Next Article in Journal
The Omics Hunt for Novel Molecular Markers of Resistance to Phytophthora infestans
Next Article in Special Issue
In Vitro Propagation, Genetic Assessment, and Medium-Term Conservation of the Coastal Endangered Species Tetraclinis articulata (Vahl) Masters (Cupressaceae) from Adult Trees
Previous Article in Journal
Transcriptomic and Physiological Response of Durum Wheat Grain to Short-Term Heat Stress during Early Grain Filling
Previous Article in Special Issue
Embryogenesis of European Radish (Raphanus sativus L. subsp. sativus Convar. Radicula) in Culture of Isolated Microspores In Vitro
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effect of LED Lighting on Physical Environment and Microenvironment on In Vitro Plant Growth and Morphogenesis: The Need to Standardize Lighting Conditions and Their Description

by
Araceli Barceló-Muñoz
1,*,
Marta Barceló-Muñoz
1 and
Alfonso Gago-Calderon
2,*
1
IFAPA—Centro de Málaga, Cortijo de la Cruz s/n, 29140 Malaga, Spain
2
Proyectos de Ingeniería, Departamento de Expresión Gráfica Diseño y Proyectos, Universidad de Malaga, 29071 Málaga, Spain
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(1), 60; https://doi.org/10.3390/plants11010060
Submission received: 18 November 2021 / Revised: 18 December 2021 / Accepted: 22 December 2021 / Published: 25 December 2021
(This article belongs to the Special Issue Advances in Plant Regeneration Ⅱ)

Abstract

:
In the last decades, lighting installations in plant tissue culture have generally been renewed or designed based on LED technology. Thanks to this, many different light quality advances are available but, with their massive implementation, the same issue is occurring as in the 1960s with the appearance of the Grolux (Sylvania) fluorescent tubes: there is a lack of a methodological standardization of lighting. This review analyzes the main parameters and variables that must be taken into account in the design of LED-based systems, and how these need to be described and quantified in order to homogenize and standardize the experimental conditions to obtain reproducible and comparable results and conclusions. We have designed an experimental system in which the values of the physical environment and microenvironment conditions and the behavior of plant tissue cultures maintained in cabins illuminated with two lighting designs can be compared. Grolux tubes are compared with a combination of monochromatic LED lamps calibrated to provide a spectral emission, and light irradiance values similar to those generated by the previous discharge lamps, achieving in both cases wide uniformity of radiation conditions on the shelves of the culture cabins. This study can help to understand whether it is possible to use LEDs as one standard lighting source in plant tissue culture without affecting the development of the cultures maintained with the previously regulated protocols in the different laboratories. Finally, the results presented from this caparison indicate how temperature is one of the main factors that is affected by the chosen light source.

1. Introduction

Until recently, white light (in cool, daylight, neutral or warm versions) or Grolux fluorescent lamps were used almost universally to supply light in in vitro growth rooms, among other reasons because of the uniform light intensity they emit. Nowadays, these fluorescents have given way to LED luminaires that, among many other options, allow us to choose the specific wavelengths we want to send to the plants depending, for example, on the species or process of interest.
In the following chapters, a technological review of fluorescent and LED lamps is presented from a general point of view and, more specifically, in terms of their possibilities related to in vitro culture. Next, a review of recently published studies that include the use of different alternatives with LEDs on in vitro plant tissue culture will be considered, focusing on the lighting variables of the installations used in each case which have more relevance on the plants. The outcome of this review will be used as the basis to establish appropriate characterization parameters of these light settings, highlighting the advantages and drawbacks of the different alternatives found, and their critical analysis.
Finally, an experimental setup with Grolux and a LED lighting system calibrated to offer a similar spectral emission was examined. This generated a controlled environment in which to apply the results of the previous characterization protocol and to compare the results obtained with these two types of artificial lighting technologies.

1.1. Fluorescent Lamps

In 1901, the American Peter Cooper Hewitt (1861–1921) patented the first mercury vapor lamp technology (U.S. Patents No. 682,690/9) and the company General Electric developed the first practical and viable fluorescent lamp (U.S. Patent No. 2,259,040) that was first sold as a general market product in 1938.
These lamps consist of a watertight glass tube with a phosphor coating that contains a mixture of noble gases and mercury vapor in a low-pressure environment. The mercury content, which according to the ROHS and WEEE European Directives must be lower than 5.0 mg per lamp, makes its waste toxic and therefore its disposal and recycling must be properly managed [1].
The collision of electrons emitted by two tungsten filaments with the mercury atoms activates this gas and generates UV radiation which is projected towards the phosphor coating that absorbs and fluorinates them, emitting other radiations in the visible spectrum. The filaments consume a significant amount of electrical power generating heat. The energy transformation in a fluorescent lamp is presented in Table 1 [2].
Adapting the operation of these lamps to use Alternating Current (AC) from the electric consumption grid requires additional elements that manage both the start-up and the variation of the alternating behavior of the electrodes as anodes and cathodes. Thus, either one of these options is needed:
  • Magnetic starter and ballast. These allow temperature increases of the lamp gas for its ignition, increasing the grid voltage with a reactance up to more than 1000 VAC. They can be accompanied by a capacitor to compensate the permanent inductive effect generated by the ballast coil. Dead cycles of the sinusoidal power signal generate a stroboscopic effect: in a 50 Hz network, the light oscillates 100 times per second, passing through its neutral position twice every 20 ms.
  • Electronic ballast. This technology replaces the ballast coil of the magnetic equipment to generate the starting voltage peak with electronic circuits that transform the network signal to high frequency voltage and currents (20–50 kHz). This allows generating light with an emission flicker at this same rate, being less perceptible to the human eye. This increase in working frequency improves the energy efficiency of the tubes by reducing their consumption and lowering the degradation of their light flux. Thus, a fluorescent tube that consumes 18 W working with a magnetic ballast will only need approximately 16 W with a high-frequency electronic device.
From the point of view of energy efficiency, this auxiliary component introduces some additional consumption in the system due to the joule effect, which can be estimated as indicated in Table 2 [3].
Another aspect that affects the energy efficiency of the fluorescent lamp technology is the working temperature. In a generalized way, they are optimized for a nominal ambient temperature of 25 °C in which they achieve their optimum performance. With this configuration, in an environment at 45 °C, its efficiency in terms of the light output performance decreases by 20% and, conversely, at 10 °C it also drops by 25% [4].

1.2. Grolux Fluorescence Technology

Fluorescent lamps developed for people in domestic or working environments (offices and factories) focus most of their radiant energy within the visible spectrum band of the green and yellow colors. That is because it is in this segment where the human eye has the greatest sensitivity. However, the radiation emitted by these discharge lamps can be significantly modified using different combinations of phosphor films to suit other purposes.
Thus, the Sylvania Company began to commercialize a new version of fluorescent lamps in September 1961 under the trade name of Grolux (US SN 154,928) [5]. These fluorescent lamps emit most of their energy in two significant chromatic peaks: 435 nm (blue) and 660 nm (red). Since then, these Grolux lamps have been used as reference sources for in vitro plant tissue culture investigations and production. The visual sensation perceived by the human eye is slightly outside the range of the definition of the white color established by the International Lighting Committee (CIE of his name in French Commission Internationale de l’Éclairage) [6] and can be classified as an emission in the purple range. The Correlated Color Temperature (CCT) generated by this lamp could be assimilated to a neutral white light (4125 K) but with a very low Color Rendering Index (CRI) (37). This indicates that this technology has a very limited capacity to allow the perception of the complete set of colors available in the visible spectrum [380–780 nm] (see Figure 1).
Some of the most significant characteristics of these tubes according to their official technical data sheets at the time of development of this publication are presented in Table 3.
There is a variant known as Grolux WS (wide spectrum) that generates an emission spectrum of light in a more distributed way, adding more emission peaks, generating a more complete sweep of photons for a more natural color reproduction. Thus, while the Grolux light has a purple glow with a slight red appearance, the Grolux WS is white in appearance with a CCT of 3700 K. Nevertheless, this second technology has no relevant presence in the scientific literature concerning in vitro plant tissue culture procedures.
During the first years of the implementation of Grolux technology, studies such as those by Lammerts [7] or Helson [5] remarked how the lack of a methodological standardization of the tests led to studies on the same plant, using both Grolux and conventional fluorescence lamps to compare their performance and produced substantially different results. Helson [5] determined that different conditions of homogeneity of the light distribution or levels of radiance (i.e., number and power of the lamps used or separation between the light sources and the plants) marked these differences.

1.3. LED Lamps

At the end of the 20th century, experiments with LED lamps related to in vitro plant growth and morphogenetic processes began to be published and, in the first decade of the 21st century, in vitro culture growth chambers are being renewed, massively replacing fluorescent lamps with others based on LEDs. This renovation is done based on lower operation energy consumption and heat generation, and an advance radiant wavelength configuration capability that can be adapted specifically to optimize the biological processes of every analyzed species [8].
The LEDs, semiconductor electronic devices, are built by putting together two pieces of crystals, generally siliceous (Si), in which dopants have been added in a controlled way with an excess (N) or deficiency (P) of electrons in their valence layer. This union is capable of emitting light with a specific monochromatic wavelength defined by Plank’s law with respect to the energy difference between the valence levels of the electron of these dopants, when it is directly polarized by passing an electric current through its two sections.
The diodes were the first-born element of electronic technology for telecommunication purposes, with the first patent of these devices was registered between 1904 and 1906. However, as early as 1907, H.J. Round highlighted the observation of light emission phenomenon in these types of semiconductors. Nevertheless, it was not until the second half of the 1920s when Oleg Vladimirovich Losev [9] began to understand the reasons why light appeared in these devices. However, was is not until well into the 21st century that they have been standardized as devices with which to build robust and efficient lighting equipment.
Nowadays there are packages for both monochromatic lights, within the entire visible spectrum, as well as white light (continuous spectrum). This is thanks to the combination of high-energy blue spectrum LEDs (440–450 nm) and the use of photon conversion layers with phosphor using the same principle as the fluorescent light (White Light Phosphor Converted or WPC), to finally emit between 15–45% of green light. These last devices are very economical due to their massive production for general lighting equipment and take advantage of the beneficial effects of green light on plant physiology, photosynthesis and growth [10].
Beyond the spectrum of the light emitted, a wide variety of LED models or packages can be found on the market depending on different parameters and fields of application: luminous power, energy efficiency, light emission angles, thermal dissipation capacity, etc. From the point of view of power, they are usually classified into four families: High Brightness [HB] (range: 40–100 mW); Mid Power [MP] (range 0.2–0.5 W); High Power [HP] (range 1–3 W) and Chip on Board [COB] arrays (range: 3–100 W).
The circulation of current through the LED generates the aforementioned emission of photons but also causes electrical losses in the inner and outer connections and within the silicon crystal itself. This has a tendency to absorb part of the radiation generated, increasing its thermal energy and, consequently, its temperature. This increase in temperature, in the long term, progressively degrades the structure of the crystal, shortening its lifetime more markedly as the values that are reached become higher and, with immediate effect, causes the crystal to have a greater tendency to continue capturing photons. Therefore, a cooling system is necessary that prevents a possible catastrophic failure from excessive positive feedback [11].
HB LEDs are designed to provide pulsating variable visual information. They are manufactured from plastic and epoxy resin enclosures without any specific heat evacuation systems. This limits their useful life in the event of prolonged continuous lighting. In this mode of use, both the life as well as the efficiency and the amount of light emitted by the HB LEDs will be considerably reduced after approximately 10,000 h of use [12]. The rest of the LED families specifically contemplate heat dissipation systems. Nevertheless, they are developed intensively in the HP and COB models incorporating encapsulation with specific thermal dissipation channels. Their mission is to connect the substrate where the silicon crystal is placed with a metallic plane of the printed circuit board coupled to the heatsink of the luminaire, which are mostly metallic (aluminum) to favor the evacuation of heat to the ambient air. In this way, the average life established for these devices is an order of magnitude higher, up to 100,000 h. Therefore, in industrialized applications of continuous production, these LED models should be used to guarantee the durability and profitability of investments in lighting equipment, avoiding HB packages [13,14].
LEDs work naturally with an electric current in continuous mode (Direct Current or DC) and at very low voltage. Consequently, they also need auxiliary equipment to adapt the AC voltage of the electric grid to their required operation values. These power supplies, or drivers, are mostly AD-DC systems with Constant Current (CC) outputs that adapt to the quantity and power of the needs of the lamp’s emitters. They have an electrical efficiency of between 85% for low power lamps up to values close to or even higher than 95% [15]. Thus, the losses of the auxiliary equipment of LED lamps are lower than those needed by the discharge lamps. Moreover, the standard flicker of the light is minimum, as the output DC values of the drivers have very limited fluctuations.
The unreliability of these LED drivers was, at first, a limiting factor in the global life of lighting equipment, but technological developments in this field have already made it possible to guarantee lifetime expectations in the vicinity of 100,000 h [16], matching the values of the HP LED emitters.
Since 1997, when Miyashita et al. published the use of LED emitters to study the effect of red light on growth and morphology of Solanum tuberosum plants cultured photoautotrophically and photomixotrophically [17], interest in this type of lighting has been increasing and a large number of works have been published with different objectives and species. This information has been collected in reviews which, in addition to describing LED technology, have focused on aspects such as: effect of LED light source on in vitro plant growth and morphogenesis [18,19,20]; secondary metabolite [18,19,20]; production and related gene expression [8,20,21]; light and in vitro competence for photosynthesis, cell wall biosynthesis and in vitro propagation of woody species [8].
However, as [8] also pointed out, the primary issues regarding the lack of protocol reproducibility among laboratories are environmental factors, light being one of those. Thus, with the implementation of LEDs the same thing is occurring as in the 1960s with the Grolux fluorescent tubes, when Lammerts [7] and Helson [5] pointed out the need for a methodological standardization of lighting conditions.

