3 is ordered similarly to Table 2, where it can be found: the author with the equation number for each model; measurable input variables in plants and their environment, with their respective units; as well as the model corresponding to the estimation of photosynthesis and its units; the number of plants (samples); the validation method used for the modeling; and the coefficient of determination (R2).
All the investigations based on the study of mathematical modeling require extensive knowledge of the photosynthesis process in order to determine the types of variables that must be measured, as well as a knowledge of how these variables interrelate to obtain a reasonable photosynthesis estimation.
Most of the mathematical models included in this review are validated with accredited models of other authors, or by comparing them with some device for measuring photosynthesis (mainly gas exchange), some others were only simulated.
The percentage analysis of the revised mathematical models shows that 20% are gray box models, 13% are black box, and 67% are white box.
It can be observed in Figure 3 that after the year 2000, the white box models had a considerable increase, separating themselves from the other types of models. The construction of the white box models is based on previous knowledge and physical intuition [45]. Besides this, white box models are more complex to model and implement in electronic systems compared with black box models [2]. On the other hand, the empirical models (black box) have been moderately developed, probably because no previous physical information is available and they are based on experimental data, trying to follow the white box. However, this does not mean that the obtained results with the black box models are not reliable, since the application of this type of model in the estimation of photosynthesis has been successful, and it is also more flexible to implement in electronic systems compared to with white box models. [47]. Unlike white box models, black box models can solve problems by estimating an unknown function, which serves as a good guide for a linear and nonlinear dynamic system. Regarding hybrid models (gray box), the first investigations began in 1991, two decades after the other types of models. Recently, interest in gray box models has increased, as it offers the possibility of combining a white box model with a black box model. In other words, it is possible to have previous information available on some processes, but at the same time, several parameters are determined from the observed data.
The models based on white box are as good as the models based on black box, and their performance depends on the domain of the application and the input data, so it is necessary to analyze them to select the best one to apply to the given problem and the best one to show the obtained results. To this end, it is essential to understand the input data, as well as the best way to display the output data, the careful formulation of the objectives is paramount and determines the scope of the built model. It is important to mention that in the last two years (2021–2022) the development of gray box and black box models has been the subject of interest, while white box models remain outdated.
In this review, it is demonstrated, with considerable evidence, that photosynthesis can be estimated through different mathematical models. These models have various characteristics, without standardizing any specific equation or variables. Among the articles on validated mathematical models with trends in mathematical behaviors are Marshall [82], validated with a rectangular hyperbolic model; Han [93] validated with an empirical function based on the Poisson distribution; [107] validated with Euler’s simple digital integration method, rectangle hyperbolic method Perin [116] validated with the hyperbolic model and Jansen [112] with the integration model according to Euler.
It is important to note that all the measurable variables used in the classification of the revised mathematical models (empirical, mechanistic, or hybrid) have diversified units, so it is impossible to standardize the units for estimating photosynthesis. Therefore, the approach is oriented towards the analysis and comparison of the behavior trend in the photosynthetic process, regardless of the specific magnitude of the reported measurement. Then, the photosynthesis estimation units depend on the type of variables in the input, since it was found that there are different mathematical models related to biochemical, physical, and agroecological variables [14,119]. Figure 4 shows the input variables used in the reviewed mathematical models, classified according to type [120,121,122].
To implement the mathematical model in an electronic system, it is necessary to consider the physical method of measuring each variable [121,123,124]. Regarding the physical method of measurement, the biochemical, physical, and agroecological variables can be sub-classified as invasive or non-invasive. Invasive measurements interfere physically or chemically with the plant; on the contrary, non-invasive measurements do not alter the natural process of the plant because, in general, there is no contact with the studied body. Invasive measurement methods have the following disadvantages: they affect the behavior of plants, they can be destructive, depending on the laboratory, and require a space with adequate facilities. Therefore, it is not in situ and the measurement is not immediate, causing a delay. In contrast, non-invasive methods have the characteristic that they can be in situ, in vivo, and non-destructive.
According to the mathematical models reviewed, Figure 5 shows the diversity of variables that have been used to estimate photosynthesis, as well as the classification of whether they are invasive or non-invasive, when measured in plants. It also presents the percentage of the number of times a certain variable is used to build the mathematical model, reinforcing the idea that it is not possible to standardize the units of photosynthesis estimation. According to Figure 5, light (photosynthetically active radiation (PAR) µmol m−2 s−1; average amount of incident energy per unit area per unit time on a surface (Irradiance) W/m2) is the most used variable for estimating photosynthesis in mathematical models since it is required in 25% of the models, followed by ambient temperature with 14%. In general, the advantage of involving biochemical variables in the models is that photosynthesis is more reliable, since they are directly related to the photosynthetic process. However, since the information on the biochemical variables is obtained by a completely invasive process, and the type of equation to be implemented requires digital elements of a better quality, the implementation price is quite high. On the other hand, involving physical variables is less invasive since it depends on the method used for the measurement, as is the case of the leaf’s area index and leaf’s temperature variables. Finally, agroecological variables are those found in the atmosphere; these variables are completely non-invasive.
Another relevant finding is that there are other important variables for estimating photosynthesis that have not been used in previous models since they interfere in this metabolic process thus opening a niche of opportunity for future models, variables such as, minerals, processed sap (glucose), relative humidity of the air and relative humidity of the leaf.
Table 4 shows the statistics of the number of input variables used in the mathematical models. It is emphasized that 31% of the models require only 2 input variables, and 5% used 7 input variables for the estimation of photosynthesis. Thus, it is shown that it is not necessary to measure all the possible input variables to obtain a photosynthesis output value. In addition, Table 4 also shows that statistics on the number of input variables used by each mathematical model are observed according to their classification. Based on this Table, it was concluded that black box models tend to use only 2 input variables; although the estimation process is unknown, the expected result is not affected since there are even white box models that also use only two input variables to estimate the photosynthesis. These statistics confirm once again that regardless of the number of input variables and the type of modeling, it does not influence the expected photosynthesis estimation.
Murchie [125] established that photosynthesis is perhaps the most studied physiological process in plant science since it supports plant productivity. However, it is notoriously sensitive to small dynamic changes in environmental conditions; this means that quantifying nature on different time scales is not easy. For this reason, most of the mathematical models reviewed in this article measure variables in steady state, that is, in controlled environments. It is important and convenient to measure the properties of a plant under steady-state conditions, but this does not always allow accurately prediction of how the plant will respond in a complex field environment [125,126,127,128,129,130]. For the analysis of physiological processes, it is important to take into account that plants are living beings, so each plant has a different photosynthetic response [11,131]. This is due to factors such as the age of the plant, its nutrition, the type of plant, its habitat, whether it is aquatic or terrestrial [132,133], and the climate in which it is found, as well as the type of species [134,135,136,137].
In the revised models, plants of different species were used to carry out the photosynthesis estimation validations, classified for both aqueous and terrestrial environments (Figure 6).
As mentioned by Gómez [138], most of the studies on photosynthetic activity in different species have been carried out on individual leaves, without considering that due to various factors, what happens in one leaf may not be what is happening in the other leaves of the same plant and the crop in general [139,140,141]. For this reason, another important aspect to consider in the analysis and validation of a mathematical model for estimating photosynthesis is the number of plants required for the experimental part of the model, since it is more significant to experiment with a complete crop than to do so with just one plant per species [11]. In this review, the number of plants used in each article was sought; however, not all the articles specified this information, leaving in doubt whether it is a significant piece of data to evaluate the reliability of the model.
Various principles and methods have been proposed to select the best model for a system among the various existing models [88,142,143,144,145]. These quality criteria, which can be used separately or in some combination, are as follows:
i.
To have the minimum number of model parameters with a reasonable error.
ii.
To have the simplest form with minimal error.
iii.
To be based on physical, chemical, and biological laws as much as possible.
iv.
To have the minimum deviation between the predicted and empirical values.
v.
To have the minimum output variance.
According to Gutiérrez [146], another useful statistic to measure the global quality of a model is the coefficient of determination (R2) since it measures the proportion or percentage of variability in the experimental data. To interpret these coefficients, it is true that 0 ≤ R2 ≤ 1, values close to 1 are desirable. In general, for prediction purposes, an adjusted coefficient of determination of at least 70% is recommended. Although the coefficient of determination is an important factor to consider for the validation of the models, only some authors showed this evidence in the literature review.
As we have seen so far, the estimation of photosynthesis is not an easy process to measure, as many factors can modify the final results, in other words, it is not possible to use the same model for all types of plants.
Years after the publication of Farquhar, Von Caemmerer and Berry’s landmark C3 photosynthesis model, photosynthesis modeling remains an active field of research. These models have allowed the formulation and testing of new hypotheses, which has led to their refinement [147].

4. Conclusions

This article summarizes current mathematical models in the context of the photosynthetic process in plants. For this, an extensive review of the current state of the art was carried out. In addition to highlighting their advantages and limitations, this review demonstrates that mathematical models consider both, qualitative and quantitative variables. Among the qualitative variables, there is the validation method, the type of crop to carry out the tests, the mathematical model, and the input variables and estimated variables with their respective units. Conversely, the quantitative variables reported are the number of input variables, the number of samples to carry out the tests, the coefficient of determination, and the percentage equivalent to the number of published articles focused on mathematical models to estimate photosynthesis according to its classification.
The reader can compare the existing types of models according to their mathematical complexity, the input variables required for the models, and the units used to estimate photosynthesis. Based on the state-of-the-art, there is no evidence of a comparison between mathematical models.
This review shows various models, from simple to very complex, that relate mathematics to the photosynthetic process.
This review presents the diversity of mathematical models for photosynthesis estimation from previously published research, focused on the mathematical formulation, complexity, validation, applications, and the analysis of the diversity of the variables with their respective units, as well as the invasiveness required for its measurement. Once these input variables have been analyzed, we conclude that the most commonly used non-invasive variables are light, ambient temperature, carbon dioxide found in the atmosphere, and the leaf’s temperature. On the other hand, the vast majority of variables used to estimate photosynthesis are invasive for the plant, causing stress. Among the most common invasive input variables to measure photosynthesis are carbon dioxide concentration in the leaf, oxygen in the leaf, the leaf’s temperature, and pressure steam, remembering at this point that the terms “invasive” or “non-invasive” depend on the techniques necessary to acquire the variables of the model [96].
Regarding the mathematical modeling described, we have inferred that some models require a wide variety of input variables, which makes them more complex, thus hindering their future application in electronic systems. Therefore, it would also be useful to develop simpler mathematical models for easy implementation.
This review supports the idea that an estimate of photosynthesis can be obtained with different mathematical equations and different measurable input variables without standardizing. This idea may be significant for future research given that, based on the research carried out in this review, it can be argued that photosynthesis units are determined based on the input variables measured.
Mathematical models provide an essential contribution to the understanding of both, mathematical and biological fields. Deepening the knowledge of mathematical modeling to estimate photosynthesis allows us to understand that there are different methodologies to obtain the same result, or at least a similar result. In biological systems, it is not always necessary to have specific definitive values, but the tendencies of the system are also important. This information can be used to develop new mathematical models to estimate photosynthesis with new variables related to the plant’s habitat and with greater relevance to be implemented in electronic systems during the development of photosynthesis estimation equipment.
Finally, as was described in this article, the photosynthesis process is not easy to estimate since there are important factors that alter said estimation, such as the type of species; the physical, chemical and agroecological characteristics of the species, as well as their control; the number of plants needed (samples); and the measurement at leaf, plant or canopy level.