2. Review Objectives

The main parameters and variables that must be taken into account in LED-based systems will be analyzed due to the wide range of possibilities they offer, contrary to what happens with Grolux fluorescent tubes with which the capability to configure factors is very scarce. These parameters and variables will therefore require a description and quantification under experimental conditions in order to homogenize and standardize the experiments and, thus, to obtain reproducible and comparable results and conclusions.
This paper reviews recently published studies that include the use of LEDs on in vitro plant tissue culture, to know, among the most relevant parameters in terms of lighting conditions, which are described in the experimental systems of the reviewed papers and what methods are used to measure or characterize these parameters. Next, the advantages and drawbacks of these alternatives used are highlighted.
In order to use LEDs as a standard lighting source in in vitro growth chambers for the maintenance of previous in vitro culture procedures without affecting their behavior, it is desirable to be able to reproduce the lighting conditions universally used in in vitro culture using LED technology. For this objective, it will be necessary to (i) configure the LED lamps so that they emit in the same spectrum as the reference Grolux fluorescent tubes, (ii) to configure desired light irradiance values and (iii) achieve large uniformity of radiation conditions on the shelves of the culture cabins.
This review is contrasted with the results obtained in an experimental system in which artificial lighting is supplied in cabins with, in some cases, Grolux fluorescent lamps (Sylvania, OH, USA) and, in others, with a LED lighting system calibrated to offer a spectral emission, irradiation level and uniformity similar to that generated by previous discharge lamps.
A greater perspective to understand the effect of LED lighting on in vitro plant growth and morphogenesis is intended which may allow the characterization of the response of each species to establish optimum intensity, distribution photometry and wavelength conditions for in vitro plant tissue culture. This will determine the most suitable LED lighting equipment and configuration for each species and biological processes.

3. Characterization of LED Luminaires for Use in Plant Tissue Culture

The large number of works published in recent years using LED technology in plant tissue culture (Table 4) show that interest in this type of light source continues to grow. This is due to the fact that the use of LEDs has opened a huge field of possibilities where light has gone from being a determining factor for photomorphogenesis, but with little manipulation capacity, to being a component that can be configured very specifically, allowing the control and manipulation of in vitro plant growth and morphogenesis.
However, while the evaluation parameters of the results of the studies and investigations analyzed are, in general, very similar or equivalent, the description of the experiments carried out reveals a clear heterogeneity of configurations and approaches from the point of view of the lighting installations. This coincides with what is indicated by Batista et al. [21], who pointed out that the primary issues regarding the lack of a reproducibility protocol among laboratories are environmental factors and that light (in quantity and, particularly, in quality) is one of those main factors.
The lighting aspects involved in the photonic processes of enhancement of biological processes are multiple and their adequate standardization or, at least, their precise characterization in experiments is essential to allow a comparison analysis between different results (as they are measured under equivalent conditions).
The most relevant design fields to take into account in this type of installation are described and analyzed in the next chapter. While in systems based on Grolux fluorescent tubes there is very little capacity to configure working parameters, apart from the mere selection of the number of lamps, their electrical power and the separation distance from the plants to achieve the radiation level desired, there are a wider range of possibilities with LED lamps.
We propose a list of possible parameters and variables to be characterized in LED installations due to their influence on the development of in vitro plant tissue cultures. Later, the information found in the description given in a comprehensive review of scientific publication will be compared with this list. The advantages and disadvantages of the different alternatives used for the descriptions of the lighting equipment and how it was configured will be highlighted. We consider that it is essential to establish these criteria to homogenize and standardize the experiments that are detailed, indicating in the description of the installation of LED luminaires, at least, the following elements using the units of measurement that facilitate more precise comparisons.
Of all these variables, two are fundamental: (A) the precise clarification of the wavelengths of the radiations that are emitted (considering three different cases of lamps with: one type of monochromatic emitter, with a combination of these and with a set of diodes that generates a continuous spectrum) and (B) the values of irradiance that are projected on the plants as a measure of the amount of energy that is being injected. However, related to the two previous variables, the guarantee of (C), the uniformity of the lighting conditions achieved on the shelves in the growth chambers is also needed. Due to the discrete nature of the encapsulated LEDs and their directional emission of light, this requires detailed planning to guarantee adequate values.

3.1. Characterization of the Emission Spectrum

This information can be obtained from the LED or lamp manufacturer or, more exactly, measured specifically with a spectrometer. It should be specified precisely and specifically:
  • In monochromatic emissions, the wavelength value of the emitters used must be indicated without simplifications of those group values, i.e., defining a radiation spectrum not by its peak wavelength but by the generic name of its color as red (that globally identifies all the range between 600–700 nm). Based on the widths of the radiation peaks offered by commercial LED packages and the variability found by the precision of the manufacturing mechanisms of these solid-state emitters, we consider that a sufficient value precision is to contemplate differential steps of 10 nm (e.g., 650, 660, 670 nm).
  • In emissions that combine several monochromatic light sources, each color should be described as in the previous case. Moreover, the ratio of each type of radiation should not be assessed solely on the basis of the relative number of LEDs of each of the colors. Since the efficiency, power, thermal behavior and degradation curve of each type of LED is not homogeneous, this cannot be considered a comparable reference. Thus, the relationship between the light sources should be characterized by the energy radiated in each frequency. This information is offered by the graphical representations of spectra of the light that can be obtained with a spectrometer or can be obtained measuring with the PAR meter the radiance of each color independently.
  • In the case of lamps with continuous frequency range emissions within the radiant spectrum, i.e., white light, it would be necessary either to establish their commercial identification, if it exists, or to show, preferably, the radiation emission diagram of the light source used obtained with a spectrometer, which gives the exact information proportions of each emitted radiation. However, since it is very difficult to accurately compare two continuous data curves from printed images, the specific values of the Correlated Color Temperature (CCT) and the Color Rendering Index (CRI) are also significant characterization elements of this type of emissions.
The CCT or the color temperature of a light source is the temperature that makes an ideal black body that radiates light of a color comparable to that of a light source. From a practical point of view, in LED lamps this variable gives a general idea of the proportion of radiation in the blue spectrum (450 nm) and the green/yellow spectrum: more blue than green in 5000 K and higher; equivalent peaks of radiation in 4000 K and more relevance for green emissions in 3000 K and lower.
The CRI measures the reproducibility of the colors of a light source in comparison with the incandescent bulb that is used as a reference of a continuous spectrum of emission from a black body. Because the red color segment has little influence on this CCT measurement, this variable, for white LED light, indicates that at values above 80 the red component of the spectrum reaches top values on commercial white LED light, while, in 60–70 CRI emissions, the number of photons within the red color range are low.

Review Results. Identification of the Spectral Values of the Light Sources Used

The three different groups of lamps found are analyzed separately: fluorescent light, LED white light, LED monochromatic light and a combination of them.
  • Fluorescent light: A major conclusion that can be obtained is that the characterization of LED lamps is carried out in greater depth than that of fluorescent lamps that are used as a control element in many of the experiments studied. From 62 papers that use fluorescent lamps as the control lighting system, 30 of them (48.4%) do not give any specific description or details about the type of lamp used; another 15 (24.2%) only give a generic description of the group of the color that is emitted (cold, neutral or warm) but without any quantitative characterization value; and only the last 17 descriptions of works (27.4%) either present the graph of the radiation spectrum of the lamps (three papers) or quantitatively determine the nature of the light source used (CTT for white light or the description of the specific model of lamp with its identification of the spectrum through the manufacturer dataset and, thus, of the radiant spectrum that is emitted on the plants) (14 papers).
  • In this assessment we highlight an increasing trend in the presentation of details of the characterization of lighting equipment in the works that have been published more recently compared to older ones, due to the use of spectrometers in in vitro plant tissue culture laboratories, among other reasons.
  • White LED light: The main deficiency detected in the characterization of these type of LED lamps refers again to the sources of white light, since the weight of each of the wavelengths in this continuous spectrum emission is not defined in a clear way. Describing only extreme wavelength values does not allow us to establish whether the main emission weight is in blue light (440–450 nm) or in the yellow-green zone. In this case, a minimum description would require an indication of the color range of the white light (cold, neutral or warm). On the other hand, the most precise description is to provide the full spectral graph of the emitted light obtained with a spectrophotometer. An intermediate solution, which offers sufficient values for comparison purposes due to the standardization that is being found in these products, is to offer the CCT and CRI values of the light source. From 57 studies analyzed that claim to use LED white light, four offer the graph of their radiation diagram; nine indicate the CCT value (with values from 2700 K to 7000 K, the most common numbers being 3000 K, 4200 K and 6500 K); eight identify only the range of CCT with one of this three options: cold, neutral and warm (the first and the last values being the most identified) and, finally, 36 indicate simply that white LED light has been used without any extended specification.
  • Monochromatic LED light: In these cases, each light source can be clearly identified by the wavelength of its emission. It is relevant at this point to note how the generalized denomination of a color does not correspond to a single spectrum but to a set of frequencies that offers slight visual differences. In total, 52 studies out of 92 clearly identify the peak wavelength values, while the remaining 40 were limited to identifying only a generic group definition (red, blue, green, etc.). As described above, it is noteworthy how each plant species reaches a point of maximum sensitivity at different frequencies, so a precise study would require using not only an emission value in each basic group of colors, but also in making a sweep in several discrete frequencies. In the cases in which this more specific discrimination is done, there are 19 studies that deal with differences in the red color, including the study of the effect of far red or infrared lights (>700 nm) and seven studies include, at least, two different frequencies inside the range of the blue color (425–475 nm). Five papers introduce emissions with UV values (<400 nm) and another three use the violet/purple region of frequencies (400–425 nm).

3.2. Amount of Radiated Energy Incident on Plants

This is a major technical value presented in most of the descriptions available in the literature but is treated in different ways.
Most common technical variables related to lighting equipment and installations are defined based on the behavior of the human eye. The luminous flux—measured in lumens—is the total amount of energy emitted by a lamp, and a technical value given by their manufacturers to identify the products. However, the lumen is a unit of measurement that expresses how bright a certain light appears to the human eye and favors the yellow/green/orange spectrum because of the human eye’s spectral sensitivity curve. The counterpart option is the Photosynthetic Photon Flux (PPF). That unit measures the quantity of photons that a light fixture emits per second within the segment of the radiation spectrum that affects the photobiological processes of the plants: the Photosynthetically Active Radiation (PAR) region. This is defined between 400 nm and 700 nm. The unit of the PPF is the µmol s−1. However, neither of these two variables are adequate to characterize an experiment, as they give the total energy emission, and it is not possible to accurately assess how much of that radiation is directed towards the plants and how much is lost in the surrounding environment.
Thus, in order to assess an experiment correctly, it is necessary to define the measurement of the amount of energy incident in the plants such as the irradiance, or the radiant flux received by a surface per unit area. This effect, again, can be represented using different units. The International System (IS) unit is the W⋅m−2 but its usefulness is more suitable for energy balance studies, as it considers the full spectrum of radiations rather than identifying the adequacy and quantification of the number of photons within the PAR region. Moreover, irradiance meters are both costly and limited in the ability to measure low irradiance values. With a lower cost and higher sensitivity in low light conditions, basic light meters measure the luminous flux per unit area (illuminance) utilizing the units of lumens per meter squared or lux. As before, this does not accurately reflect the number of photons that reach the plants with different light sources because it is adapted to the perception of incident light by the human eye (International Electrotechnical Commission EV Ref: 845-21-060), i.e., it does not quantify the radiations in the range of the green-yellow spectrum in the same way as in the blue or red segments.
Considering the above, the most appropriate measurement variable, when working with plants, is the Photosynthetic Photon Flux Density (PPFD), with µmol m−2 s−1 as its unit. This is the amount of power of the electromagnetic radiation received on a surface measured as the number of photons in the PAR region of the spectrum received in a square meter per second [118]. This variable can be measured with affordable PAR sensors which are manufactured to offer a response in which each photon is equally absorbed considering that photons of shorter wavelength (higher frequency) have more energy than photons of longer wavelength.
For standardization reasons and to facilitate processes, the PPFD values should be measured on the flat surface where the culture vessels are placed.

Review Results. Characterization of the Amount of Light Irradiated

With regard to the irradiance of light, this variable is intended to be quantified in most of the descriptions reviewed. Overall, 85 of the works analyzed (87.6%) made a description of quantitative experiments against another 12 papers (12.4%) that did not specify this information in any form. The most widely used variable is the PPDF (µmol m−2 s−1) that is an option presented in 78 works, compared to two works that indicated the characteristic radiation of the light source (W m−2) and another six of them characterized at least one experiment using illuminance (lux) values.
The values of irradiance used in the different experiments are variable. Standard values in this field are 45–50 µmol m−2 s−1, and a significant number of the studies analyzed use these values. However, others use different quantities that are found mostly within the range among 25–85 µmol m−2 s−1. Dispersions are found in those works that either: (A) include analyzing the consequences of using different radiation intensities, (B) use different values justified in previous studies that indicate that these other quantities optimize a specific biological process on a species of plant or, finally, (C) present different values of µmol m−2 s−1 for each different light source in their experiments.
Regarding this last case, it may be significant to indicate that the discrete power values of the fluorescent tubes and their unique photometric emission diagram facilitates obtaining the previous standard values based on previous experiences, choosing the height and number of tubes depending on the surface of the cabins. However, the large heterogeneity of LED lamps makes it difficult to foresee the value to be obtained correctly and dispersions from these expected values can be easily obtained if a previous analysis bases on simulations or similar is not done. Including a dimming system allows the precise configuration of the installation as desired but if that option is not available, as in basic LED lamps, the only method available to adjust the intensity sent to the plants requires the modification of the geometry of the cabins.

3.3. Uniformity of the Lighting Conditions Achieved on the Shelves in the Culture Room

One of the characteristics that makes fluorescent tubes, both cool/warm white or Grolux light, a very suitable type of lamp for supplying light in in vitro culture rooms is the high uniformity of the light that they emit and project over the plants. The continuous, homogeneous and omnidirectional nature of the light emitted by fluorescents (See Figure 2A) makes obtaining high values of uniformity much easier, even without a specific study of this parameter. To reproduce lighting conditions similar to those provided by fluorescent tubes using LED technology, it will therefore be necessary to achieve ranges of the same level of uniformity of radiation conditions on the shelves of the culture room.
Contrary to what happens with fluorescent lamps that emit continuously and in all directions of space, LEDs are discrete light sources, integrated into non-continuous matrixes that are distributed over the entire surface of their lamps. In these matrixes, emitters of different spectra are usually combined. Moreover, the light that they emit is directional and has several possible configurations depending on the type of LED used and if any secondary optic component is integrated in the lamp (see Figure 2B–D). This has the consequence that the light reflected from the walls of the chambers has a much lower relevance than with fluorescent tubes, since the light is projected in a more concentrated way towards the work area. This is an obvious advantage from the point of view of energy efficiency, but it requires that the geometric configuration of the specific lamps being used and their position in space must be analyzed in order to guarantee that the same amount of light (both in intensity and in wavelength of radiation) reaches the entire vessel-laying surface. It is necessary to measure and guarantee a high surface homogeneity of the radiation emitted in its two mentioned variables to prevent the projections generated in the in vitro cultures placed in different locations within the working zones being significantly different.
Overall, Figure 2 compares the omnidirectional emission of the fluorescent tubes (Figure 2A) (radiations are not present only in the direction of its electrical connection sockets as can be observed in the transverse plane represented by the blue curve) compared to the natural directional nature of the LEDs (Figure 2B). The latter allow a simple and highly efficient adjustment of their emission curves by means of secondary lenses, as in Figure 2C,D.
Fluorescent lamps emit a lot of light outside the reach of plants, and it is necessary to play with the reflection of the working cell (walls and ceilings) or with specific reflectors to redirect the radiations. This reflected light system is not very energy efficient, but it allows homogeneously illuminated environments to be obtained in a natural way. However, the directionality of LEDs greatly optimizes energy efficiency but requires a configuration study of each case to ensure that there is uniform incident radiation throughout the work surface for all emitted wavelengths.
The two standardized systems to measure the uniformity of light emission are the average uniformity and the extreme uniformity. The first is defined as the quotient between the minimum value of the incident radiation, measured on our useful working area divided by the average value obtained from all the discrete value calculations used to characterize this surface (see Equation (1)). On the other hand, extreme uniformity is the quotient between the minimum value and the maximum measured value. These homogeneity values should be published for each of the dominant emission frequencies used in the experimentation spaces. To calculate these uniformity values of average and minimum radiance, a regular 2D matrix of test points must be marked over the complete working plane, with a step distance smaller than the diameter of two sample vessels (i.e., 15–20 cm) and equal in both Cartesians directions.
Umean = min ( PPFD i , j ) i = 1 : n ; j = 1 : m i , j n , m PPFD i , j
With ‘n’ being the number of discrete test points in the ‘X’ axe and ‘m’ the number of discrete test points in the ‘Y’ axe.
The radiance of each test point configured should be collected using a PAR meter for every type of emitter available. Only one value is needed for monochromatic of single-type continuous spectrum LEDs, but for LED lamps with several different emitters where each spectral emission can be controlled independently, this operation must be repeated for every case where a single color can be turned on independently. As an alternative, for homogeneity measure purposes, a lux meter can also be used as sensor without loss of accuracy because the relative condition of the variable only requires a homogeneous system of measure.
Once data are collected, each minimum irradiance is identified as the numerator of the uniformity ratio and the denominator is the average radiance that is obtained adding all the values and dividing the result by the number of points measured. Our recommendation is to consider that good values of average uniformity are ratios above 0.70 and they become optimal above 0.80.

3.3.1. Review Results. Characterization of the Uniformity of Light Radiated over the Plants

None of the works analyzed in the review presented any quantitative measure or consideration of homogeneity of either intensity or wavelength of radiance values, rather they offered statements of intent of subjective visual appreciations.

3.3.2. Experimental Verification

Figure 3 presents the comparison between the Grolux fluorescent tube spectra and the LED luminaire designed by the authors to offer an equivalent functionality. This equivalence is based on the proportional reproduction of the radiation of three monochromatic wavelengths (Blue (440–450 nm), Green (540–530 nm) and Red (660 nm)). The number of emitters and their geometric placement was established to homogenize accurately the lighting values achieved throughout the working area with the plants.
As reference, a model cabin measuring 130 cm (length) × 59 cm (width) × 44 cm (height) has been designed. In this working cell, six linear LED lamps, 115 mm long, have been attached to the upper wall. Each tube is built using only one type of light source in the color spectrum and has a mechanical structure that is formed by an extruded and anodized aluminum heatsink (to improve the heat evacuation capacity by natural convection) with a diameter of 13 mm. Each of these lamps contains 24 HP monochromatic LEDs from the manufacturer CREE with a separation distance of 4.5 mm without secondary optics. There is therefore a hemispherical projection of 120° (See Figure 4). Each tube is placed in the cell 8 cm apart from each other and centered with respect to the cabin and combined alternately with a pattern: R, G, B, R, G and B.
An SQ-120 quantum sensor by APOGEE was used to measure the uniformity that is achieved throughout the useful surface of the cell. This measurement equipment was chosen based on its low thickness deployable sensor that allows the least possible interference in the measurements by any shadow. A measuring mesh matrix of 12 × 5 (60) equidistant points was set starting from four fixed positionings located on the corners of the surface located 2 cm apart from the two edge lines of the cabin.
The mean uniformity of each of the three wavelengths that make up the equivalence of Grolux tubes were independently measured. In all cases, similar average uniformities were obtained around 0.72. The minimum readings were always given at one of the four points associated with a corner. In the case of excluding the measurements of these points, the average rose to 0.79.
With the lighting conditions of spectral composition and homogeneity previously described, and using strawberry and rose as model plants, practically identical results were obtained during successive subcultures in the proliferation of these two species when using Grolux and RGB LED tubes with the equivalent configuration [119]. See Figure 5.

3.4. Temperature

Fujiwara and Kozai [120], reviewing the physical microenvironment and its effects, pointed out that the temperature of the air within the culture vessels can be considered almost the same as that outside the vessels during almost all the dark period and that, during the photoperiod with cool white fluorescent lamps, the air temperature differences between the inside and outside were less than 0.5 °C with an irradiance of 60 µmol m−2 s−1. With this data as reference, no special attention needs to be paid to the air temperature in culture vessels in conventional plant tissue culture. However, these temperature differences increase with the raise of the irradiance values, as Tani et al. [121] pointed out, being maximum at the surface of the culture medium.
Incandescent and discharge lamps have electrodes and gases (in the second case) that are activated at high temperatures for their operation, which means that they emit a significant amount of radiant heat emission (infrared) in the same direction in which they project light. In addition, the electrical losses in the auxiliary work systems and in the connections of these lamps generate an additional amount of heat that is transmitted in the air through non-forced convection, directly affecting the culture vessels. In contrast, in LED lighting systems there are no components that reach high temperatures, eliminating the direct transmission of heat by radiation.

3.4.1. Review Results. Effect of Light Source, Discharge vs. LED, on the Temperature in Plant Tissue Culture

Considering the works covered in the review process, there are no data presented in the bibliography about the effect that their light sources have on temperature in plant tissue culture. Thus, it is not possible to compare the differences produced on the temperature at which cultures are found using fluorescent or LED lighting.

3.4.2. Experimental Verification

Measuring the temperatures inside and outside the culture vessels in the experimental conditions previously described, it is possible to observe that under Grolux lighting vessels reached a temperature of 33.2 °C and condensation frequently appeared in the culture flasks, especially in those items located above or below the line of the fluorescent tube (Figure 6a). On the contrary, under LED lighting no condensation appeared in any jar and their temperatures did not exceed the 28.8 °C (Figure 6b).
The elevated temperature found on the shelves, generated by the heat transmitted conductively though the material of the racks, made the temperature inside the vessels rise more than the outside ambient values. This produces an increase in the relative humidity in the plant ambient and a decrease in the inter-gaseous change, a phenomenon that Pasqualetto [122] called the ‘greenhouse effect’, which plays an important role in the appearance of hyperhydric plants.
Observing thermographic images, it can be seen that the temperature ranges inside the vessels oscillate between 27.5–35.0 °C while using Grolux lighting (Figure 7a) and between 27.0–28.0 °C (Figure 7b) when LED lighting is used. Therefore, depending on the area of the container, there would be differences of up to 7.5 °C with Grolux compared to only 1 °C with LEDs. Therefore, although temperature is not a parameter intrinsically linked to the light source, it will have a differential influence on the temperature reached inside and outside the culture vessels and on the temperature gradient created inside these containers.

4. Future Trends: Pulsed Light

In addition to these previous parameters of which information is found to a greater or lesser extent in the research works analyzed, it would be significant to detail also the constant or pulsating light flux emission identification. Whereas fluorescents always emit with a certain fluctuation due to the dependence of the sinusoidal signal of the power supply network (50/60 Hz, depending on the country), LED lamps work on constant current condition with no effect (theoretically) of oscillations. The drivers or power supply systems of LEDs that convert the electric power from AC to DC can filter the sinusoidal fluctuations of the standard power grid to achieve light emissions with an almost zero level of oscillation, at best (flicker < 1%), or higher (>10%) in lower quality products. However, the electrical nature of LED and its fast-switching capability allows the generation of variable active width control signals with supplementary regulation devices that can be used to generate different pulsed emission conditions. Pulse Width Modulation (PWM) regulators that create variable square signals commuting the input of electric power of the light sources are a common technology used to dim, in a controlled way, the intensity of light emission of LED lamps.
Various past [123] and recent works [124] have stated that this can be a variable that significantly influences different plant processes such as biological absorption. Thus, this can be considered a variable of interest in in vitro plant tissue culture, although there are still very few studies in this field, such as [125], which leaves open a line of investigation. In this case, it is important to characterize the shape (square, sinusoidal, etc.), amplitude, width active percentage and frequency of the light pulses configured to facilitate comparison of different experiments. These values are introduced as required inputs of a lighting installation control system to obtain one specific output setting which can be measured and verified using a spectrometer with flicker measurement capability.

5. Concluding Remarks

A large number of works are being published both in a general sense and, more particularly, in the area of in vitro plant tissue culture, in which LED lighting is used and its effect on growth and morphogenesis is evaluated. The nature of technology is significantly different compared to that the fluorescence, which has been most commonly used in the implementation of artificial lighting in in vitro culture rooms. The LED flexibility offers significant advantages (configuration, control capability, etc.) but also presents the problem of establishing how to guarantee that the results of published works are extrapolated, reproducible and comparable, not only for LEDs but also for fluorescence.
This review has focused on the need to standardize lighting conditions and their description. To achieve this objective, and having considered the state-of-the-art technologies involved, the characteristics and technical variables that must be detailed have been established so as to propose a methodological standardization. In this sense, it can be considered that there are three control parameters necessary to achieve this objective: incident radiant intensity in the crops, detailed light emission spectrum and uniformity of incidence of these radiations on the plants.
The description and analysis of LED or mixed in vitro lighting installation presented in 97 scientific publications since 2018 to the present day were reviewed and they were assessed with regard to the proposed variables and methodology. It is noteworthy that in general there is still a great need to improve the amount of detail in these descriptions; however, that said, the depth and precision of the descriptions found in many recent works compared to the oldest works analyzed show a positive trend in this sense.
Together with this bibliographic study, an experimental contribution has been made in which two lighting installations (Grolux fluorescence and LED) have been designed and configured to offer equivalent lighting performance considering the generated irradiance obtained in the shelves, emission spectrum—built based on the three main peaks of wavelength generated—and uniformity of both previous parameters. This experimental setup has made it possible to verify that the results obtained in the in vitro proliferation of strawberry and rose, as model plants, are practically identical with LED lamps configured and installed to match the emission of the Grolux fluorescents.
Finally, another noteworthy result of the study is how, although temperature is not a parameter intrinsically linked to the light sources, the lamps chosen—LED or fluorescence—will have a differential influence on the temperature reached inside and outside the culture vessels and on the gradients of temperature created. The increment on the environmental and microenvironmental temperatures generated is smaller when using LED technology.