Author Contributions

Conceptualization, A.E.-C. and R.G.-C.; methodology, J.P.-O.; formal analysis, M.A.R.-L. and F.J.P.-P.; investigation, L.d.C.G.-R. and A.E.-C.; writing—original draft preparation, L.d.C.G.-R.; writing—review and editing, R.G.-C. and A.I.B.-G.; visualization, A.I.B.-G.; supervision, A.E.-C.; project administration, A.E.-C.; funding acquisition, A.E.-C. and J.P.-O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Tecnológico Nacional de México, project 2wlk40 (11309).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

Authors thank to Consejo Nacional de Ciencia y Tecnología (CONACyT), Mexico, for the Ph.D. grant of L.d.C.G.-R.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

aLeaf’s age Equation (5)
a*Total concentration of PSUs Equations (14) and (24)
ACO2 assimilation Equation (1); Photosynthetic rate Equations: (10), (12), (19) and (22); Photosynthesis Equation (18); Net photosynthesis Equation (11)
AcAssimilation rate limited by activation and kinetic properties of the Rubisco enzyme Equations (12) and (22)
AjNet photosynthesis rate limited by light Equations (12) and (22)
AnCO2 assimilation Equation (23); Photosynthetic rate Equations (26) and (38); Photosynthesis Equation (9)
ageAge of plant Equation (31)
BBiomass Equation (29)
CCarbon dioxide in the leaf Equations: (1), (10), (19), (25) and (30); Carbon dioxide from the atmosphere Equations (2), (6), (8), (18), (20), (21) and (34)
C1CO2 Concentration in the leaf Equation (3)
CaCO2 in the atmosphere Equations (12) and (23)
CbsCO2 concentration Equation (9)
CcCO2 partial pressure at the Rubisco carboxylate sites Equation (19)
CcCO2 offset point Equation (3)
CiCO2 partial pressures Equations (10) and (11); Internal CO2 concentration in the leaf Equation (23); Carbon dioxide from the atmosphere Equation (22)
ClChlorophyll Equation (13)
CLACumulative leaf area of entire crown Equation (26)
CsCO2 concentration on the leaf surface Equation (23)
CUOAccumulated absorption O3 Equation (23)
DConversion coefficient Pg Equations (6) and (20); Stem diameter Equation (29)
DO2Dissolved oxygen concentration Equations (35), (36) and (38)
EIrradiance Equations (13) and (32)
E1Light utilization efficiency Equation (3)
EaStart of photosynthesis saturation Equation (17)
EcCO2 utilization efficiency Equation (3)
EURRadiation use efficiency Equation (34)
EXTExtinction coefficient of PAR inside the cover Equation (34)
fPS2 fraction Equation (32)
FFluorescence Equation (37)
f (I, T)Function of light and temperature in the respiration of the micro algae Equation (37)
f (age)Age function Equation (31)
f (PPFD)Photosynthetic photon flux density function Equation (31)
f(SWC)Soil water content function Equation (31)
f(T1)Temperature function Equation (3)
f(Ta)Air temperature function Equation (31)
f(Ts)Soil temperature function Equation (31)
f(VPD)vapor pressure deficit function Equation (31)
F(z)Cumulative density leaf area between the surface of the canopy and height z Equation (27)
f1(T)Absolute Maximum mathematical function due to leaf’s temperature Equation (5)
f2 (a)Function that describes the pattern for the age of the leaf Equation (5)
FAO3Response relationship treatment to control photosynthesis Equation (23)
FC02Relative dimensionless factor that estimates the effect on the growth of the concentration of CO2 in the air Equation (34)
Fi,dPhotosynthesis gross Equation (33)
F’mFluorescence flux Equation (37)
gbConductance of the boundary layer of the leaf Equation (23)
gminMinimum stomatal conductance Equation (12)
gnStomatal conductance Equation (12)
gsStomatal conductance Equation (23)
HHeigh Equation (29)
iSerial number Equation (26)
ILight intensity or irradiance Equations (2), (15), (16), (17), (30), (36) and (39)
I*CO2 photosynthetic compensation point Equation (18)
i,dTime, day Equation (33)
IAFLeaf’s area index Equation (34)
IavIrradiance Equation (35)
IcIrradiance concentration Equation (15)
IlIrradiance Equation (21)
IoLight flux density in foliage Equation (20); Light intensity Equation (27); Quanta absorbed by the light Equation (11)
IpPAR incidence Equation (9)
IPARRadiation Equation (34)
IPPFPhotosynthetic photon flux Equation (25)
I’RPAR intensity of half-saturation Equation (28)
JPotential rate of full-chain electron transport Equations (10), (12) and (22)
J (CO2, I)Maximum photosynthesis rate with Cbs saturation at a given irradiance Equation (9)
JmElectron transport potential rate Equation (9)
kRepresents the fractional rate of photosynthesis vs. irradiance Equations (12) and (13)
KConstant rate of photoclimatory changes in cellular chlorophyll Equations (24) and (26)
K (I)Light extinction coefficient Equations (6), (20) and (27)
K (O)Function of enzymatic kinetics Equation (12)
K1 [RuBP]Carbon fixed by photosynthesis Equation (7)
K2 [RuBP]2Carbon used by photorespiration Equation (7)
K5 [TP]Carbon used by dark respiration Equation (7)
KaPPFD extinction on chlorophyll Equation (11)
KcMichaelis constant for carboxylation Equations (1), (9), (10), (18) and (22)
KmCRubisco constant for CO2 Equation (19)
KmORubisco constant for O2 Equation (19)
KoMichaelis constant for oxygenation Equations (1), (9), (10), (18) and (22)
KO3Leaf resistance ratio for O3 Equation (23)
kp, ksConstan Equations (14) and (24)
LFoliar area index Equation (20)
LAIFoliar area index Equation (6)
LFmaxMaximum leaf photosynthesis rate Equation (6)
LMALeaf’s mass per area Equations (26) and (38)
mLeaf light transmission coefficient Equations (6), (20) and (27)
msatRate of light saturated carboxylation Equation (11)
NpLeaf nitrogen content Equation (11)
nsatLeaf nitrogen content Equation (11)
OOxygen from the atmosphere Equations (9) and (18), Oxygen in the leaf Equations (1), (10), (12), (13), (17), (19), (22) and (24)
O(O)O2 partial pressures Equation (10)
O3Oxygen in the leaf Equation (23)
ObsO2 concentration in the bundle-sheath cell Equations (9) and (19)
PPhotosynthetic rate Equations (8), (13), (14), (15), (17), (24) and (37)
P*Photosynthesis rate without photoinhibition Equation (4)
P(T)Photosynthesis reduction factor Equation (20)
PARRadiation Equations (26), (34) and (38)
PARtPAR incident on top of crown Equation (26)
PARvpPAR intensity reaching the epidermis of the leaf Equation (28)
PCARBPotential biomass accumulation per plant Equation (34)
PgTotal gross photosynthesis Equations (6), (20) and (27)
PgmaxMaximum gross photosynthesis Equation (25)
PGRED(T)Function to modify Pg, below suboptimal temperatures Equation (6)
pHDegree of acidity or alkalinity of the plant Equations (35), (36) and (39)
PijPhotosynthesis Equation (28)
P l,gPhotosynthetic rate Equation (21)
PLTPOPPlant density per unit area Equation (34)
PmMaximum photosynthesis rate Equation (14)
Pm (O)Oxygen release rate Equation (24)
Pm (T, a)Maximum net photosynthesis rate per unit of leaf area over all PPFD levels, temperature and ages Equation (5)
Pm (T, C, N)Photosynthetic value with saturation of Il Equation (21)
PmaxMaximum photosynthetic rate Equations (2) and (3); Specific velocity coefficient for photosynthesis calculated by a saturated speed-dependent hyperbolic function of light and is a function of temperature Equations (17) and (28)
PmmMaximum endogenous photosynthesis capacity Equation (25)
PnNet photosynthetic rate Equations (2), (7), (16), (30) and (31)
PnCO2Net photosynthesis rate limited by CO2 Equation (25)
PnlPhotosynthesis rate Equation (25)
Pn1maxMaximum net photosynthesis Equation (25)
PnmaxMaximum net photosynthesis Equation (31)
PO2Net photosynthesis rate Equations (36) and (39)
PO2 (DO2)Net photosynthesis rate in function of dissolved oxygen concentration Equations (36) and (39)
PO2 (I)Net photosynthesis rate in function of irradiance Equations (36) and (39)
PO2 (pH)Net photosynthesis rate in function of pH Equations (36) and (39)
PO2(T)Net photosynthesis rate in function of temperature Equations (36) and (39)
PO 2 B Rate of gross photosynthesis Equation (32)
PPFDFlux density of photosynthetic photon Equations (6) and (31)
PpotPhotosynthetic potential rate Equation (3)
QeLeaf’s quantum efficiency Equation (6)
qpPhotochemical cooling, a measure of the fraction of open reaction centers under ambient photon precipitation rates. Equation (32)
RLeaf’s breathing Equation (8)
r*mDisappearance of PSU activated Equations (14) and (24)
RaRadiation Equation (33)
RCLAIRelative area of the accumulated leaf Equation (26)
RdDark respiration Equations (1), (2), (9), (25), (26) and (31)
Rd (C)Respiratory CO2 released Equation (18)
Rd (I)Leaf respiration in light or day Equations (12) and (22)
RDINCRelative depth in the crown Equations (26) and (38)