Author Contributions

Conceptualization, A.B.-M.; investigation, M.B.-M., A.B.-M. and A.G.-C.; methodology, A.G.-C. and A.B.-M.; formal analysis A.G.-C. and M.B.-M.; writing—original draft A.G.-C., A.B.-M. and M.B.-M.; writing—review & editing A.B.-M., M.B.-M. and A.G.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the RTA2014–00056–C02–01 (INIA-AEI) and the AVA2019.008 projects (20% Junta de Andalucía, 80% FEDER).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors would like to thank Alfonso Gago Bohórquez for helping us to initiate this collaboration and to Marta Carrera Amalia-Cabeza and Josefa Alcalá for their technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

Glossary

ASA: ascorbic acidDHA: dehydroascorbic acid
B: blueDR: deep reed
BA: 6-benzyladenineDW: dry weight
BCF: bioconcentration factorFLs: fluorescent lamps
C: controlfR: far red
CCC: chlorocholine chlorideFW: fresh weight
CHO: carbohydrateG: Green
CHL: chlorophyllHP-LED: high-power LED
CNOC: carotenoidHPLC: high-performance liquid chromatography
CsLHY: geneIAA: indole-3-acetic acid, an auxin
Cv: cultivarIR: infrared
CW: cool WhiteLEDs: light emitting diodes
DEGs: differentially expressed genes MDA: malondialdehyde
mTR: meta-topolin ribosideSE: somatic embryos
MW: mint whiteSOD: superoxide dismutase
NW: neutral whiteSpd: spermidine
P: purpleTSIIA: tanshinone IIA
PA: polyamineV: violet
PAR: photosynthetically active radiationVOCs: volatile organic compounds
Pas: free polyaminesW: white
PGRs: plant growth regulatorsWW: warm white
PLBs: protocorm-like bodiesY: yellow
POD: peroxidases2iP: N6 -(2-isopentenyl) adenine
PPFD: photosynthetic photon fux densityR: red
Put: Putrescine