RlightNonphotorespiratory mitochondrial CO2 release in light Equation (16)
RO2Photosynthesis rate Equation (35)
RO2(DO2)Photosynthesis rate in function of dissolved oxygen concentration Equation (35)
RO2(I)Leaf respiration in light or day Equation (39)
RO2(Iav)Photosynthesis rate in function of irradiance Equation (35)
RO2(pH)Photosynthesis rate in function of pH Equation (35)
RO2(T)Photosynthesis rate in function of temperature Equation (35)
RuBPRibulose bisphosphate Equation (7)
Sc/oRelative CO2/O2 specificity factor for Rubisco Equation (19)
SWCSoil water content Equation (31)
tTime Equations (4) and (24), Temperature Equation (28)
TLeaf’s temperature Equations (5), (10), (20) and (22), Ambient temperature Equations (6), (15), (18), (21), (23), (30), (33), (34), (35), (36) and (39)
T1Ambient temperature Equation (3)
TaAmbient temperature Equation (31)
TairAmbient temperature Equation (38)
TSsoil temperature Equation (31)
TPTriose phosphate Equation (7)
vStoichiometric factor Equation (37)
VcmMaximum saturated CO2 carboxylation rate Equation (9)
VcmaxMaximum carboxylation rate Equations (1), (10), (18), (19) and (22)
VmMaximum carboxylation rate Equation (12)
VPDPressure steam Equation (26); Vapor pressure deficit Equations (30), (31) and (38)
WcMinimum carboxylation rates limited by Rubisco activity Equation (9)
WCAssimilation rate limited by activation and kinetic properties of the Rubisco enzyme Equation (10)
WjRuBP regeneration Equation (9)
WJRate limited by the regeneration capacity of Ribulose bisphosphate (RuBP) Equation (10)
WpCapacity of utilization of the products of photosynthesis Equation (10)
xMaximum per crop Equation (33)
x*Functional activation fraction PSUs Equation (24)
Xe*PSU steady state fraction functions activated under continuous illumination Equation (14)
zHeight Equation (27)
αQuantum efficiency Equations (2), (9), (11) and (21)
α (Cl)Specific absorption of Chlorophyll Equation (13)
α (I)Light utilization efficiency Equations (8), (20) and (27)
l Photochemical efficacy of the leaf in the absence of oxygen Equation (25)
ΓCO2 compensation point in the absence of dark breathing Equations (10), (12) and (22)
Γ *Compensation point in the absence of dark Equations (1) and (19); CO2 partial pressure for the compensation of oxygenation and carboxylation reactions Equation (11)
εRadiation use efficiency Equation (33)
nPS2Relationship between PS2 reaction centers and chlorophyll Equation (32)
θCurvature factor dimensionless 0 > θ < 1 Equations (8), (21) and (37)
θ(C)Relationship between physical and total resistance to CO2 diffusion Equation (2)
θ(C, I)Degree of curvature of the CO2 response of net light-saturated photosynthesis Equation (25)
σModel Parameter Equation (37)
σPS2Functional absorption of PS2 Equation (32)
τLeaf conductance to CO2 transfer Equations (8), (9) and (20)
ϕQuantum performance of oxygen evolution Equation (13)
ϕCIrradiance-dependent parameters Equation (32)
ΦpLight intensity Equation (37)
ΦRCQuantum performance on the PS2 Equation (32)
ψSolar elevation angle Equation (26)
ΨLight inhibition function Equation (4)

References

  1. Marín, M.A. Control Climático y Ciclo de Cultivo. Horticultura 2001, 152, 28–34. [Google Scholar]
  2. Stirbet, A.; Lazár, D.; Guo, Y.; Govindjee, G. Photosynthesis: Basics, History and Modelling. Ann. Bot. 2020, 126, 511–537. [Google Scholar] [CrossRef]
  3. Daniell, H.; Lin, C.-S.; Yu, M.; Chang, W.-J. Chloroplast Genomes: Diversity, Evolution, and Applications in Genetic Engineering. Genome Biol. 2016, 17, 134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Blankenship, R.E. Molecular Mechanisms of Photosynthesis; John Wiley & Sons: New York, NY, USA, 2021. [Google Scholar]
  5. Maxwell, K.; Johnson, G.N. Chlorophyll Fluorescence—A Practical Guide. J. Exp. Bot. 2000, 51, 659–668. [Google Scholar] [CrossRef] [PubMed]
  6. Tillery, B.; Enger, E.; Ross, F. Integratsssed Science; Higher Education; McGraw-Hill: New York, NY, USA, 2012. [Google Scholar]
  7. Shipman, J.; Wilson, J.D.; Higgins, C.A.; Lou, B. An Introduction to Physical Science; Cengage Learning: Boston, MA, USA, 2020. [Google Scholar]
  8. Azcon-Bieto, J.; Talón, M. Fundamentos de Fisiología Vegetal; McGraw-Hill: Interamericana de España, Barcelona, Spain, 2000. [Google Scholar]
  9. Cheng, Y.; He, D. A Photosynthesis Continuous Monitoring System for CAM Plants. Int. J. Agric. Biol. Eng. 2019, 12, 141–146. [Google Scholar] [CrossRef] [Green Version]
  10. West-Eberhard, M.J.; Smith, J.A.C.; Winter, K. Photosynthesis, Reorganized. Science 2011, 332, 311–312. [Google Scholar] [CrossRef]
  11. Fereres Castiel, E.; Martín, V.; Francisco, J. Fitotecnia: Principios de Agronomía para Una Agricultura Sostenible; Mundi-Prensa Libros: Madrid, Spain, 2017. [Google Scholar]
  12. Taiz, L.; Zeiger, E.; Møller, I.M.; Murphy, A. Plant Physiology and Development; Sinauer Associates Incorporated: Sunderland, MA, USA, 2015. [Google Scholar]
  13. Climático, C.; Biodiversidad, Y. Documento Técnico V Del IPCC; IPCC: Geneva, Switzerland, 2002. [Google Scholar]
  14. Bidwell, R.G.S. Fisiología Vegetal; AGT: Mexico City, Mexico, 1993. [Google Scholar]
  15. Murchie, E.H.; Pinto, M.; Horton, P. Agriculture and the New Challenges for Photosynthesis Research. New Phytol. 2009, 181, 532–552. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, Y.; Xiao, X.; Wu, X.; Zhou, S.; Zhang, G.; Qin, Y.; Dong, J. A Global Moderate Resolution Dataset of Gross Primary Production of Vegetation for 2000–2016. Sci. Data 2017, 4, 170165. [Google Scholar] [CrossRef] [Green Version]
  17. Ryu, Y.; Berry, J.A.; Baldocchi, D.D. What Is Global Photosynthesis? History, Uncertainties and Opportunities. Remote Sens. Environ. 2019, 223, 95–114. [Google Scholar] [CrossRef]
  18. Wu, A.; Hammer, G.L.; Doherty, A.; von Caemmerer, S.; Farquhar, G.D. Quantifying Impacts of Enhancing Photosynthesis on Crop Yield. Nat. Plants 2019, 5, 380–388. [Google Scholar] [CrossRef]
  19. Morales, F.; Ancín, M.; Fakhet, D.; González-Torralba, J.; Gámez, A.L.; Seminario, A.; Soba, D.; Ben Mariem, S.; Garriga, M.; Aranjuelo, I. Photosynthetic Metabolism under Stressful Growth Conditions as a Bases for Crop Breeding and Yield Improvement. Plants 2020, 9, 88. [Google Scholar] [CrossRef] [Green Version]
  20. Sinclair, T.R.; Rufty, T.W.; Lewis, R.S. Increasing Photosynthesis: Unlikely Solution for World Food Problem. Trends Plant Sci. 2019, 24, 1032–1039. [Google Scholar] [CrossRef]
  21. Shevela, D.; Bjorn, L.O. Photosynthesis: Solar Energy for Life; World Scientific Publishing: Singapore, 2018. [Google Scholar]
  22. Tenhunen, J.D.; Weber, J.A.; Yocum, C.S.; Gates, D.M. Development of a Photosynthesis Model with an Emphasis on Ecological Applications. Oecologia 1976, 26, 101–119. [Google Scholar] [CrossRef]
  23. Field, C.B.; Ball, J.T.; Berry, J.A. Photosynthesis: Principles and Field Techniques. In Plant Physiological Ecology; Springer: New York, NY, USA, 2000; pp. 209–253. [Google Scholar]
  24. Wolfbeis, O.S. Materials for Fluorescence-Based Optical Chemical Sensors. J. Mater. Chem. 2005, 15, 2657–2669. [Google Scholar] [CrossRef]
  25. Millan-Almaraz, J.R.; Guevara-Gonzalez, R.G.; Troncoso, R.R.-; Osornio-Rios, R.A.; Torres-Pacheco, I. Advantages and Disadvantages on Photosynthesis Measurement Techniques: A Review. Afr. J. Biotechnol. 2009, 8, 7340–7349. [Google Scholar] [CrossRef]
  26. Busch, F.A. Photosynthetic Gas Exchange in Land Plants at the Leaf Level. In Photosynthesis; Springer: New York, NY, USA, 2018; pp. 25–44. [Google Scholar]
  27. Walker, B.J.; Busch, F.A.; Driever, S.M.; Kromdijk, J.; Lawson, T. Survey of Tools for Measuring In Vivo Photosynthesis. Photosynthesis 2018, 1770, 3–24. [Google Scholar]
  28. Bassow, S.L.; Bazzaz, F.A. How Environmental Conditions Affect Canopy Leaf-Level Photosynthesis in Four Deciduous Tree Species. Ecology 1998, 79, 2660–2675. [Google Scholar] [CrossRef]
  29. Schulze, E.-D. A New Type of Climatized Gas Exchange Chamber for Net Photosynthesis and Transpiration Measurements in the Field. Oecologia 1972, 10, 243–251. [Google Scholar] [CrossRef]
  30. Hunt, S. Measurements of Photosynthesis and Respiration in Plants. Physiol. Plant. 2003, 117, 314–325. [Google Scholar] [CrossRef]
  31. Takahashi, M.; Ishiji, T.; Kawashima, N. Handmade Oxygen and Carbon Dioxide Sensors for Monitoring the Photosynthesis Process as Instruction Material for Science Education. Sens. Actuators B Chem. 2001, 77, 237–243. [Google Scholar] [CrossRef]
  32. Biosciences LI-6400XT System. Photosyntesis, Fluorescence, Respiration. 2014. Available online: https://www.licor.com/env/products/photosynthesis/LI-6400XT/ (accessed on 28 March 2022).
  33. Kawachi, N.; Sakamoto, K.; Ishii, S.; Fujimaki, S.; Suzui, N.; Ishioka, N.S.; Matsuhashi, S. Kinetic Analysis of Carbon-11-Labeled Carbon Dioxide for Studying Photosynthesis in a Leaf Using Positron Emitting Tracer Imaging System. IEEE Trans. Nucl. Sci. 2006, 53, 2991–2997. [Google Scholar] [CrossRef]
  34. Taiz, L. Plant Physiology, 5th ed.; Sinauer Associates Inc.: Sunderland, MA, USA, 2006. [Google Scholar]
  35. Hermand, J.-P. Photosynthesis of Seagrasses Observed in Situ from Acoustic Measurements. In Proceedings of the Oceans’ 04 MTS/IEEE Techno-Ocean’04 (IEEE Cat. No. 04CH37600), Kobe, Japan, 9–12 November 2004; Volume 1, pp. 433–437. [Google Scholar]
  36. Velten, K. Mathematical Modeling and Simulation: Introduction for Scientists and Engineers; John Wiley & Sons: New York, NY, USA, 2009. [Google Scholar]
  37. Hesketh, J.D. Predicting Photosynthesis for Ecosystem Models: Volume II; CRC Press: Boca Raton, FL, USA, 2017; ISBN 978-1-351-07595-4. [Google Scholar]
  38. Zufferey, V.; Murisier, F.; Schultz, H.R. A Model Analysis of the Photosynthetic Response of Vitis vinifera L. Cvs Riesling and Chasselas Leaves in the Field: I. Interaction of Age, Light and Temperature. VITIS 2000, 39, 19–26. [Google Scholar]
  39. Polerecky, L.; Lott, C.; Weber, M. In Situ Measurement of Gross Photosynthesis Using a Microsensor-Based Light-Shade Shift Method. Limnol. Oceanogr. Methods 2008, 6, 373–383. [Google Scholar] [CrossRef]
  40. Torres-Pacheco, I.; Padilla-Medina, J.A.; Osornio-Rios, R.A.; de Jesus Romero-Troncoso, R.; Villasentilde, C.; Rico-Garcia, E.; Guevara-Gonzalez, R.G. Description of Photosynthesis Measurement Methods in Green Plants Involving Optical Techniques, Advantages and Limitations. Afr. J. Agric. Res. 2011, 6, 2638–2647. [Google Scholar]
  41. Magney, T.S.; Frankenberg, C.; Fisher, J.B.; Sun, Y.; North, G.B.; Davis, T.S.; Kornfeld, A.; Siebke, K. Connecting Active to Passive Fluorescence with Photosynthesis: A Method for Evaluating Remote Sensing Measurements of Chl Fluorescence. New Phytol. 2017, 215, 1594–1608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Aziz, H.A.; Dianursanti, D. Production of Microalgae Nannochloropsis oculata Biomass in a Bubble Column Photobioreactor with Integrated Lighting through Adjustment of Light Intensity and Air Flow Rate. AIP Conf. Proc. 2018, 2024, 020040. [Google Scholar] [CrossRef]
  43. Kitić, G.; Tagarakis, A.; Cselyuszka, N.; Panić, M.; Birgermajer, S.; Sakulski, D.; Matović, J. A New Low-Cost Portable Multispectral Optical Device for Precise Plant Status Assessment. Comput. Electron. Agric. 2019, 162, 300–308. [Google Scholar] [CrossRef]
  44. Thornley, J.H.M.; France, J. Mathematical Models in Agriculture: Quantitative Methods for the Plant, Animal and Ecological Sciences; CABI: Wallingford, UK, 2007; ISBN 978-0-85199-010-1. [Google Scholar]
  45. Ljung, L. Black-Box Models from Input-Output Measurements. In Proceedings of the IMTC 2001. In Proceedings of the 18th IEEE Instrumentation and Measurement Technology Conference. Rediscovering Measurement in the Age of Informatics (Cat. No.01CH 37188), Budapest, Hungary, 21–23 May 2001; Volume 1, pp. 138–146. [Google Scholar]
  46. Hu, P.; Sun, Y.; Zhang, Y.; Dong, J.; Zhang, X. Application of WDNN for Photosynthetic Rate Prediction in Greenhouse. In Proceedings of the IEEE 2nd International Conference on Big Data, Artificial Intelligence and Internet of Things Engineering (ICBAIE), Nanchang, China, 26–28 March 2021; pp. 331–336. [Google Scholar]
  47. Guidotti, R.; Monreale, A.; Ruggieri, S.; Turini, F.; Giannotti, F.; Pedreschi, D. A Survey of Methods for Explaining Black Box Models. ACM Comput. Surv. 2018, 51, 93. [Google Scholar] [CrossRef] [Green Version]
  48. Ljung, L.; Glad, T. Modeling of Dynamic Systems; PTR Ptrnyice-Hall: Hoboken, NJ, USA, 1994. [Google Scholar]
  49. Von Caemmerer, S. Steady-State Models of Photosynthesis. Plant Cell Environ. 2013, 36, 1617–1630. [Google Scholar] [CrossRef]
  50. Wohlfahrt, G.; Gu, L. The Many Meanings of Gross Photosynthesis and Their Implication for Photosynthesis Research from Leaf to Globe. Plant Cell Environ. 2015, 38, 2500–2507. [Google Scholar] [CrossRef] [Green Version]
  51. Gest, H. History of the Word Photosynthesis and Evolution of Its Definition. In Discoveries in Photosynthesis; Govindjee, Beatty, J.T., Gest, H., Allen, J.F., Eds.; Advances in Photosynthesis and Respiration; Springer: Dordrecht, The Netherlands, 2005; pp. 39–42. ISBN 978-1-4020-3324-7. [Google Scholar]
  52. Carril, E.P.-U. Fotosíntesis: Aspectos Básicos. REDUCA 2011, 2, 39. [Google Scholar]
  53. Ruiz-Torres, N.; Rincon, F.; Martinez, J.; Reyna, J.; Esquivel, M. Sistemas de producción y asimilación de CO2 EN la variedad jaguan de maíz criollo mejorado. In Proceedings of the XII Simposio Internacional y VII Congreso Nacional de Agricultura Sostenible, Puebla, Puebla, 11–16 November 2013; pp. 1872–1886, ISBN 978-607-715-178-4. [Google Scholar]
  54. Carvajal, M. Investigación Sobre la Absorción de CO2 por los Cultivos más Representativos; CSIC: Madrid, Spain, 2022. [Google Scholar]
  55. Barrales, L.V.; Peña, I.R.; de la Reguera, P.B.F. Nota Científica—Validación de Modelos: Un Enfoque Aplicado (Model Validation: An Applied Approach); Instituto de Investigaciones Agropecuarias: Santiago, Chile, 2004. [Google Scholar]
  56. Medina-Peralta, S.; Vargas-Villamil, L.; Navarro-Alberto, J.; Canul-Pech, C.; Peraza-Romero, S. Comparación de medidas de desviación para validar modelos sin sesgo, sesgo constante o proporcional. Univ. Cienc. 2010, 26, 255–263. [Google Scholar]
  57. Mayer, D.G.; Butler, D.G. Statistical Validation. Ecol. Model. 1993, 68, 21–32. [Google Scholar] [CrossRef]
  58. Mitchell, P.L. Misuse of Regression for Empirical Validation of Models. Agric. Syst. 1997, 54, 313–326. [Google Scholar] [CrossRef]
  59. Paez, T.L. Introduction to Model Validation. In Introduction to Model Validation; Sandia National Lab: Orlando, FL, USA, 2009; p. 11. [Google Scholar]
  60. Zhu, X.-G.; Wang, Y.; Ort, D.R.; Long, S.P. E-Photosynthesis: A Comprehensive Dynamic Mechanistic Model of C3 Photosynthesis: From Light Capture to Sucrose Synthesis. Plant Cell Environ. 2013, 36, 1711–1727. [Google Scholar] [CrossRef] [Green Version]
  61. Zhu, X.-G.; Song, Q.; Ort, D.R. Elements of a Dynamic Systems Model of Canopy Photosynthesis. Curr. Opin. Plant Biol. 2012, 15, 237–244. [Google Scholar] [CrossRef]
  62. Bernard, O.; Rémond, B. Validation of a Simple Model Accounting for Light and Temperature Effect on Microalgal Growth. Bioresour. Technol. 2012, 123, 520–527. [Google Scholar] [CrossRef]
  63. Bitaubé Pérez, E.; Caro Pina, I.; Pérez Rodríguez, L. Kinetic Model for Growth of Phaeodactylum Tricornutum in Intensive Culture Photobioreactor. Biochem. Eng. J. 2008, 40, 520–525. [Google Scholar] [CrossRef]
  64. Moisan, J.R.; Moisan, T.A.; Abbott, M.R. Modelling the Effect of Temperature on the Maximum Growth Rates of Phytoplankton Populations. Ecol. Model. 2002, 153, 197–215. [Google Scholar] [CrossRef]
  65. Rojano-Aguilar, A.; Arias, A.; Lopez-Cruz, I. Modelos Matemáticos de Hortalizas En Invernadero: Trascendiendo La Contemplación de La Dinámica de Cultivos. Rev. Chapingo Ser. Hortic. 2005, 11, 257–267. [Google Scholar] [CrossRef]
  66. Romero Cuéllar, J.; Torrente Trujillo, A.; Obregón Neira, N.; Torres, A. Acople de un modelo de conductancia estomática, fotosíntesis y transpiración para el cultivo de maracuyá (Passiflora edulis Sims). Acta Agronómica 2013, 62, 42–51. [Google Scholar]
  67. Walker, A.P.; Johnson, A.L.; Rogers, A.; Anderson, J.; Bridges, R.A.; Fisher, R.A.; Lu, D.; Ricciuto, D.M.; Serbin, S.P.; Ye, M. Multi-Hypothesis Comparison of Farquhar and Collatz Photosynthesis Models Reveals the Unexpected Influence of Empirical Assumptions at Leaf and Global Scales. Glob. Chang. Biol. 2021, 27, 804–822. [Google Scholar] [CrossRef] [PubMed]
  68. Mengoli, G.; Agustí-Panareda, A.; Boussetta, S.; Harrison, S.P.; Trotta, C.; Prentice, I.C. Ecosystem Photosynthesis in Land-Surface Models: A First-Principles Approach Incorporating Acclimation. J. Adv. Model. Earth Syst. 2022, 14, e2021MS002767. [Google Scholar] [CrossRef]
  69. Fee, E.J. A Numerical Model for the Estimation of Photosynthetic Production, Integrated Over Time and Depth, in Natural Waters1. Limnol. Oceanogr. 1969, 14, 906–911. [Google Scholar] [CrossRef]
  70. Thornley, J.H.M. Acclimation of Photosynthesis to Light and Canopy Nitrogen Distribution: An Interpretation. Ann. Bot. 2004, 93, 473–475. [Google Scholar] [CrossRef] [Green Version]
  71. Pettersson, G.; Ryde-Pettersson, U. A Mathematical Model of the Calvin Photosynthesis Cycle. Eur. J. Biochem. 1988, 175, 661–672. [Google Scholar] [CrossRef] [Green Version]
  72. Obeid, H.; Rendall, A.D. The Minimal Model of Hahn for the Calvin Cycle. arXiv 2018, arXiv:1812.00620. [Google Scholar] [CrossRef]
  73. Kim, D.; Kang, W.H.; Hwang, I.; Kim, J.; Kim, J.H.; Park, K.S.; Son, J.E. Use of Structurally-Accurate 3D Plant Models for Estimating Light Interception and Photosynthesis of Sweet Pepper (Capsicum annuum) Plants. Comput. Electron. Agric. 2020, 177, 105689. [Google Scholar] [CrossRef]
  74. Hinojo-Hinojo, C.; Castellanos, A.E.; Llano-Sotelo, J.; Peñuelas, J.; Vargas, R.; Romo-Leon, J.R. High Vcmax, Jmax and Photosynthetic Rates of Sonoran Desert Species: Using Nitrogen and Specific Leaf Area Traits as Predictors in Biochemical Models. J. Arid Environ. 2018, 156, 1–8. [Google Scholar] [CrossRef]
  75. Tsujimoto, K.; Hikosaka, K. Estimating Leaf Photosynthesis of C3 Plants Grown under Different Environments from Pigment Index, Photochemical Reflectance Index, and Chlorophyll Fluorescence. Photosynth. Res. 2021, 148, 33–46. [Google Scholar] [CrossRef]
  76. Bernacchi, C.J.; Pimentel, C.; Long, S.P. In Vivo Temperature Response Functions of Parameters Required to Model RuBP-Limited Photosynthesis. Plant Cell Environ. 2003, 26, 1419–1430. [Google Scholar] [CrossRef]
  77. Center, B.; Verma, B.P. A Fuzzy Photosynthesis Model for Tomato. Trans. ASAE 1997, 40, 815–821. [Google Scholar] [CrossRef]
  78. Valenzuela, I.C.; Baldovino, R.G.; Bandala, A.A.; Dadios, E.P. Optimization of Photosynthetic Rate Parameters Using Adaptive Neuro-Fuzzy Inference System (ANFIS). In Proceedings of the 2017 International Conference on Computer and Applications (ICCA), Doha, United Arab Emirates, 6–7 September 2017; pp. 129–134. [Google Scholar]
  79. Farquhar, G.D.; von Caemmerer, S.; Berry, J.A. A Biochemical Model of Photosynthetic CO2 Assimilation in Leaves of C3 Species. Planta 1980, 149, 78–90. [Google Scholar] [CrossRef] [Green Version]