References

  1. Lecler, M.T.; Zimmermann, F.; Silvente, E.; Masson, A.; Morèle, Y.; Remy, A.; Chollot, A. Improving the work environment in the fluorescent lamp recycling sector by optimizing mercury elimination. Waste Manag. 2018, 76, 250–260. [Google Scholar] [CrossRef]
  2. Thomson, G. The Museum Environment, 2nd ed.; Elsevier Science: Burlington, VT, USA, 2013; ISBN 13-978-0750620413. [Google Scholar] [CrossRef]
  3. Ahmed, S.; Amir, F.; Riaz-ul-Hasnain, S.; Shahwar, D.; Jamil, S. Electronic ballast circuit configurations for fluorescent lamps. In Proceedings of the 2015 Power Generation System and Renewable Energy Technologies (PGSRET), Islamabad, Pakistan, 10–11 June 2015; pp. 1–8. [Google Scholar] [CrossRef]
  4. Hammer, E.E. Effects of ambient temperature on the performance of bent tube fluorescent lamps. IEEE Trans. Ind. Appl. 1989, 25, 216–223. [Google Scholar] [CrossRef]
  5. Helson, V.A. Comparison of Gro-Lux and cool-white fluorescent lamps with and without incandescent as light sources used in plant growth rooms for growth and development of tomato plants. Can. J. Plant Sci. 1965, 45, 461–466. [Google Scholar] [CrossRef]
  6. Teunissen, C.; Van der Heijden, F.H.F.W.; Poort, S.H.M.; De Beer, E. Characterising user preference for white LED light sources with CIE colour rendering index combined with a relative gamut area index. Light Res. Technol. 2017, 49, 461–480. [Google Scholar] [CrossRef] [Green Version]
  7. Lammerts, W.E. Comparative effect of Grolux and incandescent light on growth of camellias. In American Camellia Yearbook; Pyron, J.H., Ed.; American Camellia Society: Ft. Valley, GA, USA, 1964; pp. 158–162. [Google Scholar]
  8. Batista, D.; Felipe, S.H.S.; Silva, T.D.; de Castro, K.M.; Mamedes-Rodrigues, T.C.; Miranda, N.A.; Ríos-Ríos, A.M.; Faria, D.V.; Fortini, E.A.; Chagas, K.; et al. Light quality in plant tissue culture: Does it matter. In Vitro Cell Dev. Plant 2018, 54, 195–215. [Google Scholar] [CrossRef]
  9. Nair, G.B.; Dhoble, S.J. A perspective perception on the applications of light-emitting diodes. Luminescence 2015, 30, 1167–1175. [Google Scholar] [CrossRef] [PubMed]
  10. Golovatskaya, I.F.; Karnachuk, R.A. Role of green light in physiological activity of plants. Russ. J. Plant Physiol. 2015, 62, 727–740. [Google Scholar] [CrossRef]
  11. Luo, X.; Hu, R.; Liu, S.; Wang, K. Heat and fluid flow in high-power LED packaging and applications. Prog. Energy Combust. Sci. 2016, 56, 1–32. [Google Scholar] [CrossRef] [Green Version]
  12. Arik, M.; Becker, C.A.; Weaver, S.E.; Petroski, J. Thermal management of LEDs: Package to system. In Proceedings of the Third International Conference on Solid State Lighting, San Diego, CA, USA, 5–7 August 2003; International Society for Optics and Photonics: Bellingham, WA, USA, 2004; Volume 5187, pp. 64–75. [Google Scholar]
  13. Fan, H.B.; Li, X.P.; Shen, J.X.; Chen, M. An effective prediction method for LED lumen maintenance. In Proceedings of the 2012 13th International Conference on Electronic Packaging Technology & High Density Packaging, Guilin, China, 13–16 August 2012; pp. 1560–1563. [Google Scholar] [CrossRef]
  14. Hegedüs, J.; Hantos, G.; Poppe, A. Reliability issues of mid-power LEDs. In Proceedings of the 2019 25th International Workshop on Thermal Investigations of ICs and Systems (THERMINIC), Lecco, Italy, 25–27 September 2019; pp. 1–6. [Google Scholar] [CrossRef]
  15. Yadlapalli, R.T.; Narasipuram, R.P.; Kotapati, A. An overview of energy efficient solid state LED driver topologies. Int. J. Energy Res. 2020, 44, 612–630. [Google Scholar] [CrossRef]
  16. Padmasali, A.N.; Kini, S.G. A generalized methodology for predicting the lifetime performance of LED luminaire. IEEE Trans. Electron. Devices 2020, 67, 2831–2836. [Google Scholar] [CrossRef]
  17. Miyashita, Y.; Kimura, Y.; Kubota, C.; Kozai, T. Effects of red light on the growth and morphology of potato plantlets in vitro: Using emitting diodes (LEDs) as a light source for micropropagation. Acta Hortic. 1997, 418, 169–173. [Google Scholar] [CrossRef]
  18. Gupta, S.D.; Jatothu, B. Fundamentals and applications of light-emitting diodes (LEDs) in in vitro plant growth and morphogenesis. Plant. Biotechnol. Rep. 2013, 7, 211–220. [Google Scholar] [CrossRef]
  19. Agarwal, A.; Gupta, S.D. Impact of light-emitting diodes (LEDs) and its potential on plant growth and development in controlled-environment plant production system. Curr. Biotechnol. 2016, 5, 28–43. [Google Scholar] [CrossRef]
  20. Jung, W.S.; Chung, I.M.; Hwang, M.H.; Kim, S.H.; Yu, C.Y.; Ghimire, B.K. Application of light-emitting diodes for improving the nutritional quality and bioactive compound levels of some crops and medicinal plants. Molecules 2021, 26, 1477. [Google Scholar] [CrossRef] [PubMed]
  21. Hashim, M.; Ahmad, B.; Drouet, S.; Hano, C.; Abbasi, B.H.; Anjum, S. Comparative effects of different light sources on the production of key secondary metabolites in plants in vitro cultures. Plants 2021, 10, 1521. [Google Scholar] [CrossRef]
  22. Araújo, D.X.; Rocha, T.T.; de Carvalho, A.A.; Bertolucci, S.K.V.; Medeiros, A.P.R.; Ribeiro, F.N.S.; Barbosa, S.M.; Pinto, J.E.B.P. Photon flux density and wavelength influence on growth, photosynthetic pigments and volatile organic compound accumulation in Aeollanthus suaveolens (Catinga-de-mulata) under in vitro conditions. Ind. Crops Prod. 2021, 168, 113597. [Google Scholar] [CrossRef]
  23. Attaya, A.S. LED Light Technology as a Source of Illumination and a Promising Method for Stevia rebaudiana Elite Propagation. Egypt J. Agron. 2021, 43, 123–132. [Google Scholar] [CrossRef]
  24. Boonsnongcheep, P.; Pichetpongtorn, P.; Juntosuth, P.; Yowsong, T.; Vansanit, P.; Preeprame, S.; Kanchanapoom, T.; Putalun, W. LED and methyl jasmonate enhance a strong antioxidant compound, verbascoside, production in Clerodendrum indicum and Acanthus ebracteatus. J. Plant Biochem. Biot. 2021, 1–9. [Google Scholar] [CrossRef]
  25. Coelho, A.D.; de Souza, C.K.; Bertolucci, S.K.V.; de Carvalho, A.A.; Santos, G.C.; de Oliveira, T.; Marques, E.A.; Salimena, J.P.; Pinto, J.E.B.P. Wavelength and light intensity enhance growth, phytochemical contents and antioxidant activity in micropropagated plantlets of Urtica dioica L. Plant Cell Tissue Organ. Cult. 2021, 145, 59–74. [Google Scholar] [CrossRef]
  26. Costa, É.L.G.; Farnese, F.D.S.; de Oliveira, T.C.; Rosa, M.; Rodrigues, A.A.; Resende, E.C.; Januario, A.H.; Silva, F.G. Combinations of blue and red LEDs increase the morphophysiological performance and furanocoumarin production of Brosimum gaudichaudii Trécul in vitro. Front. Plant Sci. 2021, 1440. [Google Scholar] [CrossRef] [PubMed]
  27. Gnasekaran, P.; Rahman, Z.A.; Chew, B.L.; Appalasamy, S.; Mariappan, V.; Subramaniam, S. Development of micropropagation system of Zingiber officinale var. rubrum Theilade using different spectrum light-emitting diode (LED) irradiation. Ind. Crops Prod. 2021, 170, 113748. [Google Scholar] [CrossRef]
  28. Gris, T.; Pinheiro, M.V.M.; Thiesen, L.A.; Webler, A.R.; Junges, D.L.; Holz, E.; Naibo, I.; Batista, D.S.; Otoni, W.C.; Schmidt, D. Light quality and sealing type affect in vitro growth and development of Capsicum frutescens cultivars. An. Acad. Bras. Ciências 2021, 93. [Google Scholar] [CrossRef]
  29. Matos, A.V.C.D.S.D.; Oliveira, B.S.D.; Oliveira, M.E.B.S.D.; Cardoso, J.C. AgNO3 improved micropropagation and stimulate in vitro flowering of rose (Rosa x hybrida) cv. Sena. Ornam. Hortic. 2021, 27, 33–40. [Google Scholar] [CrossRef]
  30. Manivannan, A.; Soundararajan, P.; Park, Y.G.; Jeong, B.R. Physiological and proteomic insights into red and blue light-mediated enhancement of in vitro growth in Scrophularia kakudensis—A potential medicinal plant. Front. Plant Sci. 2021, 11, 2205. [Google Scholar] [CrossRef]
  31. Siatkowska, K.; Chraniuk, M.; Bollin, P.; Banasiuk, R. Light emitting diodes optimisation for secondary metabolites production by Droseraceae plants. J. Photochem. Photobiol. B Biol. 2021, 224, 112308. [Google Scholar] [CrossRef]
  32. Sonthisut, M.; Wongpanya, R.; Phonphoem, A.; Phonphoem, W.P. Enhancement of 1-deoxynojirimycin production in mulberry (Morus spp.) using LED irradiation. Plant Cell Tissue Organ. Cult. 2021, 1–10. [Google Scholar] [CrossRef]
  33. Sorgato, J.C.; Mudolon, E.D.; Guimarães, F.F.; Soares, J.S.; Ribeiro, L.M. Light sources on the germination and initial in vitro establishment of Schomburgkia crispa Lindl., a species of the Brazilian Cerrado. Ciência Rural. 2021, 51. [Google Scholar] [CrossRef]
  34. Weremczuk-Jeżyna, I.; Hnatuszko-Konka, K.; Lebelt, L.; Grzegorczyk-Karolak, I. The protective function and modification of secondary metabolite accumulation in response to light stress in Dracocephalum forrestii shoots. Int. J. Mol. Sci. 2021, 22, 7965. [Google Scholar] [CrossRef]
  35. Winson, K.W.S.; Chew, B.L.; Sathasivam, K.; Subramaniam, S. Effect of amino acid supplementation, elicitation and LEDs on Hylocereus costaricensis callus culture for the enhancement of betalain pigments. Sci. Hortic. 2021, 289, 110459. [Google Scholar] [CrossRef]
  36. Yu, R.; Li, F.; Wang, G.; Ruan, J.; Wu, L.; Wu, M.; Yang, C.; Shan, Q. In vitro regeneration of the colorful fern Pteris aspericaulis var. tricolor via green globular bodies system. In Vitro Cell Dev. Biol. Plant 2021, 57, 225–234. [Google Scholar] [CrossRef]
  37. Aasim, M.; Aydın, S.; Karataş, M.; Aydın, M.E.; Soğukpınar, C.; Sevinc, C. Bioaccumulation potential of in vitro regenerated plants of Ceratophyllum demersum against Chromium—A lab study. Asian J. Agric. Biol. 2020, 8, 233–239. [Google Scholar] [CrossRef]
  38. Cappai, F.; Garcia, A.; Cullen, R.; Davis, M.; Munoz, P.R. Advancements in low-chill blueberry Vaccinium corymbosum L. tissue culture practices. Plants 2020, 9, 1624. [Google Scholar] [CrossRef]
  39. Centofante, A.R. Light quality on the morphoanatomy and physiology of Campomanesia pubescens (DC.) O. Berg. Seedlings. Sci. Hortic. 2020, 108765. [Google Scholar] [CrossRef]
  40. Chaves, M.C.; Freitas, J.C.E.; Nery, F.C.; Paiva, R.; de Oliveira Prudente, D.; Costa, B.G.P.; Daubermann, A.G.; Bernardes, M.M.; Grazul, R.M. Influence of colorful light-emitting diodes on growth, biochemistry, and production of volatile organic compounds in vitro of Lippia filifolia (Verbenaceae). J. Photochem. Photobiol. B 2020, 212, 112040. [Google Scholar] [CrossRef]
  41. Chen, Y.C.; Chang, C.; Lin, H.L. Topolins and red light improve the micropropagation efficiency of passion fruit (Passiflora edulis Sims) ‘Tainung No. 1’. Hortscience 2020, 55, 1337–1344. [Google Scholar] [CrossRef]
  42. Chen, L.L.; Zhang, K.; Gong, X.C.; Wang, H.Y.; Gao, Y.H.; Wang, X.Q.; Zeng, Z.H.; Hu, Y.G. Effects of different LEDs light spectrum on the growth, leaf anatomy and chloroplast ultrastructure of potato plantlets in vitro and minituber production after transplanting in the greenhouse. J. Integr. Agric. 2020, 19, 108–119. [Google Scholar] [CrossRef]
  43. Chen, C.C.; Lee, M.R.; Wu, C.R.; Ke, H.J.; Xie, H.M.; Tsay, H.S.; Agrawal, D.C.; Chang, H.C. LED lights affecting morphogenesis and isosteroidal alkaloid contents in Fritillaria cirrhosa D. Don-An Important Chinese Medicinal Herb. Plants 2020, 9, 1351. [Google Scholar] [CrossRef]
  44. Da Costa Silveira, A.A.; Lopes, F.J.F.; Sibov, S.T. Micropropagation of Bambusa oldhamii Munro in heterotrophic, mixotrophic and photomixotrophic systems. Plant Cell Tissue Organ. Cult. 2020, 141, 315–326. [Google Scholar] [CrossRef]
  45. De Carvalho, A.A.; Bertolucci, S.K.V.; da Cunha Honorato, A.; Rocha, T.T.; Silva, S.T.; Pinto, J.E.B.P. Influence of light spectra and elicitors on growth and ascaridole content using in vitro cultures of Dysphania ambrosioides L. Plant Cell Tissue Org. 2020, 143, 277–290. [Google Scholar] [CrossRef]
  46. Dogan, M. The effectiveness of light emitting diodes on shoot regeneration in vitro from shoot tip tissues of Limnophila aromatica (Lamk.) Merr. and Rotala rotundifolia (Buch-Ham. ex Roxb) Koehne. Biotech. Histochem. 2020, 95, 225–232. [Google Scholar] [CrossRef]
  47. Gam, D.T.; Khoi, P.H.; Ngoc, P.B.; Linh, L.K.; Hung, N.K.; Anh, P.T.L.; Thu, N.T.; Hien, N.T.T.; Khanh, T.D.; Ha, C.H.H. LED lights promote growth and flavonoid accumulation of Anoectochilus roxburghii and are linked to the enhanced expression of several related genes. Plants 2020, 9, 1344. [Google Scholar] [CrossRef]
  48. He, C.; Zeng, Y.; Fu, Y.; Wu, J.; Liang, Q. Light quality affects the proliferation of in vitro cultured plantlets of Camellia oleifera Huajin. Peer J. 2020, 8, e10016. [Google Scholar] [CrossRef]
  49. Johnson, R.E.; Kong, Y.; Zheng, Y. Elongation growth mediated by blue light varies with light intensities and plant species: A comparison with red light in arugula and mustard seedlings. Environ. Exp. Bot. 2020, 169, 103898. [Google Scholar] [CrossRef]
  50. Li, R.N.; You, J.; Miao, C.; Kong, L.; Long, J.H.; Yan, Y.Z.; Xu, Z.G.; Liu, X.Y. Monochromatic lights regulate the formation, growth, and dormancy of in vitro grown Solanum tuberosum L. microtubers. Sci. Hortic. 2020, 261, 108947. [Google Scholar] [CrossRef]
  51. Manoh, H.P.; Toruan-Mathius, Y.N. Growth and development of oil palm shoots under different light qualities. Biotropia J. 2020, 28, 21–28. [Google Scholar] [CrossRef]
  52. Mølmann, J.A.; Johansen, T.J. Sprout growth inhibition and photomorphogenic development of potato seed tubers (Solanum tuberosum L.) under different LED light colours. Potato Res. 2020, 63, 199–215. [Google Scholar] [CrossRef]
  53. Neto, A.R.; Chagas, E.A.; Costa, B.N.S.; Chagas, P.C.; Vendrame, W.A. Photomixotrophic growth response of sugarcane in vitro plantlets using different light intensities and culture vessel types. In Vitro Cell Dev. Biol.-Plant 2020, 56, 504–514. [Google Scholar] [CrossRef]
  54. Oliveira, T.d.R.; Aragão, V.P.M.; Moharana, K.C.; Fedosejevs, E.; do Amaral, F.P.; Sousa, K.R.; Santa-Catarina, C. Light spectra affect the in vitro shoot development of Cedrela fissilis Vell. (Meliaceae) by changing the protein profile and polyamine contents. Biochim. Biophys. Acta Proteins Proteom. 2020, 1868, 140529. [Google Scholar] [CrossRef]
  55. Piątczak, E.; Kuźma, Ł.; Kozłowska, W.; Lisiecki, P.; Szemraj, M.; Płachno, B.J.; Zielińska, S. Phenylethanoid and iridoid glycosides production in Rehmannia elata NE Brown ex Prein. in vitro shoot cultures and their biological activity. Ind. Crops Prod. 2020, 158, 113050. [Google Scholar] [CrossRef]
  56. Pundir, R.K.; Pathak, A.; Joshi, N.; Bagri, D.S.; Upadhyaya, C.P. Irradiation studies of LED light spectra on the growth and development of potato (Solanum tuberosum L.). Plant Sci. Today 2020, 7, 406–416. [Google Scholar] [CrossRef]
  57. Santos, G.C.; Cardoso, F.P.; Martins, A.D.; Pascual, M.; Ossani, P.C.; Queiroz, J.M.; Rezende, R.A.L.S.; Dória, J. Effect of light and sucrose on photoautotrophic and photomixotrophic micropropagation of Physalis angulata. Biosci. J. 2020, 36, 1353–1367. [Google Scholar] [CrossRef]
  58. Siedlarz, P.; Bany, S.; Rybka, K. Quantitative changes in DNA methylation induced by monochromatic light in barley regenerants obtained by androgénesis. Biuletyn. IHAR 2020, 288, 47–51. [Google Scholar] [CrossRef]
  59. Silva, T.D.; Batista, D.S.; Fortini, E.A.; de Castro, K.M.; Sousa Felipe, S.H.; Fernandes, A.M.; de Jesus Sousa, R.M.; Chagas, K.; Siqueira da Silva, J.V.; de Freitas Correia, L.; et al. Blue and red light affects morphogenesis and 20-hydroxyecdisone N content of in vitro Pfaffia glomerata accessions. J. Photochem. Photobiol. B 2020, 203, 111761. [Google Scholar] [CrossRef]
  60. da Silva, D.; Imakawa, A.M.; Guedes, K.R.V.; Bruno, F.M.S.; Sampaio, P.D.T.B. In vitro regeneration and callogenesis of Libidibia férrea. J. Exp. Agric. Int. 2020, 42, 14–24. [Google Scholar] [CrossRef]
  61. Souza, D.M.S.C.; Xavier, A.; Miranda, N.A.; Gallo, R.; Otoni, W.C. Light quality, 6- benzyladenine and number of subcultives for in vitro multiplication of hybrid clones of Corymbia. Sci. For. 2020, 48, e3282. [Google Scholar] [CrossRef]
  62. Stepanova, A.Y.; Solov’yova, A.I.; Salamaikina, S.A. Influence of spectral light composition on flavones formation in callus culture of Scutellaria baicalensis Georgi. Pharmacogn. Mag. 2020, 16, 156–160. [Google Scholar] [CrossRef]
  63. Usman, H.; Ullah, M.A.; Jan, H.; Siddiquah, A.; Drouet, S.; Anjum, S.; Giglioli-Guviarc’h, N.; Hano, C.; Abbasi, B.H. Interactive effects of wide-spectrum monochromatic lights on phytochemical production, antioxidant and biological activities of Solanum xanthocarpum callus cultures. Molecules 2020, 25, 2201. [Google Scholar] [CrossRef] [PubMed]
  64. Yeow, L.C.; Chew, B.L.; Sreeramanan, S. Elevation of secondary metabolites production through light-emitting diodes (LEDs) illumination in protocorm-like bodies (PLBs) of Dendrobium hybrid orchid rich in phytochemicals with therapeutic effects. Biotechnol. Rep. 2020, 27, e00497. [Google Scholar] [CrossRef]
  65. Yu, L.L.; Song, C.M.; Sun, L.J.; Li, L.L.; Xu, Z.G.; Tang, C.M. Effects of light-emitting diodes on tissue culture plantlets and seedlings of rice (Oryza sativa L.). J. Integr. Agric. 2020, 19, 1743–1754. [Google Scholar] [CrossRef]
  66. Xu, Y.; Yang, M.; Cheng, F.; Liu, S.; Liang, Y. Effects of LED photoperiods and light qualities on in vitro growth and chlorophyll fluorescence of Cunninghamia lanceolata. Plant Biol. 2020, 20, 269. [Google Scholar] [CrossRef]
  67. Zárate-Salazar, J.R.; de Souza, L.M.; de Morais, M.B.; Neto, L.P.; Willadino, L.; Gouveia-Neto, A.; Ulisses, C. Light-emitting diodes and gas exchange facilitation minimize hyperhydricity in Lippia grata: Physiological, biochemical and morpho anatomical aspects. S. Afr. J. Bot. 2020, 135, 164–171. [Google Scholar] [CrossRef]
  68. Zhao, J.; Thi, L.T.; Park, Y.G.; Jeong, B.R. Light quality affects growth and physiology of Carpesium triste Maxim. cultured in vitro. Agriculture 2020, 10, 258. [Google Scholar] [CrossRef]
  69. Zielińska, S.; Piątczak, E.; Kozłowska, W.; Bohater, A.; Jezierska-Domaradzka, A.; Kolniak-Ostek, J.; Matkowski, A. LED illumination and plant growth regulators’ effects on growth and phenolic acids accumulation in Moluccella laevis L. in vitro cultures. Acta Physiol. Plant 2020, 42, 72. [Google Scholar] [CrossRef] [Green Version]
  70. Zushi, K.; Suehara, C.; Shirai, M. Effect of light intensity and wavelengths on ascorbic acid content and the antioxidant system in tomato fruit grown in vitro. Sci. Hortic. 2020, 274, 109673. [Google Scholar] [CrossRef]
  71. Bhaya, M.H.M.; Al-RazzaqSalim, S. Impacts of plant growth regulators and light quality on banana (Musa spp.) micropropagation. Plant Arch. 2019, 19, 1379–1385. [Google Scholar]
  72. Boonsnongcheep, P.; Sae-foo, W.; Banpakoat, K.; Channarong, S.; Chitsaithan, S.; Uafua, P.; Puthaa, W.; Kerdsiria, K.; Putalun, W. Artificial color light sources and precursor feeding enhance plumbagin production of the carnivorous plants Drosera burmannii and Drosera indica. J. Photochem. Photobiol. B 2019, 199, 111628. [Google Scholar] [CrossRef]
  73. Cióc, M.; Kalisz, A.; Zupnik, M.; Pawłowska, B. Different LED light intensities and 6-benzyladenine concentrations in relation to shoot development, leaf architecture, and photosynthetic pigments of Gerbera jamesonii Bolus in vitro. Agronomy 2019, 9, 358. [Google Scholar] [CrossRef] [Green Version]
  74. Chen, L.; Yang, Y.; Jiang, Y.; Zhao, J.; Zang, H.; Wang, X.; Hu, Y.; Xue, X. RNA-Seq analysis reveals differential responses of potato (Solanum tuberosum L.) plantlets cultured in vitro to red, blue, green, and white light-emitting diodes (LEDs). J. Plant. Growth Regul. 2019, 38, 1412–1427. [Google Scholar] [CrossRef]
  75. Das Bhowmik, S.S.; Cheng, A.Y.; Long, H.; Tan, G.Z.H.; Hoang, T.M.L.; Karbaschi, M.R.; Williams, B.; Higgins, T.J.V.; Mundree, S.G. Robust genetic transformation system to obtain non-chimeric transgenic chickpea. Front. Plant Sci. 2019, 10, 524. [Google Scholar] [CrossRef] [Green Version]
  76. De Hsie, B.S.; Bueno, A.I.S.; Bertolucci, S.K.V.; de Carvalho, A.A.; da Cunha, S.H.B.; Martins, E.R.; Pinto, J.E.B.P. Study of the influence of wavelengths and intensities of LEDs on the growth, photosynthetic pigment, and volatile compounds production of Lippia rotundifolia Cham in vitro. J. Photochem. Photobiol. B 2019, 198, 111577. [Google Scholar] [CrossRef] [PubMed]
  77. De Oliveira Prudente, D.; de Souza, L.B.; Paiva, R.; Domiciano, D.; de Carvalho, P.A.; Nery, F.C. Goji berry (Lycium barbarum L.) in vitro multiplication improved by light-emitting diodes (LEDs) and 6-benzylaminopurine. In Vitro Cell Dev. Biol. Plant 2019, 55, 258–264. [Google Scholar] [CrossRef]
  78. Dogan, M. The effects of different agar concentrations and light emitting diodes on in vitro shoot regeneration of Hygrophila polysperma (Roxb.) T. Anderson. Fresenius Environ. Bull. 2019, 28, 5382–5389. Available online: https://hdl.handle.net/11492/2524 (accessed on 17 November 2021).
  79. Faria, D.V.; de Freitas Correia, L.N.; de Souza, M.V.C.; Ríos Ríos, A.M.; Vital, C.E.; Batista, D.S.; Cardoso Costa, M.G.; Otoni, W.C. Irradiance and light quality affect two annatto (Bixa orellana L.) cultivars with contrasting bixin production. J. Photochem. Photobiol. B 2019, 197, 111549. [Google Scholar] [CrossRef] [PubMed]
  80. Gupta, S.D.; Kumar, A.; Agarwal, A. Impact of light-emitting diodes (LEDs) on the growth and morphogenesis of encapsulated shoot buds of Curculigo orchioides Gaertn., an endangered medicinal herb. Acta Physiol. Plant 2019, 41, 50. [Google Scholar] [CrossRef]
  81. Hu, J.; Leng, Y.; Jiang, Y.; Ni, S.; Zhang, L. Efect of light quality on regeneration and naphthoquinones accumulation of Arnebia euchroma. Plant Biotechnol. Rep. 2019, 13, 353–360. [Google Scholar] [CrossRef]
  82. Jiang, L.; Wang, Z.; Jin, G.; Lu, D.; Li, X. Responses of Favorita potato plantlets cultured in vitro under fluorescent and light-emitting diode (LED) light sources. Am. J. Potato Res. 2019, 96, 396–402. [Google Scholar] [CrossRef]
  83. Kepenek, K. Photosynthetic effects of light-emitting diode (LED) on in vitro-derived strawberry (Fragaria x Ananassa cv. Festival) plants under in vitro conditions. Erwerbs-Obstbau 2019, 61, 179–187. [Google Scholar] [CrossRef]
  84. Khan, T.; Ullah, M.A.; Garros, L.; Hano, C.; Abbasi, B.H. Synergistic effects of melatonin and distinct spectral lights for enhanced production of anti-cancerous compounds in callus cultures of Fagonia indica. J. Photochem. Photobiol. B 2019, 190, 163–171. [Google Scholar] [CrossRef]
  85. Lerin, J.; Aragão, V.P.M.; Reis, R.S.; Silveira, V.; Santa-Catarina, C. Proteomic profile and polyamine contents are modulated by light source to promote in vitro shoot development in Cariniana legalis (Martius) O. Kuntze (Lecythidaceae). Plant Cell Tissue Org. 2019, 137, 329–342. [Google Scholar] [CrossRef]
  86. Li, S.; Zhou, L.; Wu, S.; Liu, L.; Huang, M.; Lin, S.; Ding, G. Effects of LED light on Acacia melanoxylon bud proliferation in vitro and root growth ex vitro. Open Life Sci. 2019, 14, 349–357. [Google Scholar] [CrossRef] [PubMed]
  87. Lotfi, M.; Mars, M.; Werbrouck, S. Optimizing pear micropropagation and rooting with light emitting diodes and trans-cinnamic acid. Plant Growth Regul. 2019, 88, 173–180. [Google Scholar] [CrossRef]
  88. Maślanka, M.; Prokopiuk, B. Bulb organogenesis of Tulipa tarda in vitro cultures in relation to light environment. Acta Agric. Scand. B Soil Plant Sci. 2019, 69, 398–404. [Google Scholar] [CrossRef]
  89. Mehraj, H.; Alam, M.; Habiba, S.U.; Mehbub, H. LEDs combined with CHO sources and CCC priming PLB regeneration of Phalaenopsis. Horticulturae 2019, 5, 34. [Google Scholar] [CrossRef] [Green Version]
  90. Meng, X.; Wang, Z.; He, S.; Shi, L.; Song, Y.; Lou, X.; He, D. LED-supplied red and blue light alters the growth, antioxidant status, and photochemical potential of in vitro-grown Gerbera jamesonii plantlets. Hort. Sci. Technol. 2019, 37, 473–489. [Google Scholar] [CrossRef]
  91. Miler, N.; Kulus, D.; Woźny, A.; Rymarz, D.; Hajzer, M.; Wierzbowski, K.; Szeffs, L. Application of wide-spectrum light-emitting diodes in micropropagation of popular ornamental plant species: A study on plant quality and cost reduction. In Vitro Cell Dev. Biol. Plant 2019, 55, 99–108. [Google Scholar] [CrossRef] [Green Version]
  92. Nadeem, M.; Abbasi, B.H.; Younas, M.; Ahmad, W.; Zahir, A.; Hano, C. LED-enhanced biosynthesis of biologically active ingredients in callus cultures of Ocimum basilicum. J. Photochem. Photobiol. B 2019, 190, 172–178. [Google Scholar] [CrossRef]
  93. Nakonechnaya, O.V.; Gafitskaya, I.V.; Burkovskaya, E.V.; Khrolenko, Y.A.; Grishchenko, O.V.; Zhuravlev, Y.N.; Subbotin, E.P.O.; Kulchin, Y.N. Effect of light intensity on the morphogenesis of Stevia rebaudiana under in vitro conditions. Russ. J. Plant Physiol. 2019, 66, 656–663. [Google Scholar] [CrossRef]
  94. Pinheiro, M.V.M.; Schmidt, D.; Diel, M.I.; dos Santos, J.; Thiesen, L.A.; Vitalli de Azevedo, G.C.; Holz, E. In vitro propagation of alpinia cultivars in different light sources. Ornam. Hortic. 2019, 25, 49–54. [Google Scholar] [CrossRef]
  95. Ren, X.; Liu, Y.; Jeong, B.R. A two-stage culture method for zygotic embryos effectively overcomes constraints imposed by hypocotyl and epicotyl seed dormancy in Paeonia ostii ‘Fengdan’. Plants 2019, 8, 356. [Google Scholar] [CrossRef] [Green Version]
  96. Watcharatanona, K.; Ingkaninanc, K.; Putalun, W. Improved triterpenoid saponin glycosides accumulation in in vitro culture of Bacopa monnieri (L.) Wettst with precursor feeding and LED light exposure. Ind. Crops Prod. 2019, 134, 303–308. [Google Scholar] [CrossRef]
  97. Xu, Y.; Liang, Y.; Yang, M. Efects of composite LED light on root growth and antioxidant capacity of Cunninghamia lanceolata tissue culture seedlings. Sci. Rep. 2019, 9, 9766. [Google Scholar] [CrossRef] [PubMed]
  98. Zielińskaa, S.; Dryśb, A.; Piątczakc, E.; Kolniak-Ostekd, J.; Podgórskae, M.; Oszmiańskid, J.; Matkowski, A. Effect of LED illumination and amino acid supplementation on phenolic compounds profile in Agastache rugosa in vitro cultures. Phytochem. Lett. 2019, 31, 12–19. [Google Scholar] [CrossRef]
  99. Bantis, F.; Karamanoli, K.; Ainalidou, A.; Radoglou, K.; Constantinidou, H.I.A. Light emitting diodes (LEDs) affect morphological, physiological and phytochemical characteristics of pomegranate seedlings. Sci. Hortic. 2018, 234, 267–274. [Google Scholar] [CrossRef]
  100. Cioć, M.; Szewczyk, A.; Żupnik, M.; Kalisz, A.; Pawłowska, B. LED lighting affects plant growth, morphogenesis and phytochemical contents of Myrtus communis L. in vitro. Plant Cell Tissue Organ. Cult. 2018, 132, 433–447. [Google Scholar] [CrossRef] [Green Version]
  101. Chen, G.J.; Lee, M.S.; Lin, M.K.; Ko, C.Y.; Chang, W.T. Blue light decreases tanshinone IIA content in Salvia miltiorrhiza hairy roots via genes regulation. J. Photochem. Photobiol. B 2018, 183, 164–171. [Google Scholar] [CrossRef]
  102. Chen, L.; Xue, X.; Yang, Y.; Chen, F.; Zhao, J.; Wang, X.; Khan, A.T.; Hu, Y. Effects of red and blue LEDs on in vitro growth and microtuberization of potato single-node cuttings. Front. Agric. Sci. Eng. 2018, 5, 197–205. [Google Scholar] [CrossRef] [Green Version]
  103. Jeong, B.R.; Sivanesan, I. Impact of light quality and sucrose on adventitious shoot regeneration and bioactive compound accumulation in Ajuga multiflora Bunge. Sci. Hortic. 2018, 236, 222–228. [Google Scholar] [CrossRef]
  104. Larraburu, E.E.; Correa, G.S.; Llorente, B.E. In vitro development of yellow lapacho (Bignoniaceae) using high power light emitting diode. Rev. Árvore 2018, 42. [Google Scholar] [CrossRef]
  105. Lazzarini, L.E.S.; Bertolucci, S.K.V.; Pacheco, F.V.; dos Santos, J.; Silva, S.T.; de Carvalho, A.A.; Pinto, J.E.B.P. Quality and intensity of light affect Lippia gracilis Schauer plant growth and volatile compounds in vitro. Plant Cell Tissue Organ. Cult. 2018, 135, 367–379. [Google Scholar] [CrossRef]
  106. Li, R.; Huang, W.; Wang, X.; Liu, X.; Xu, Z. Effects of yellow, green, and different blue spectra on growth of potato plantlets in vitro. HortScience 2018, 53, 541–546. [Google Scholar] [CrossRef]
  107. Mamedes-Rodrigues, T.C.; Batista, D.S.; Napoleão, T.A.; Cruz, A.C.F.; Fortini, E.A.; Nogueira, F.T.S.; Romanel, E.; Otoni, W.C. Lignin and cellulose synthesis and antioxidative defense mechanisms are affected by light quality in Brachypodium distachyon. Plant Cell Tissue Organ. Cult. 2018, 133, 1–14. [Google Scholar] [CrossRef]
  108. Muneer, S.; Park, Y.G.; Jeong, B.R. Red and blue light emitting diodes (LEDs) participate in mitigation of hyperhydricity in in vitro grown carnation genotypes (Dianthus caryophyllus). J. Plant Growth Regul. 2018, 37, 370–379. [Google Scholar] [CrossRef]
  109. Nguyen, Q.H.; Park, Y.G.; Jeong, B.R. Promoted growth and development of carnation plantlets in vitro by ventilation and combined red and blue light. Flower Res. J. 2018, 26, 166–178. Available online: http://db.koreascholar.com/article.aspx?code=365555%27 (accessed on 17 November 2021). [CrossRef]
  110. Nhut, D.T.; Tung, H.T.; Tanaka, M. Enhanced growth and development of Cymbidium and Phalaenopsis plantlets cultured in vitro under light-emitting diodes. In Orchid Propagation: From Laboratories to Greenhouses-Methods and Protocols; Lee, Y.I., Yeung, E.T., Eds.; Springer Protocols Handbooks; Humana Press: New York, NY, USA, 2018; pp. 209–223. [Google Scholar] [CrossRef]
  111. Pawłowska, B.; Żupnik, M.; Szewczyk Taranek, B.; Cioć, M. Impact of LED light sources on morphogenesis and levels of photosynthetic pigments in Gerbera jamesonii grown in vitro. Hortic. Environ. Biotechnol. 2018, 59, 115–123. [Google Scholar] [CrossRef]
  112. Prokopiuk, B.; Cioć, M.; Maślanka, M.; Pawłowska, B. Effects of light spectra and benzyl adenine on in vitro adventitious bulb and shoot formation of Lilium regale EH Wilson. Propag. Ornam. Plants 2018, 18, 12–18. Available online: http://www.journal-pop.org/References/Vol_18_1(12-18).pdf (accessed on 17 November 2021).
  113. Rodrigues, P.H.V.; Arruda, F.; Forti, V.A. Slow-grown in vitro conservation of Heliconia champneiana cv. Splash under different light spectra. Sci. Agric. 2018, 75, 163–166. [Google Scholar] [CrossRef] [Green Version]
  114. Sant’Ana, C.R.D.O.; Paiva, R.; Reis, M.V.D.; Silva, D.P.C.D.; Silva, L.C. In vitro propagation of Campomanesia rufa: An endangered fruit species. Ciência Agrotecnol. 2018, 42, 372–380. [Google Scholar] [CrossRef]
  115. Souza, D.M.S.C. Fontes de luz na Micropropagação de Clones Híbridos de Corymbia. Dissertação Apresentada à Univ Federal de Viçosa, Como Parte das Exigências do Programa de Pós-Graduação em Ciência Florestal, para Obtenção do Título de Magister Scientiae. Brasil. 2018. Available online: http://www.locus.ufv.br/handle/123456789/19917 (accessed on 17 November 2021).
  116. Tung, H.T.; Nam, N.B.; Huy, N.P.; Luan, V.Q.; Hien, V.T.; Phuong, T.T.B.; Le, D.T.; Loc, N.H.; Nhut, D.T. A system for large scale production of chrysanthemum using microponics with the supplement of silver nanoparticles under light-emitting diodes. Sci. Hortic. 2018, 232, 153–161. [Google Scholar] [CrossRef]
  117. Wei, X.; Zhang, X.; Mao, J.; Liang, G.; Tong, H. Different light qualities modify morphology and expression of CsLHY in in vitro plantlets of Camillia sinensis. Int. J. Agric. Biol. 2018, 20, 676–682. [Google Scholar] [CrossRef]
  118. Mottus, M.; Sulev, M.; Baret, F.; Reinart, A.; Lopez, R. Photosynthetically Active Radiation: Measurement and Modeling. In Encyclopedia of Sustainability Science and Technology; Meyers., R.A., Ed.; Springer: New York, NY, USA, 2012. [Google Scholar] [CrossRef]
  119. Gago-Calderon, A.; Barceló-Muñoz, M.; Barceló-Muñoz, A. Descripción y estandarización de la iluminación LED para su utilización en el cultivo in vitro de plantas. In Proceedings of the XIV Reunión de la SECIVTV, El Ejido, Almería, Spain, 8–10 September 2021; Available online: https://riuma.uma.es/xmlui/handle/10630/22829 (accessed on 17 November 2021).
  120. Fujiwara, K.; Kozai, T. Physical microenvironment and its effects. In Automation and Environmental Control in Plant Tissue Culture; Aitken-Christie, T., Kozai, M., Smith, L., Eds.; Springer: Dordrecht, The Netherlands; Kluwer Academic Publishers: Alphen aan den Rijn, The Netherlands, 1995; pp. 319–369. [Google Scholar] [CrossRef]
  121. Tani, A.; Kiyota, M.; Taira, T.; Aiga, I. The effects oflight intensity, flow rate of aeration and plantlets on changes in temperatures in the vessel for plantlet culture. Plant Tissue Cult. Lett. 1991, 9, 133–135. [Google Scholar] [CrossRef]
  122. Pasqualetto, P.L. Vitrification in plant tissue culture. In Plant Aging; Springer: Boston, MA, USA, 1990; pp. 133–137. [Google Scholar] [CrossRef]
  123. Bula, R.; Tennessen, D.J.; Morrow, R.C.; Tibbitts, T.W. Light emitting diodes as a plant lighting source. HortScience 1991, 26, 203–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Olvera-Gonzalez, E.; Escalante-Garcia, N.; Myers, D.; Ampim, P.; Obeng, E.; Alaniz-Lumbreras, D.; Castaño, V. Pulsed LED-Lighting as an alternative energy savings technique for vertical farms and plant factories. Energies 2021, 14, 1603. [Google Scholar] [CrossRef]
  125. Huang, H.; Tang, H.F.; Chen, Y.; Zheng, X.F.; Hu, Y.; Wang, R.H.; Huang, M.L. Effect of intense pulsed light on Trichophyton rubrum growth in vitro. J. South. Med. Univ. 2017, 37, 853–857. [Google Scholar] [CrossRef]
Figure 1. Spectral diagram and values of light characterization of the emission of a Grolux fluorescent tube.
Figure 1. Spectral diagram and values of light characterization of the emission of a Grolux fluorescent tube.
Plants 11 00060 g001
Figure 2. Common photometric diagrams found in led lamps for in vitro culture. (A) Fluorescent tube lamp (B) 120° hemispheric distribution (regular pattern of most HP LEDs without external lenses) (C) 60° concentrated distribution (regular pattern of many HB LED without external lenses). (D) Homogenized 120° distribution. Optimal lens design for chambers (i.e., PMMA C17720 EMERALD lens by Ledil).
Figure 2. Common photometric diagrams found in led lamps for in vitro culture. (A) Fluorescent tube lamp (B) 120° hemispheric distribution (regular pattern of most HP LEDs without external lenses) (C) 60° concentrated distribution (regular pattern of many HB LED without external lenses). (D) Homogenized 120° distribution. Optimal lens design for chambers (i.e., PMMA C17720 EMERALD lens by Ledil).
Plants 11 00060 g002
Figure 3. Comparison of the light spectra of (a) Grolux fluorescent and (b) R + G + B LED equivalence.
Figure 3. Comparison of the light spectra of (a) Grolux fluorescent and (b) R + G + B LED equivalence.
Plants 11 00060 g003
Figure 4. Test cabins with (a) Grolux fluorescent lamps and (b) LED Grolux spectrum equivalence.
Figure 4. Test cabins with (a) Grolux fluorescent lamps and (b) LED Grolux spectrum equivalence.
Plants 11 00060 g004
Figure 5. Effect of the light spectra on in vitro proliferation of (a) strawberry and (b) rose with Grolux fluorescent (blue line) and R + G + B LED equivalence (red line).
Figure 5. Effect of the light spectra on in vitro proliferation of (a) strawberry and (b) rose with Grolux fluorescent (blue line) and R + G + B LED equivalence (red line).
Plants 11 00060 g005
Figure 6. Thermographic images of the temperature outside the vessels under (a) Grolux fluorescent lighting and (b) LED lighting. (Thermographic camera model 875. Manufacturer Testo).
Figure 6. Thermographic images of the temperature outside the vessels under (a) Grolux fluorescent lighting and (b) LED lighting. (Thermographic camera model 875. Manufacturer Testo).
Plants 11 00060 g006
Figure 7. Thermographic images of the temperature in the vessels under (a) Grolux fluorescent lighting and (b) LED lighting. (Thermographic camera model 875. Manufacturer Testo).
Figure 7. Thermographic images of the temperature in the vessels under (a) Grolux fluorescent lighting and (b) LED lighting. (Thermographic camera model 875. Manufacturer Testo).
Plants 11 00060 g007
Table 1. Average distribution of the transformation of electrical energy in fluorescent tubes.
Table 1. Average distribution of the transformation of electrical energy in fluorescent tubes.
Energy ConversionVisible LightUltra Violet (UV)Infra-Red (IR)Heat
Fluorescent lamps23%1%30%46%
Table 2. Range of energy losses depending on the type of auxiliary ballast of fluorescent lamps [3].
Table 2. Range of energy losses depending on the type of auxiliary ballast of fluorescent lamps [3].
Conventional Magnetic BallastHigh Efficiency Magnetic BallastElectronic Ballast
20–25%14–16%8–11%
Table 3. Range of energy losses depending on the type of auxiliary ballast of fluorescent lamps.
Table 3. Range of energy losses depending on the type of auxiliary ballast of fluorescent lamps.
Technical ParameterQuantity
Mercury content of the lamp2.80 mg
Control gear requiredElectronic Ballast (+8–10% Power)
Average life (Nominal)14,000 h
Operating temperature range−15–40 °C (Maximum luminous flux @ 25 °C)
Light colorGrolux (Chromaticity Not adjustable)
Table 4. Experimental LED lighting conditions on in vitro plant tissue culture. (A): Lamps, spectrum. (B): Radiation, measurement equipment, laying distance.
Table 4. Experimental LED lighting conditions on in vitro plant tissue culture. (A): Lamps, spectrum. (B): Radiation, measurement equipment, laying distance.
SpeciesExperimental Lighting ConditionsParameters StudiedRemarksReferences
(A)(B)
Aeollanthus suaveolensLEDs: W, G, Y, B, R, R:B (1:2.5, 2.5:1; 1:1)
FL: CW (C)
Graph spectrum but not given definition values or charactersitics
PPFD LED (μmol m−2 s−1): W [20, 57, 78, 102 and 139]; Monocromatic LED: undefined
PPFD FL (μmol m−2 s−1): 42
SPECTRA PEN Z850 spectrometer; Qubit Systems
QSO-S Procheck + Sensor-PAR Photon Flux device; Decagon Devices-Pullman.
Shoot and root no. and length; leaf no.; shoot, root and leaf DW; leaf area; total DWGrowth was optimal with a PPFD of 139 μmol m−2 s−1, CNOC and CHL content, increased at 20 μmol m−2 s−1. W LEDs and FL were similar for DW and growth, but R, B and Y LED inhibited the last[22]
Stevia rebaudiana BertoniLED (nm): W (420) -undefined-, R (660), B (460)
IP65 model, SMD 5050 RBG (Techno Lite®).
FL: (4000–7000 nm) (C) -Typing error and undefined-
Illuminance (klux): CW FL 1, 2, 3 and 4
PPFD LED (μmol m−2 s−1): 40–50
Shoot induction %, no. and length; leaf no.; FW; DW; callus formation; CNOC and CHL contentHigh-quality plants can be grown in vitro under B, R and W LEDs for mass propagation and genetic improvement[23]
Clerodedum indicum and Acanthus ebracteatsLED (nm): W (400–700), B (425–500) and R (600–700) -undefined-Irradiation: 61.5 W·m2
30 cm between light sources and plants
Verbascoside levelMethyjasmonate and B LEDs enhanced verbascoside level[24]
Urtica dioica L.LED: B, R, W and B:R (1:1, 2.5:1, 1:2.5) -undefined-
FL: CW (C) -undefined-