  80. Farquhar, G.D.; von Caemmerer, S.; Berry, J.A. Models of Photosynthesis. Plant Physiol. 2001, 125, 42–45. [Google Scholar] [CrossRef] [Green Version]
  81. Wang, Q.; Chun, J.A.; Fleisher, D.; Reddy, V.; Timlin, D.; Resop, J. Parameter Estimation of the Farquhar—Von Caemmerer—Berry Biochemical Model from Photosynthetic Carbon Dioxide Response Curves. Sustainability 2017, 9, 1288. [Google Scholar] [CrossRef] [Green Version]
  82. Marshall, B.; Biscoe, P.V. A Model for C3 Leaves Describing the Dependence of Net Photosynthesis on Irradiance. J. Exp. Bot. 1980, 31, 29–39. [Google Scholar] [CrossRef]
  83. Kano, A.; Gavel, C.H.M.V. Design and Test of a Simulation Model of Tomato Growth and Yield in a Greenhouse. J. Jpn. Soc. Hortic. Sci. 1988, 56, 408–416. [Google Scholar] [CrossRef] [Green Version]
  84. Pahl-Wostl, C.; Imboden, D.M. DYPHORA—A Dynamic Model for the Rate of Photosynthesis of Algae. J. Plankton Res. 1990, 12, 1207–1221. [Google Scholar] [CrossRef]
  85. Lieth, J.H.; Pasian, C.C. A Model for Net Photosynthesis of Rose Leaves as a Function of Photosynthetically Active Radiation, Leaf Temperature, and Leaf Age. J. Am. Soc. Hortic. Sci. 1990, 115, 486–491. [Google Scholar] [CrossRef] [Green Version]
  86. Jones, J.W.; Dayan, E.; Allen, L.H.; van Keulen, H.; Challa, H. A Dynamic Tomato Growth and Yield Model (TOMGRO). Trans. ASAE 1991, 34, 663–672. [Google Scholar] [CrossRef]
  87. Hahn, B.D. Photosynthesis and Photorespiration: Modelling the Essentials. J. Theor. Biol. 1991, 151, 123–139. [Google Scholar] [CrossRef]
  88. Pachepsky, L.B.; Haskett, J.D.; Acock, B. An Adequate Model of Photosynthesis—I Parameterization, Validation and Comparison of Models. Agric. Syst. 1996, 50, 209–225. [Google Scholar] [CrossRef]
  89. Chen, D.-X.; Coughenour, M.B.; Knapp, A.K.; Owensby, C.E. Mathematical Simulation of C4 Grass Photosynthesis in Ambient and Elevated CO2. Ecol. Model. 1994, 73, 63–80. [Google Scholar] [CrossRef]
  90. Nikolov, N.T.; Massman, W.J.; Schoettle, A.W. Coupling Biochemical and Biophysical Processes at the Leaf Level: An Equilibrium Photosynthesis Model for Leaves of C3 Plants. Ecol. Model. 1995, 80, 205–235. [Google Scholar] [CrossRef]
  91. Kull, O.; Kruijt, B. Leaf Photosynthetic Light Response: A Mechanistic Model for Scaling Photosynthesis to Leaves and Canopies. Funct. Ecol. 1998, 12, 767–777. [Google Scholar] [CrossRef]
  92. Chen, J.M.; Liu, J.; Cihlar, J.; Goulden, M.L. Daily Canopy Photosynthesis Model through Temporal and Spatial Scaling for Remote Sensing Applications. Ecol. Model. 1999, 124, 99–119. [Google Scholar] [CrossRef] [Green Version]
  93. Han, B.-P. Photosynthesis–Irradiance Response at Physiological Level: A Mechanistic Model. J. Theor. Biol. 2001, 213, 121–127. [Google Scholar] [CrossRef]
  94. Rubio, F.C.; Camacho, F.G.; Sevilla, J.M.F.; Chisti, Y.; Grima, E.M. A Mechanistic Model of Photosynthesis in Microalgae. Biotechnol. Bioeng. 2003, 81, 459–473. [Google Scholar] [CrossRef] [Green Version]
  95. Ye, Z.-P. A New Model for Relationship between Irradiance and the Rate of Photosynthesis in Oryza Sativa. Photosynth. 2007, 45, 637–640. [Google Scholar] [CrossRef]
  96. Ye, Z.-P.; Suggett, D.J.; Robakowski, P.; Kang, H.-J. A Mechanistic Model for the Photosynthesis–Light Response Based on the Photosynthetic Electron Transport of Photosystem II in C3 and C4 Species. New Phytol. 2013, 199, 110–120. [Google Scholar] [CrossRef]
  97. Bernacchi, C.J.; Rosenthal, D.M.; Pimentel, C.; Long, S.P.; Farquhar, G.D. Modeling the Temperature Dependence of C3 Photosynthesis. In Photosynthesis in Silico: Understanding Complexity from Molecules to Ecosystems; Laisk, A., Nedbal, L., Govindjee, Eds.; Advances in Photosynthesis and Respiration; Springer: Dordrecht, The Netherlands, 2009; pp. 231–246. ISBN 978-1-4020-9237-4. [Google Scholar]
  98. Yin, X.; Struik, P.C. C3 and C4 Photosynthesis Models: An Overview from the Perspective of Crop Modelling. NJAS—Wagening. J. Life Sci. 2009, 57, 27–38. [Google Scholar] [CrossRef] [Green Version]
  99. Torres, J.C.; Ríos, E.; Peña, R.; Martínez, D.; Sánchez, P.; Reyes, F.; Romero, G. Evaluación del Modelo de Jones para Desarrollar un Sistema por Computadora que Permita Determinar el Crecimiento del Frijol en Ambiente Controlado. In Proceedings of the Congreso Anual 2010 de la Asociación de México de Control Automático, Puerto Vallarta, Jalisco, México, 6–8 October 2010; p. 7. [Google Scholar]
  100. Johnson, I.R.; Thornley, J.H.M.; Frantz, J.M.; Bugbee, B. A Model of Canopy Photosynthesis Incorporating Protein Distribution through the Canopy and Its Acclimation to Light, Temperature and CO2. Ann. Bot. 2010, 106, 735–749. [Google Scholar] [CrossRef] [Green Version]
  101. Egea, G.; Verhoef, A.; Vidale, P.L. Towards an Improved and More Flexible Representation of Water Stress in Coupled Photosynthesis–Stomatal Conductance Models. Agric. For. Meteorol. 2011, 151, 1370–1384. [Google Scholar] [CrossRef]
  102. Lombardozzi, D.; Sparks, J.P.; Bonan, G.; Levis, S. Ozone Exposure Causes a Decoupling of Conductance and Photosynthesis: Implications for the Ball-Berry Stomatal Conductance Model. Oecologia 2012, 169, 651–659. [Google Scholar] [CrossRef]
  103. García-Camacho, F.; Sánchez-Mirón, A.; Molina-Grima, E.; Camacho-Rubio, F.; Merchuck, J.C. A Mechanistic Model of Photosynthesis in Microalgae Including Photoacclimation Dynamics. J. Theor. Biol. 2012, 304, 1–15. [Google Scholar] [CrossRef]
  104. Janka, E.; Körner, O.; Rosenqvist, E.; Ottosen, C.-O. A Coupled Model of Leaf Photosynthesis, Stomatal Conductance, and Leaf Energy Balance for Chrysanthemum (Dendranthema grandiflora). Comput. Electron. Agric. 2016, 123, 264–274. [Google Scholar] [CrossRef]
  105. Liu, Q.; Xie, L.; Li, F. Dynamic Simulation of the Crown Net Photosynthetic Rate for Young Larix Olgensis Henry Trees. Forests 2019, 10, 321. [Google Scholar] [CrossRef] [Green Version]
  106. Hozumi, K.; Kirita, H. Estimation of the Rate of Total Photosynthesis in Forest Canopies. Shokubutsugaku Zasshi 1970, 83, 144–151. [Google Scholar] [CrossRef]
  107. Wetzel, R.L.; Neckles, H.A. A Model of Zostera Marina L. Photosynthesis and Growth: Simulated Effects of Selected Physical-Chemical Variables and Biological Interactions. Aquat. Bot. 1986, 26, 307–323. [Google Scholar] [CrossRef]
  108. De la Torre, D.d.P.; Cabrera, J.D.; Calderón, Ó.A.; Hernández, F.J.; Pérez, J.J. Estimación de biomasa y contenido de carbono de Pinus cooperi Blanco, en Pueblo Nuevo, Durango. Madera Bosques 2007, 13, 35–46. [Google Scholar] [CrossRef]
  109. Romdhonah, Y.; Fujiuchi, N.; Takahashi, N.; Nishina, H.; Takayama, K. Empirical Model for the Estimation of Whole-Plant Photosynthetic Rate of Cherry Tomato Grown in a Commercial Greenhouse. Environ. Control Biol. 2021, 59, 117–124. [Google Scholar] [CrossRef]
  110. Shimada, R.; Takahashi, K. Diurnal and Seasonal Variations in Photosynthetic Rates of Dwarf Pine Pinus Pumila at the Treeline in Central Japan. Arct. Antarct. Alp. Res. 2022, 54, 1–12. [Google Scholar] [CrossRef]
  111. Kolber, Z.; Falkowski, P.G. Use of Active Fluorescence to Estimate Phytoplankton Photosynthesis in Situ. Limnol. Oceanogr. 1993, 38, 1646–1665. [Google Scholar] [CrossRef]
  112. Jansen, D.M. Uso de integración según Gauss para calcular la fotosíntesis en modelos de Simulación de Crecimiento de Cultivos. Rev. Matemática Teoría Apl. 1995, 2, 1–15. [Google Scholar] [CrossRef] [Green Version]
  113. Sau, F.; Cedrón, F.X.L.; Confalones, A.E.; Lizaso, J.I. Modelos de Simulación del Cultivo de Maíz: Fundamentos y Aplicaciones en España. Pastos 2012, 40, 117–138. [Google Scholar]
  114. Costache, T.A.; Acién Fernández, F.G.; Morales, M.M.; Fernández-Sevilla, J.M.; Stamatin, I.; Molina, E. Comprehensive Model of Microalgae Photosynthesis Rate as a Function of Culture Conditions in Photobioreactors. Appl. Microbiol. Biotechnol. 2013, 97, 7627–7637. [Google Scholar] [CrossRef]
  115. Ippoliti, D.; Gómez, C.; del Mar Morales-Amaral, M.; Pistocchi, R.; Fernández-Sevilla, J.M.; Acién, F.G. Modeling of Photosynthesis and Respiration Rate for Isochrysis Galbana (T-Iso) and Its Influence on the Production of This Strain. Bioresour. Technol. 2016, 203, 71–79. [Google Scholar] [CrossRef] [PubMed]
  116. Perin, G.; Bernardi, A.; Bellan, A.; Bezzo, F.; Morosinotto, T. A Mathematical Model to Guide Genetic Engineering of Photosynthetic Metabolism. Metab. Eng. 2017, 44, 337–347. [Google Scholar] [CrossRef]
  117. Liu, Q.; Xie, L.; Dong, L.; Li, F. Dynamic Simulation of the Multilayer Crown Net Photosynthetic Rate and Determination of the Functional Crown for Larch (Larix Olgensis) Trees. New For. 2021, 52, 1011–1035. [Google Scholar] [CrossRef]
  118. Sánchez Zurano, A.; Gómez Serrano, C.; Acién-Fernández, F.G.; Fernández-Sevilla, J.M.; Molina-Grima, E. Modeling of Photosynthesis and Respiration Rate for Microalgae–Bacteria Consortia. Biotechnol. Bioeng. 2021, 118, 952–962. [Google Scholar] [CrossRef]
  119. Jordan, D.B.; Ogren, W.L. The CO2/O2 Specificity of Ribulose 1,5-Bisphosphate Carboxylase/Oxygenase. Planta 1984, 161, 308–313. [Google Scholar] [CrossRef]
  120. Smith, N.G.; Dukes, J.S. Plant Respiration and Photosynthesis in Global-Scale Models: Incorporating Acclimation to Temperature and CO2. Glob. Change Biol. 2013, 19, 45–63. [Google Scholar] [CrossRef]
  121. Aguirre, E.A.; Pira, D.D.B.; Rodríguez, J.S.S.; Mesa, H.C.B.; Castro, D.A.V. Red de sensores inalámbricos para el monitoreo de variables agroecológicas en cultivos bajo invernadero. Publ. Investig. 2019, 13, 53–60. [Google Scholar] [CrossRef]
  122. Manrique, J.T.; Martínez y Ojeda, J.; Méndez, G.M.; Gil, A.M.; Martínez, A.R. LA Agroecología como Propuesta de Modelo de Producción Aplicado al Cultivo de Chile Habanero en Peto, Yucatán. Rev. Mex. Agronegocios 2014, 35, 969–978. [Google Scholar]
  123. Hawkesford, M.J.; Lorence, A.; Hawkesford, M.J.; Lorence, A. Plant Phenotyping: Increasing Throughput and Precision at Multiple Scales. Funct. Plant Biol. 2016, 44, v–vii. [Google Scholar] [CrossRef] [Green Version]
  124. Osuna, C.d.G. Sistema Basado en Brquitectura Híbrida FPGA-Uc Para la Medición Remota de las Variables Fisiológicas en Plantas. Master’s Thesis, Universidad Autonomia de Queretaro, Santiago, Mexico, 2013. [Google Scholar]
  125. Murchie, E.H.; Kefauver, S.; Araus, J.L.; Muller, O.; Rascher, U.; Flood, P.J.; Lawson, T. Measuring the Dynamic Photosynthome. Ann. Bot. 2018, 122, 207–220. [Google Scholar] [CrossRef] [Green Version]