PPFD LED (μmol m−2 s−1): 26, 51, 69, 94 and 130
PPFD FL (μmol m–2 s–1): 42
Portable spectrometer SPECTRA PEN Z850; Qubit Systems-Kingston.
QSO-S Pro Check +Par Photon Flux Sensor; Decagon Devices; Pullman
Plant and root length, DW, leaf area, photosynthetic pigments, phenolics, favonoids and antioxidant activityA PPFD of 94 μmol m−2 s−1 stimulated growth and DW.
The antioxidant activity was directly proportional to the increase in light intensity
[25]
Brosimum gaudichaudiiLED (nm): W (400–700), B (400–490), R (645–700) and R:B (1:1, 1:3, 3:1) -undefined-PPFD (μmol m−2 s−1): 100 ± 5
PAR sensor, model APG-SQ-316; Apogee
Shoot length, leaf no. and area, stem and leaf dry matter, total dry matter
R induced elongation. B and R:B enhanced the content of furanocoumarin, but B:R decreased the growth[26]
Zingiber officinale TheiladeLED (nm): W (400−700), B (460), G (530), R (660), fR (720) and R:B (1:1) (400 and 660)
LED Tubes, OSRAM Opto Semiconductors
PPFD (μmol m−2 s−1): W 15.3, fR 20.8, B 15.8, G 17.0, R 29.3, R:B 44.8
30 cm between light and plants
No. of Microsprouts, shoots, roots and leaves; shoot length; porphyrin, CHL and CNOC contentR LED improved rooting and shoot length, P stimulated micropropagation. G, R and P enhanced microshoot growth considerably, but fR and B, stimulated it too[27]
Rosa x hybrida cv. SenaCW FL cold bulbs 40 W & W LED bulbs 10W -undefined- Shoot regeneration %, flowering induction %, abnormal flowers %W LEDs (10W) reduced the % abnormal flowers in vitro, similar to W FLs (40 W)[28]
Capsicum frutescensFL (Luz do Dia Especial) (40 W, Osram) -undefined-
LEDs: W, R, B and R:B [6:4] (TEC LAMP®) -undefined-
PPFD (μmol m−2 s−1): 72Germination %; length; nodes no., no. of green and senescent leaves; FW, DW, CNOC, CHLPorous membranes, R and B:R improved growth and development[29]
Scrophularia kakudensisLED (nm): R (621 and 710), B (450 and 475) (4 lamps PSLED1203-50A, Force Lighting)
FL: CW 40 W, Philips -undefined-
PPFD (µmol m−2 s−1): 50
30 cm between light sources and plants
Growth, stomatal ultrastructure, phenols and flavonoids, activities of antioxidant enzymes, and protein expressionB or R LEDs improved the micropropagation. R and B elicited the synthesis of secondary metabolites[30]
DroseraceaeLED: CW (7000 K), NW (4000 K) and WW (3000 K)
FL: NW 3600 K Philips Master TL-D 58 W/835
Illuminance (lx): 5000 measured:center of the shelf.
PPFD (μmol m−2 s−1): FL 33.7 ± 11.4; LED 7000 K: 44.2 ± 15.3; LED 4000 K: 40.9 ± 13.7; LED 3000 K: 43.8 ± 14.8
GL Spectrolux spectrometer; GL Optic
Shoot length, growth, secondary metabolites productionLEDs results as a more efficient, eco-friendly and economically reasonable source of light for big scale in vitro production than FLs[31]
Morus sppLED (nm): R:G:B (635/520/452), R (635) and B (452).
FL: W (C) Graph spectrum but not given definition values or charactersitics
PPFD (μmol m−2 s−1): 10 and 50
Spectral Colour Illumino meter; Hangzhou Hopoo
Deoxynojirimycin (DNJ) contentR LED particularly increased the DNJ production[32]
Schomburgkia crispa Lindl.LEDs: WW (3000 K) and CW
(6500 K)
FL: CW 6500 K [C]
PPFD LED (μmol m–2 s–1): WW 128; CW 58; and CW 108
PPFD FL (μmol m–2 s–1): CW FL 23
Germination %3000 K LED induced a faster establishment and low mortality[33]
Dracocephalum forrestiiLED (nm): B (430), R (670), B:R (7:3) and NW (430–670) -Graph spectrum but not given definition values-
FL: CW [C] -undefined-
Spectrometer BTS256-LED Tester; Gigahertz-Optik
PPFD (μmol m–2 s–1): 40
Shoot proliferation and length; FW; DW; photosynthetic pigments; secondary compounds with therapeutic valueW LED induced the best proliferation. B was optimum for biomass and micropropagation. LEDs increaseed secondary metabolites accumulation[34]
Hylocereus costaricensisLEDs (nm): CW (400–700) [C], MW LEDs (540), R (660), B (440) and R:B (660 and 440).
Osram Opto Semiconductors
Darkness
FL: CW Phillips TLD, 36 W
Graphs of relative spectral distribution given
PPFD (μmol m–2 s–1):
CW 30, MW 17.5, R 37.9, B 18.5, R:B 61.1, CW FL 30
SPIC200 portable spectral irradiance colourimeter; Everfine Corporation
Betalain pigment, phenolic and flavonoid content, antioxidant activityR enhanced betalain content in Y and R callus. B and R + B light improved antioxidant properties. R enhanced phenolics, flavonoids and antioxidant activity in R callus[35]
Pteris aspericaulis var. tricolorLEDs (nm): W (450–470, 570–590), B (440) and R (660)PPFD (μmol m−2 s−1): 40GGB differentiation frequency and FW, plant height, leaf morphologyLEDs did not affected GGB multiplication and differentiation. R promoted elongation, but B inhibited it[36]
Ceratophyllum demersumLED: W -undefined-Iluminance (lux): 1500FW, DW, Cr uptake, BCF
In vitro plants are useful for Cr phytoremediation
[37]
Vaccinium corymbosum L.LED: R:B (77:23) and W
-undefined-
FL: W -undefined-
PPFD (μmol m−2 s−1): 55 ± 12No. of shoots, FW, DWR:B LEDs induced more production of meristems and biomass than W LED or FLs[38]
Campomanesia pubescens (DC.)LED (nm): R (600–700), B (400–490), W (400–700) and R:B (1:1)
-undefined-
PPFD (μmol m−2 s−1): 50 ± 5
PPFD sensor QSO-S;
Decagon Devices, Pullman
Biometry; leaf area and anatomy; MDA, DW; CHL, CNOCR increased MDA, with oxidative damages. W, B:R enhanced biometry[39]
Lippia filifoliaLED (nm): W (400–700) [C], B (450), R (653), R:B (664 and 448)Spectrorad. R Tide USB 650 UV, Ocean Optics™Height, shoots no., FW, CHL, CNOC, othersR:B enhanced the growth and the regeneration[40]
Passiflora edulis SimsLED (nm): Combination of B (450), G (525), R (660), and IR (730)PPFD (μmol m–2 s–1): 42Height, bud, root no., CHL contentR and mTR promoted a reliable propagation[41]
Solanum tuberosum L.LED: R (660), B (450), R:B (65:35) and R:B:G (520) (45:20:35)
FL: W (C) -undefined-
PPFD (μmol m–2 s–1): 100Stem, root length and Φ, health index, leaf area, FW, DW, starch, othersB, R:B, R:B:G LEDs were better than FLs for micropropagation[42]
Fritillaria cirrhosa D. DonLED (nm): R (660), B (450), fR (730), CW (5000 K), WW (2700 K)
R:B (8:1), R:G (525):B (7:1:1)
and R:B:fR (1:1:1)
PPFD (μmol m–2 s–1):
CW 57; WW 56; R:G:B 56; R:B 57;
B 57; R 56; fR 10; R:B:fR 56
No. of SE, FW, alkaloid contentR, fR increased alkaloid content. R, fR, R:B:fR enhanced FW and SE nº[43]
Bambusa oldhamii MunroLED (nm): B (455), R (630), R:B (30:70; 70:30) (TEC-LUX LED)
FL: W (C) -undefined-
PPFD FL (μmol m–2 s–1): 40Shoot height, no. of shoots and leavesB:R, TDZ and gas exchange increased shoot proliferation[44]
Dysphania ambrosioides L.LED: B, R, W and B:R [1:1, 2:1, 1:2] -undefined-PPFD (μmol m–2 s–1):
LED: 42; FL CW: 60
Shoot length; leaf, shoot and root dry biomass; VCOW, B:R [2:1] increased the growth. B reduced Z-ascaridole content[45]
Limnophila aromatica & Rotala rotundifoliaLED: W, R, B, W:R:B (2:1:1), W:R:B (1:2:1), W:R:B (1:1:2) and W:R:B [1:1:1] -undefined-
FL: W [C] -undefined-
Iluminance (lux): 1500
Luxometer PCE-EM 888
Regeneration %, no. and length of shootsW:R:B [1:2:1] was the most effective for in vitro propagation[46]
Anoectochilus roxburghiiLED (nm): R (630), B (465), B:R (20:80), B:R:W (13.8: 72.4: 13.8; 13.8: 58.6: 27.6)
FL: (C) -undefined-
PPFD (μmol m–2 s–1): 30 ± 2.13
Distance between plants and lights: 40–60 cm
Stem diameter, leaf no. and area, height, FW, DW, root no., stem anatomy, flavonoidsBR [1:4] enhanced the flavonoids content, the propagation and the medicinal value[47]
Camellia oleífera HuajinLED (nm): R (640), B (450), R:B (4:1) and R:B (1:4)
LED: W (C) -undefined-
PPFD (μmol m–2 s–1): 50 ± 5Proliferation, length, height, CHL, CNOC, leaf anatomy, proteinsR:B [4:1] induced the highest proliferation coefficient[48]
Brassica eruca, ‘Rocket’ and Brassica juncea, ‘Ruby Streaks’LED (nm): R (665), B (440)
(Pro-Series 325 by LumiGrow)
PPFD (μmol m–2 s–1): 20, 70, 120, 250, 450 and 650Dry mass, hypocotyl and petiole elongation, size, plant coloringB LED promoted elongation, but this varied with light intensity and plant species.[49]
Solanum tuberosum L.LED (nm): B (440), B (460), G (520), Y (590), R (620) and R (660)PPFD (μmol m–2 s–1): 65Height, stem diameter, leaf and microtuber no., CHOLight reduced the production cycle, and increased microtubers[50]
Elaeis guineensisLED (nm): R (660), B (460) and R:B [3:1] (660 and 460)
LED W: 10,000 K;
FL: W 6500 K, Phillips T5 28 W;
DARK
FW, leaf no., shoot height, rootsR:B induced the growth. R enhanced the rooting[51]
Solanum tuberosum L.LED (nm): Comb of UV-A (380) + UV-A (400) + B (450) + G (520) + R (660) + fR (735)
FL: WW 2700 K (+C)
DARK (-C)
PPFD LED max (μmol m–2 s–1): 1000
PPFD FL (μmol m–2 s–1): 2.8–4.6
Sprout length, growth vigourR at low irradiances, reduced elongation; fR at hight ones, reduced it according to the cv.[52]
Saccharum officinarum L.R:B LEDs (Philips Green Power R/B 150 43 W)
DR:W LEDs (Philips GP DR/W 150 33 W)
FL: W (OSRAM Sylvania)
PPDF (μmol m−2 s−1):
R:B 70, 120 and 200 ± 25;
DR:W 70 and 120; W FL 200
Shoot height, leaf no., yellow leaf %, shoot and root no., root length, FW, DW, CHLIncreases in light intensity stimulated plant height and leaf nº, without negative effects[53]
Cedrela fissilis VellLED: W:mB, W:mB:DR, W:mB:DR:fR -undefined-
LED (nm): B (425–490), DR (620–700) and fR (700–740)
FL [C] -undefined-
Length and no. of shoots, FW, DWBA, WmBdR: enhanced FW, DW and length. Proteins were identified[54]
Rehmannia elata N.E. Brown ex PreinW LED (8000–10,000 K)
RB:B:lR:R:DR:fR (5:10:10:35:35:5)
LED (nm): RB (430), B (460), lR (610), R (630), DR (660), fR (730)
PPFD (μmol m–2 s–1): 40Shoot %, axillary shoots no., leaf area and width, roots no. and length, CHLPAR illumination and PGRs enhanced the regeneration[55]
Solanum tuberosum L.LED (nm): W (C) -undefined-; R (650), B (460), R:B (3:7), R:B (1:1) and R:B (7:3)PPFD (μmol m–2 s–1):
W 100, R and B 210
Height, stem diameter, branches and leaves no., leaf area, FW, DW, health index, pigments, starch, soluble proteins, sugars and phenolics, ROS and ascorbate contentR:B (3:7) was optimal for plant development and growth[56]
Physalis angulataLuminaire: Screen filtered by Polysack’s Cromatinet® black photoconverter mesh with 50% shading with 6 lampsLED (nm): B (450), R (660), B:R, G (525), Y (590)
LED CW (7000 k)
O2, CO2, stem and root length, leaf area, nodal segments, CHL, CNOC, leaf anatomyFiltered natural light, allowed photoautotrophic propagation. LEDs did not promote it[57]
Hordeum vulgare L.LED (nm): B (454.63), G (525.95), R (630.84)
DARK in callus induction
Quantitative analysis of DNA methylationThe methylation depends on the light conditions[58]
Pfaffia glomerataLED (nm): R (665), B (440), R:B (1:1), R:B (1:3) and R:B (3:1)PPFD (μmol m−2 s−1): 80Stem and root length, leaf area, DW, CHL, CNOCR and B LEDs enhanced biomass and 20E production[59]
Libidibia ferreaLED: R:B and W -undefined-PPFD (μmol m–2 s–1): 31 ± 1
Shoots length and no., multiplication, buds no.
R:B stimulated the growth[60]
Corymbia. citriodora × C. torelliana and C. torelliana × C. citriodoraLED: CW -undefined-
LED (nm): R:B (450 and 660)
FL: -undefined-
Shoots length, no.; vigor, oxidation, CHL, CNOC contentR:B LEDs and sucrose enhanced elongation[1]
Hybrid Corymbia clonesLED: CW -undefined-
LED (nm): R:B (450 and 660)
FL: -undefined-
Length, no., vigor and oxidation of axillary shootsR:B, BA and ninth subculture, enhanced multiplication[61]
Scutellaria baicalensisLED (nm): B (420–480), R (600–650), WW (400–800)
DARK [C]
PPFD (µmol m−2 s−1): 1Growth of callus, flavones contentLight increased the callus. B induced flavones content[62]
Solanum xanthocarpumLED (nm): B (460), G (510), R (660) and Y (570)
LED W (400–700 nm) -undefined-
DARK
PPFD (µmol m−2 s−1): 45–50Flavonoids, phenolics, phytochemicals, antioxidant activityW increased the biomass. B, enhanced phytochemicals and phenolics content[63]
Dendrobium Enopi x Dendrobium Pink LadyLED CW (400–700 nm)
LED (nm): fR (730), R (660), G (530), B (440) and B:R