  126. Athanasiou, K.; Dyson, B.C.; Webster, R.E.; Johnson, G.N. Dynamic Acclimation of Photosynthesis Increases Plant Fitness in Changing Environments. Plant Physiol. 2010, 152, 366–373. [Google Scholar] [CrossRef] [Green Version]
  127. Kaiser, E.; Morales, A.; Harbinson, J.; Kromdijk, J.; Heuvelink, E.; Marcelis, L.F.M. Dynamic Photosynthesis in Different Environmental Conditions. J. Exp. Bot. 2015, 66, 2415–2426. [Google Scholar] [CrossRef] [Green Version]
  128. Matthews, J.S.A.; Vialet-Chabrand, S.; Lawson, T. Acclimation to Fluctuating Light Impacts the Rapidity of Response and Diurnal Rhythm of Stomatal Conductance. Plant Physiol. 2018, 176, 1939–1951. [Google Scholar] [CrossRef] [Green Version]
  129. Poorter, H.; Fiorani, F.; Pieruschka, R.; Wojciechowski, T.; van der Putten, W.H.; Kleyer, M.; Schurr, U.; Postma, J. Pampered inside, Pestered Outside? Differences and Similarities between Plants Growing in Controlled Conditions and in the Field. New Phytol. 2016, 212, 838–855. [Google Scholar] [CrossRef]
  130. Vialet-Chabrand, S.; Matthews, J.S.A.; Simkin, A.J.; Raines, C.A.; Lawson, T. Importance of Fluctuations in Light on Plant Photosynthetic Acclimation. Plant Physiol. 2017, 173, 2163–2179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Kumar, V.; Sharma, A.; Soni, J.K.; Pawar, N. Physiological Response of C3, C4 and CAM Plants in Changeable Climate. Pharma Innov. J. 2017, 6, 70–79. [Google Scholar]
  132. Iversen, L.L.; Winkel, A.; Baastrup-Spohr, L.; Hinke, A.B.; Alahuhta, J.; Baattrup-Pedersen, A.; Birk, S.; Brodersen, P.; Chambers, P.A.; Ecke, F.; et al. Catchment Properties and the Photosynthetic Trait Composition of Freshwater Plant Communities. Science 2019, 366, 878–881. [Google Scholar] [CrossRef] [PubMed]
  133. Raven, J.A. Photosynthesis in Aquatic Plants. In Ecophysiology of Photosynthesis; Schulze, E.-D., Caldwell, M.M., Eds.; Springer Study Edition; Springer: Berlin/Heidelberg, 1995; pp. 299–318. ISBN 978-3-642-79354-7. [Google Scholar]
  134. Godin, C. Representing and Encoding Plant Architecture: A Review. Ann. For. Sci. 2000, 57, 413–438. [Google Scholar] [CrossRef]
  135. Johnson, G.N.; Lawson, T.; Murchie, E.H.; Raines, C. Photosynthesis in Variable Environments. J. Exp. Bot. 2015, 66, 2371–2372. [Google Scholar] [CrossRef] [Green Version]
  136. Nithia, S.; Shanthi, N. Enhance Solar UVB (280–320 nm) Radiation. Adv. Biores. 2018, 9, 232–239. [Google Scholar] [CrossRef]
  137. Toro, G.; Flexas, J.; Escalona, J.M. Contrasting Leaf Porometer and Infra-Red Gas Analyser Methodologies: An Old Paradigm about the Stomatal Conductance Measurement. Theor. Exp. Plant Physiol. 2019, 31, 483–492. [Google Scholar] [CrossRef]
  138. Gomez, L.; Montoya, E.C.; López, Y.; Riaño, N.M. Estimación de La Fotosíntesis En Plantas Completas y Validación de Un Modelo Matemático Para El Cafeto Coffea arabica L. In Proceedings of the Simposio Internacional Medición y Monitoreo de la Captura de Carbono en Ecosistemas Forestales, Valdivia, Chile, 18–20 October 2001. [Google Scholar]
  139. Bourgault, M.; James, A.T.; Dreccer, M.F.; Bourgault, M.; James, A.T.; Dreccer, M.F. Pot Size Matters Revisited: Does Container Size Affect the Response to Elevated CO2 and Our Ability to Detect Genotypic Variability in This Response in Wheat? Funct. Plant Biol. 2016, 44, 52–61. [Google Scholar] [CrossRef]
  140. Dambreville, A.; Griolet, M.; Rolland, G.; Dauzat, M.; Bédiée, A.; Balsera, C.; Muller, B.; Vile, D.; Granier, C.; Dambreville, A.; et al. Phenotyping Oilseed Rape Growth-Related Traits and Their Responses to Water Deficit: The Disturbing Pot Size Effect. Funct. Plant Biol. 2016, 44, 35–45. [Google Scholar] [CrossRef]
  141. Dornhoff, G.M.; Shibles, R.M. Varietal Differences in Net Photosynthesis of Soybean Leaves1. Crop Sci. 1970, 10, 42–45. [Google Scholar] [CrossRef]
  142. Anderson, J. Simulation: Methodology and Application in Agricultural Economics. Rev. Mark. Agric. Econ. 1974, 42, 9675. [Google Scholar]
  143. Gass, S.I. Feature Article—Decision-Aiding Models: Validation, Assessment, and Related Issues for Policy Analysis. Oper. Res. 1983, 31, 603–631. [Google Scholar] [CrossRef] [Green Version]
  144. Himmelblau, D.M. Process Analysis by Statistical Methods; John Wiley & Sons, Inc.: New York, NY, USA, 1970. [Google Scholar]
  145. Shannon, R.E. Systems Simulation; the Art and Science. In Proceedings of the 1998 Winter Simulation Conference, Washington, DC, USA, 13–16 December 1975. [Google Scholar]
  146. Gutiérrez Pulido, H.; de la Vara Salazar, R. Analysis and Design of Experiments; McGraw Hill: New York, NY, USA, 2004. [Google Scholar]
  147. Ubierna, N.; Cernusak, L.A. Preface: Advances in Modelling Photosynthetic Processes in Terrestrial Plants. Photosynth. Res. 2019, 141, 1–3. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Classification of mathematical models. The diagram shows the three main classifications of the white box, black box and gray box models, and their sub-classifications of the white box or mechanistic, and black box or empirical models. Mechanistic and empirical models can be deterministic or stochastic; and in turn, they can be continuous or discrete.
Figure 1. Classification of mathematical models. The diagram shows the three main classifications of the white box, black box and gray box models, and their sub-classifications of the white box or mechanistic, and black box or empirical models. Mechanistic and empirical models can be deterministic or stochastic; and in turn, they can be continuous or discrete.
Applsci 12 05537 g001
Figure 2. The three dimensions of an SQM mathematical model, where the (S) systems are ranked at the top of the bar; immediately below the bar, there is a list of objectives that the mathematical models in each of the segments can have (which is Q); at the lower end are the corresponding mathematical structures (M) ranging from algebraic equations (Aes) to differential equations (Des). (a) Classification of mathematical models between black and white box models. (b) Classification of mathematical models in the SQM space. Modified from [36].
Figure 2. The three dimensions of an SQM mathematical model, where the (S) systems are ranked at the top of the bar; immediately below the bar, there is a list of objectives that the mathematical models in each of the segments can have (which is Q); at the lower end are the corresponding mathematical structures (M) ranging from algebraic equations (Aes) to differential equations (Des). (a) Classification of mathematical models between black and white box models. (b) Classification of mathematical models in the SQM space. Modified from [36].
Applsci 12 05537 g002
Figure 3. Mathematical models reported in the literature. The graph shows the representation of the number of published articles focused on mathematical models to estimate photosynthesis over time, classified by decades from 1970 to 2022.
Figure 3. Mathematical models reported in the literature. The graph shows the representation of the number of published articles focused on mathematical models to estimate photosynthesis over time, classified by decades from 1970 to 2022.
Applsci 12 05537 g003
Figure 4. Input variables used in the reviewed mathematical models, classified by biochemical, physical, and agroecological type.
Figure 4. Input variables used in the reviewed mathematical models, classified by biochemical, physical, and agroecological type.
Applsci 12 05537 g004
Figure 5. Input variables used by the mathematical models to estimate photosynthesis; invasiveness based on their measurement; as well as the equivalent percentage to the number of times that the models use these variables.
Figure 5. Input variables used by the mathematical models to estimate photosynthesis; invasiveness based on their measurement; as well as the equivalent percentage to the number of times that the models use these variables.
Applsci 12 05537 g005
Figure 6. Plant species used in the mathematical models classified by aqueous and terrestrial environments. Refs. [39,72,73,74,76,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111].
Figure 6. Plant species used in the mathematical models classified by aqueous and terrestrial environments. Refs. [39,72,73,74,76,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111].
Applsci 12 05537 g006
Table 1. Classification of the methods used for photosynthesis estimation and their general description.
Table 1. Classification of the methods used for photosynthesis estimation and their general description.
Methods Used for Photosynthesis EstimationDescription
DestructiveInvolves cutting a whole plant or only a portion of it to estimate the photosynthetic activity based on the accumulation of dry matter in the plant, from the point of germination until it is cut [25].
ManometricConsists of direct measurement of oxygen (O2) pressure or carbon dioxide (CO2) in an isolated chamber with photosynthetic organisms [30].
ElectrochemicalUses electrochemical electrodes to measure O2, CO2, or pH in aqueous solutions of the sample to detect variations that depend on photosynthetic activity [30].
Gas exchangeConsists of isolating the sample for analysis in a closed chamber to quantify the CO2 concentration [29,31]. Concentrated CO2 gas is detected by an infrared gas sensor (called IRGA for Infra-Red Gas Analysis sensors) [32]
Carbon isotopesUses carbon isotopes such as 11C, 12C, and 14C to produce incorporated CO2 with radioactivity. This methodology is applied to analyze samples in isolated and illuminated chambers to produce a maximum fixation of radioactive CO2 during photosynthesis [33,34]. The main disadvantage is that it is destructive as it fixes a radioactive compound onto the sample; and its precision depends on lighting conditions.
Acoustic wavesBased on the principle of sound wave distortion in the medium in which waves propagate. The technique involves placing an acoustic transmitter on the seabed area where you want to monitor photosynthetic activity. The disadvantage is that it dependent on water conditions and sensitive to environmental disturbances [35].
FluorescenceWay in which a certain amount of light energy absorbed by chlorophylls is dissipated. The fluorescence emission can be analyzed and quantified, which provides information on the electron transport rate, the quantum yield and the existence of photoinhibition of photosynthesis. Indeed, fluorescence is used in various ways, and it has different applications. The interested reader is referred to reference [5,8,24].
Mathematical modelingEquation or a set of equations that represent a system’s behavior. There is a correspondence between the model variables and the observable quantities [36].
Table 2. Model equations for estimating photosynthesis.
Table 2. Model equations for estimating photosynthesis.
EquationAuthorMechanistic Mathematical Models
(1)Farquhar [79] A = V C m a x C Γ * C + K C 1 O / K O R d
(2)Marshall [82] P n = α I P m a x θ P n 1 θ α I + P m a x θ P n R d
(3)Kano [83] P p o t = P m a x I P m a x E 1 + I C 1 C c 1 E c + C 1 C c f T 1
(4)C.Pahl-Wostl [84] P t = P * t 1 + ψ t
(5)Lieth [85] P m ( T , a )   =   P m , T , a · f 1 ( T ) · f 2 ( a )
(6)Jones [86] P g = D · L F m a x · P G R E D T K   l n 1 m L F m a x + Q e · K P P F D 1 m L F m a x + Q e · K · P P F D · e x p K · L A I
(7)Hahn [87] P n = 36,000   · 44   h ·   ( k 1 R U B P k 2 R U B P 2 3 k 5 T P ) )
(8)Pachepsky [88] P = α I τ C α I + τ C R A c o c k , 1991
P = α I 1 + α 2 I 2 τ C 2 R H a r l e y T e n h u n e n , 1991
P = 2 α I 1 + α I τ C + 1 + α I τ C 2 4 α I τ C θ R   E v a n s F a r q u h a r , 1991
(9)Chen [89] A n = 1 0.5 O b s τ C b s m i n W c , W j R d
W c = V c m C b s C b s + K c 1 + K c K o ,   W j = J C b s C b s + O b s / τ ,   J = α I P 1 + α 2 I P 2 / J m 2 1 / 2
(10)Nikolov [90] A = m i n W C , W J , W P
W C = V c m a x C i Γ C i + K c 1 + O / K o ,   W p = V c m a x 2 ,   W J = J C i Γ 4.5 C i + 2 Γ
(11)Kull [91] A = 1 Γ * C i · m s a t   n s a t + α m 1 1 r r I o e k a   n 3 N s a t e k a   n 3 N p
(12)Chen [92] A = m i n A c , A j
A = 1.27 2 g n g m i n ( a 1 2 2 g n 2 g m i n 2 + c 1 2 g n g m i n 2 2 a g n + b 4 a d + 2 a g m i n + b 4 a e 1 2 + b 2 4 a c 8 a 3 2 l n 2 a g n + b + 2 a 1 2 d 2 a g m i n + b + 2 a 1 2 e )
where Aj:
a = 2.3 Γ + C a 2 ,
b = 0.4 4.3 Γ C a J + 2 2.3 Γ + C a R d
c = 0.2 J R d 2
where Ac:
a = K + C a 2 ,
b = 2 2 Γ + k C a V m + 2 C a + k R d
c = V m R d 2
for both:
d = a g n 2 + b g n + c 1 / 2
e = a g m i n 2 + b g m i n + c 1 / 2
(13)Han [93] P = α E ϕ
(14)Rubio [94] P = k p · r * m · a * k s + a * = P m · X e * k + X e *
P m = k p · r * · m
(15)Ye [95] P I = α I β I I + γ I I I c
(16)Ye [96] P n = α p I β p I I + γ p I   I R l i g h t
(17)Polerecky [39] P I = P m a x 1 e x p I / E a
(18)Bernacchi [97] A = 1 I * C C · V c m a x C + K C 1 + O / K O R d
(19)Yin [98]C3 plants
A = C c Γ * X 1 C c + X 2 R d
for the limited Rubisco part:
X1: Vcmax, X2: KmC (1 + O/KmO):
for the limited transport part
X 1 : J / 4 , X 2 = 2   Γ *
C4 plants
A = C c γ * O b s X 1 C c + X 2 O b s + X 3 R d
γ * : 0.5/Sc/o, X2: KmC/KmO, X3 = KmC
(20)Torres [99] P g = D τ C P T K l n ε K · I O + 1 m τ C α K · I O e x p K · L + 1 m τ C
(21)Johnson [100] P l , g = 1 2 θ α I l + P m α I l + P m 2 4 θ · α I l · P m 1 / 2
(22)Egea [101] A = m i n A c , A j R d
A c = V c m a x C i Γ C i + K c 1 + O K o ,   A j = J C i Γ 4 C i + 2 Γ
(23)Lombardozzi [102] A n = C a C s g b / 1.4 = C s C i g b / 1.6
A n = C a C s g b / 1.4 F A O 3 = C s C i g b / 1.6 F A O 3
F A O 3 = 1.0421 0.2399 * C U O , C U O = K O 3 g s O 3
(24)García-Camacho [103] P t = k p · r * · m · a t k s + a t = P m · x t k t + x t
(25)Janka [104] P n l = P g m a x 1 e x p l · I P P F P g m a x R d
P g m a x = P n l m a x + R d
P n l m a x = P n C O 2 + P m m P n C O 2 + P m m 2 4 θ · P n C O 2 · P m m 2 θ
(26)Liu [105] A n = a 0 · l n L M A V P D · R D I N C + a 1
1 e x p b 0 V D P + b 1 · P A R a 0 l n L M A V P D R D I N C + a 0 R d
P A R i = P A R t · e x p k i · R C L A I i s i n ψ
K = f d 0 ψ + d 1 C L A + d 2 · R D I N C e 0 + C L A + e 1
Empirical mathematical models
(27)Hozumi & Kirita [106] P g = b 0 a 0 1 α K 1 e K 1 α F z
b 0 1 m a 0 2 I 0 1 α K 2 l n 1 m + a 0 I 0 K 1 m + a 0 I 0 K e K 1 α F z
(28)Wetzel [107] P i j = P m a x t P A R v p / I R + P A R v p
(29)Pimienta de la Torre [108] B = a + b D + c D 2 + H , B = a + b D + c D 2 + d H D 2 ,
B = a + b D 2 + c H D 2 , B = a + b D + c H , B = a + D b + H c
(30)Romdhonah [109] P n = α 0 + α 1 I + α 2   T + α 3 C + α 4   V P D
(Linear model (G-model))
P n = β 0 + β 1 I + β 2   T + β 3 C + β 4   V P D + β 5 I · T + β 6 I · C + β 7 I · V D P + β 8 T · C + + β 9 T   · V D P + β 10 C · V D P
(Linear interaction model (I-model))
P n = γ 0 + γ 1 I + γ 2   T + γ 3 C + γ 4   V P D + γ 5   I 2 + γ 6   T 2 + γ 7   C 2 + γ 8   V D P I 2
(Linear squared model (S-model))
P n = δ 0 + δ 1 I + δ 2   T + δ 3 C + δ 4   V P D + δ 5 I · T + δ 6 I · C + δ 7 I · V D P + δ 8 T · C + + δ 9 T   · V D P + δ 10 C · V D P + δ 11   I 2 + δ 12   T 2 + δ 13   C 2 + δ 14   V D P I 2
(Linear interaction-squared model (IS-model))
(31)Shimada [110] P n = P n m a x · f   P P F D · f   T a · f   V P D · f   S W C · f   T s · f   a g e R d
Gray box mathematical models
(32)Kolber [111] P O 2 B E = σ P S 2 · Φ R C · q p E · ϕ C E f · n P S 2 · E
(33)Jansen [112] F i , d = F x , i , d 1 e x p ε i , d · R a , i , d / F x , i , d
F i , d = F x f 1 T i , d , ε i , d = ε x f 2 T i , d
Functions f1 and f2 differ by crop, and by crop history
(34)Sau [113]PCARB = EUR · IPAR/PLTPOP · FC02
IPAR = PAR · [1 − exp(-EXT · IAF)]
(35)Costache [114] R O 2 = R O 2 I a v · R O 2   T ·   R O 2 p H ·   R O 2   D O 2
(36)Ippoliti [115] P O 2 = P O 2 I · P O 2   T ¯ ·   P O 2 p H ¯ ·   P O 2   D O 2 ¯
(37)Perin [116] P = θ · σ · Φ P · I
ϕ P = 1 v F m F F m
(38)Liu [117] A n = ( a 0 · l n L M A V P D . R D I N C   + a 1 ) · 1 e x p a 2 · V P D + a 3 · P A R a 0 · l n L M A V P D R D I N C + a 1        a 4   · e a 5 ·   T a i r
(39)Sánchez [118] P O 2 = P O 2 I · P O 2   T ¯ ·   P O 2 p H ¯ ·   P O 2   D O 2 ¯ R O 2 I
Table 3. Major features of the models for estimating photosynthesis.
Table 3. Major features of the models for estimating photosynthesis.
Equation [Ref]Input VariablesEstimated VariableNo. of
Samples
Validation MethodR2
Mechanistic mathematical models
(1) [79]Ambient temperature [°C]
CO2 concentration in the leaf [mbar]
Light intensity
µ E m 1 s 2 μ m o l p h o t o n s m 2 s 1
Oxygen concentration in the leaf [mbar]
CO2 assimilation
μ m o l m 2 s 1
-Gas exchange-
(2) [82]CO2 in the atmosphere at [bar]
Irradiance W m 2
Net photosynthetic rate
g C O 2 m 2 h 1
-Rectangular
hyperbolic model
-
(3) [83]Ambient temperature [°C]
Concentration in the leaf g m 3
Photosynthetically active irradiation W m 2
Photosynthetic potential rate
g m m 2 s 1
60Simulation-
(4) [84]Light intensity µ E m 1 s 2
Time [min]
Photosynthesis rate
[Relative Units (bel)]
-Gas exchange system
LI-COR 6000
(First generation)
-
(5) [85]Photosynthetically active radiation PAR μ m o l m 2 s 1
Leaf’s age [days]
Leaf’s temperature [°C]
Maximum net photosynthesis rate m g m 2 s 1 15Gas exchange system
LI-COR 6000(First generation)
0.95
(6) [86]Ambient temperature [°C]
Foliar area index m o l p h o t o n / m 2 d
CO2 in the atmosphere [Ppm]
Light (photosynthetic photon flux
density, PPFD) [PAR]
Photosynthetic rate
g C H O 2 / m 2 d
264Previous research and simulation-
(7) [87]CO2 concentration in the leaf [No information]
Ribulose bisphosphate [No information]
Triose phosphate [No information]
Net photosynthetic rate
m g   C O 2 d m 2 h r 1
-Simulation-
(8) [88]Leaf’s Breathing μ m o l C O 2 m 2 s 1
Light μ m o l p h o t o n s m 2 s 1
CO2 in the atmostphere μ l l 1
Ambient temperature constant [°C]
Photosynthetic rate
μ m o l C O 2 m 2 s 1
-Analysis of the models by F-test -
(9) [89]Ambient temperature [°C]
Oxygen in the atmosphere [mbar]
CO2 in the leaf [Ppm]
Irradience μ m o l m 2 s 1
Rubisco μ m o l m o l 1
Net photosynthesis
μ m o l m 2 s 1
-Photosynthesis
measurement model
experiment by Collatz
0.95
(10) [90]CO2 in the leaf µ m o l / m o l
Oxygen in the leaf [Pa]
Leaf’s temperature [°C]
Photosynthetic rate
μ m o l m 2 s 1
-Gas exchange system
Li-COR 6200
(Second generation)
0.64
(11) [91]CO2 in the leaf [Ppm]
Light m m o l   m 2
Nitrogen m m o l   m 2
Chlorophyll m m o l   m 2
Net photosynthesis
μ m o l m 2 s 1
-Gas exchange system
Li-COR 6200
-
(12) [92]CO2 in the atmosphere µ m o l / m o l
Oxygen in the leaf [Pa]
Ambient temperature [°C]
Stomatal conductance m m s 1
Light μ m o l m 2 s 1
Net foliar photosynthesis rate μ m o l m 2 s 1 -FvCB model0.72
(13) [93]Chlorophyll m 2 m g C H l a 1
Oxygen in the leaf m o l O 2 e E i n s t e i n 1
Irradiance E i n s t e i n m 2 S 1
Photosynthetic rate m o l O 2 m g C H 1 a 1 S 1 -Empirical function based on the Poisson distribution-
(14) [94]Irradiance µ E m 1 s 2
Energy Composition m o l c e l 1 s 1
Concentration of power supplies
activated m o l c e l 1
Photosynthesis
m o l c e l 1 s 1
-J. Neal Phillips, Jr. and Jack Myers Growth Rate Model-
(15) [95]PAR irradiance μ m o l m 2 s 1
Ambient temperature [°C]
Net photosynthesis μ m o l C O 2 m 2 s 1 -Non-rectangular
and Rectangular
hyperbolic model
Binomial regression method
-
(16) [96]Nonphotorespiratory mitochondrial CO2 release in light μ m o l C O 2 m 2 s 1
Light intensity μ m o l   p h o t o n s   m 2 s 1
Net photosynthetic rate μ m o l C O 2 m 2 s 1 -Simulation0.99
(17) [39]Light intensity μ m o l p h o t o n s m 2 s 1
O2 on the leaf [mM]
Gross rate of photosynthesis m m o l m 3 s 1 2 of
microbial
1 of corals
Comparison between the same simulated and implemented model0.95
(18) [97]CO2 in the atmosphere μ m o l m o l 1
Oxygen in the atmosphere m m o l m o l 1
Ambient temperature [°C]
Photosynthesis
μ m o l m 2 s 1
-FvCB model-
(19) [98]CO2 in the leaf μ m o l C O 2 m 2 s 1
Oxygen in the leaf [µ bar]
Stomatal conductance
Net photosynthesis rate μ m o l C O 2 m 2 s 1 -FvCB model-
(20) [99]CO2 in the atmosphere
μ m o l C O 2 m o l 1
Leaf’s temperature [°C]
Light μ m o l f o t o n e s m 2 s 1
Foliar Area Index m 2 h o j a m 2 s u e l o
CO2 conductance in the leaf μ m o l C O 2 m 2 l e a f s 1
Light flux density in foilage μ m o l p h o t o n s m 2 s o i l m 1
Plant population density [m2]
Rate of gross photosynthesis g C H 2 O m 2 g 1 -Simulation-
(21) [100]Irradiance μ m o l f o t o n e s m 2 s 1
Ambient temperature [°C]
CO2 in the atmosphere
μ m o l C O 2 m o l 1
Nitrogen
Photosynthetic rate μ m o l C O 2 m 2 s 1 -Photosynthesis measurement model by Johnson-
(22) [101]CO2 in the leaf μ m o l m o l 1
CO2 in the atmosphere μ m o l m o l 1
Oxigen in the leaf [Pa]
Leaf’s temperature [°C]
Light μ m o l m 2 s 1
Photosynthetic rate
μ m o l m 2 s 1
-FvCB model-
(23) [102]CO2 concentration in the atmosphere μ m o l C O 2 m o l 1
CO2 concentration on the leaf surface μ m o l C O 2 m o l 1
Internal CO2 concentration in the leaf μ m o l C O 2 m o l 1
Stomatal conductance m o l H 2 O m 2 s 1
Light μ m o l f o t o n e s m 2 s 1
Ambient temperature [°C]
Oxygen in the leaf m m o l m 2
Carbon assimilation μ m o l C O 2 m 2 s 1 5Gas exchange system
LI-COR 6400
(Third generation)
-
(24) [103]Oxygen in the leaf m g O 2 C e l l 1 s 1
Disappearance of PSU activated μ m o l P S U C e l l 1 s 1
Total concentration of PSUs [PSU Cell−1]
Time [s]
Photosynthesis rate
m g O 2 C e l l 1 s 1
-Mathematical
modeling previously developed by
García-Camacho
-
(25) [104]Leaf’s temperature [°C]
Light μ m o l m 2 s 1
CO2 concentration in the leaf μ m o l m o l 1
Photosynthesis rate
μ m o l m 2 s 1
80Biochemical model of photosynthesis-
(26) [105]Radiation [PAR]
CO2 in the leaf μ m o l m 2 s 1
Leaf’s temperature [°C]
Pressure steam [KPa]
Leaf’s mass per area [gm−2]
Photosynthetic net rate
μ m o l m 2 s 1
5Gas exchange system
LI-COR 6400
(Third generation) andJackknife technique
0.85
Empirical mathematical models
(27) [106]Light intensity [Klux]
Height [m]
Total gross photosynthesis m g C O 2 / d m 2 l a n d a r e a h r 5 species
Plot (Q 15, 20 m × 20 m)
Monsi Saeki’s
mathematical model
0.60
(28) [107]Ambient temperature [°C]
Irradiation µ E m 1 s 2
Photosynthesis
m g C g C 1 h o r a 1
-Euler’s simple digital integration method, rectangle hyperbolic method-
(29) [108]Stem diameter [cm]
°H: Heigh [m]
Biomass
[Kg]
663Comparison between the same models0.99
(30) [109]Photosynthetically active radiation W m 1
Air temperature [°C]
CO2 concentration μ m o l m o l 1
Vapor pressure déficit [KPa]
Net photosynthetic rate
μ m o l s 1 p e r   p l a n t
-Test data0.82
(G-model)
0.91
(I-model)
0.90
(S-model)
0.92
(IS-model)
(31) [110]Photosynthetic photon flux density μ m o l m 2 s 1
Air temperature [°C]
Soil temperature [°C]
Vapor pressure deficit [Kpa]
Soil water content [unitless]
Age [Year]
Photosynthetic rate
μ m o l C O 2 m 2 s 1
-Simulation0.86
Gray box mathematical models
(32) [111]Irradiance µ E m 1 s 2
Fluorescence m g C m g C H l 1 h 1
Rate of gross photosysthesis m o l C H l m o l O 2 1 61“Pump and probe technique”
Forometry
0.86
(33) [112]Ambient temperature [°C]
Radiation J m 2 S 1
Time [ hours = i, day = d]
Photosynthesis gross k g C O 2 h a 1 h o r a 1 -Integration model according to
Euler
0.96
(34) [113]Radiation g M J 1
CO2 in the atmosphere
Plants surface [plant m−2]
Ambient temperature [°C]
Potential biomass accumulation per plant
[g planta−1 d−1]
-Simulation modelNo information
(35) [114]Irradiance μ E m 2 s 1
Temperature [°C]
pH [unitless]
Dissolved oxygen concentration m g l
Photosynthesis rate
m o l   O 2   g   b i o m a s s
-Data obtained by simultaneous
modifications of variables
0.74
(36) [115]Irradiance μ E m 2 s 1
Temperature [°C]
pH [unitless]
Dissolved oxygen concentration m g l
Net photosynthesis rate
m g   O 2   g   b i o m a s s · h
-Experimental data assembly from an industrial-scale outdoor tubular photobioreactor (T-PBRs)0.89
(37) [116]Light μ m o l f o t o n e s m 2 s 1
Fluorescence [chl]
Photosynthesis
g O 2 / g c h l h
-Hyperbolic model0.85
(38) [117]Leaf mass per area g m 2
Photosynthetically active radiation μ m o l m 2 s 1
Air temperature [°C]
Vapor pressure defcit [KPa]
Net photosynthetic rate
μ m o l m 2 s 1
5 treesGas exchange systemLI-6400XT0.87
(39) [118]Irradiance μ E m 2 s 1
Temperature [°C]
pH [unitless]
Dissolved oxygen m g l
Net photosynthesis rate
m g   O 2   g   b i o m a s s · h
-Data obtained by simultaneous
modifications of variables
0.60
(33) [112]Ambient temperature [°C]
Radiation J m 2 S 1
Time [ hours = i, day = d]
Photosynthesis gross k g C O 2 h a 1 h o r a 1 -Integration model according to
Euler
0.96
(34) [113]Radiation g M J 1
CO2 in the atmosphere
Plants surface [plant m−2]
Ambient temperature [°C]
Potential biomass accumulation per plant
[g planta−1 d−1]
-Simulation modelNo information
(35) [114]Irradiance μ E m 2 s 1
Temperature [°C]
pH [unitless]
Dissolved oxygen concentration m g l
Photosynthesis rate
m o l   O 2   g   b i o m a s s
-Data obtained by simultaneous
modifications of variables
0.74
(36) [115]Irradiance μ E m 2 s 1
Temperature [°C]
pH [unitless]
Dissolved oxygen concentration m g l
Net photosynthesis rate
m g   O 2   g   b i o m a s s · h
-Experimental data assembly from an industrial-scale outdoor tubular photobioreactor (T-PBRs)0.89
(37) [116]Light μ m o l f o t o n e s m 2 s 1
Fluorescence [chl]
Photosynthesis
g O 2 / g c h l h
-Hyperbolic model0.85
(38) [117]Leaf mass per area g m 2
Photosynthetically active radiation μ m o l m 2 s 1
Air temperature [°C]
Vapor pressure defcit [KPa]
Net photosynthetic rate
μ m o l m 2 s 1
5 treesGas exchange system
LI-6400XT
0.87
(39) [118]Irradiance μ E m 2 s 1
Temperature [°C]
pH [unitless]
Dissolved oxygen m g l
Net photosynthesis rate
m g   O 2   g   b i o m a s s · h
-Data obtained by simultaneous
modifications of variables
0.60
For each model, the equation number (Equation) is reported in Table 2. The symbol (-) means no information.
Table 4. Reported input variables in mathematical models for estimating photosynthesis. The table shows the representation of the percentage equivalent to the number of input variables used by the reviewed mathematical models for white, black, or gray box models.
Table 4. Reported input variables in mathematical models for estimating photosynthesis. The table shows the representation of the percentage equivalent to the number of input variables used by the reviewed mathematical models for white, black, or gray box models.
Mathematical Models
Number of Input VariablesWhite Box ModelsBlack Box ModelsGray Box ModelsPercentage Equivalent to the Number of Input Variables Used by Models
78%--5%
515%20%-13%
423%20%50%28%
331%-13%23%
223%60%38%31%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

García-Rodríguez, L.d.C.; Prado-Olivarez, J.; Guzmán-Cruz, R.; Rodríguez-Licea, M.A.; Barranco-Gutiérrez, A.I.; Perez-Pinal, F.J.; Espinosa-Calderon, A. Mathematical Modeling to Estimate Photosynthesis: A State of the Art. Appl. Sci. 2022, 12, 5537. https://doi.org/10.3390/app12115537

AMA Style

García-Rodríguez LdC, Prado-Olivarez J, Guzmán-Cruz R, Rodríguez-Licea MA, Barranco-Gutiérrez AI, Perez-Pinal FJ, Espinosa-Calderon A. Mathematical Modeling to Estimate Photosynthesis: A State of the Art. Applied Sciences. 2022; 12(11):5537. https://doi.org/10.3390/app12115537

Chicago/Turabian Style

García-Rodríguez, Luz del Carmen, Juan Prado-Olivarez, Rosario Guzmán-Cruz, Martín Antonio Rodríguez-Licea, Alejandro Israel Barranco-Gutiérrez, Francisco Javier Perez-Pinal, and Alejandro Espinosa-Calderon. 2022. "Mathematical Modeling to Estimate Photosynthesis: A State of the Art" Applied Sciences 12, no. 11: 5537. https://doi.org/10.3390/app12115537

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