PPFD (µmol m−2 s−1):
CW 4.6, 5.2 and 17.0; fR 1.1, 9.1 and 20.8; R 1.3, 15.4 and 29.3; G 0.8, 6.2 and 16.9; B 0.9, 6.7 and 15.7; B:R 2.0, 20.3 and 44.8
Phenolics content, secondary metabolites accumulationB:R increased flavonoids.
R FL pre-illumination reduced LED effects on metabolites production
[64]
Oryza sativa L. cv. NipponbareFL W and
LED: B, B:R (3:1, 1:1; 1:3), R
PPFD (μmol m–2 s–1): 50Differentiation and budding rate, healthy index, plant no., length, CNOC, CHL, othersB improved callus regeneration. B:R [1:1] enhanced rice factory seedling cultivation[65]
Cunninghamia lanceolataLED: R:B (88.9:11.1), R:B:P (80:10:10), R:B:P:G (72.7:9.1:9.1:9.1), R:B:G (12.7:3.9:83.4) [C]PPFD (μmol·m−2 s−1): 20 & 30Plant height and no., rooting %, root no. and length, area, CHLR:B:P:G was the best for in vitro growth[66]
Lippia grata SchauerLED R:B [5:1;1:1] and R:G:B
LED (nm): B (460), R (640)
and G (530)
FL: W -undefined-
Non-ventilation under FL
PPFD (μmol m–2 s–1): 60
Sprouts no., roots, leaves, FW, hyperhydricity, height, sucrose, CHO and CNOC contentHyperhydricity was reduced under R:B and sealing; W decreased it in leaves[67]
Carpesium triste Maxim.LED: CW -undefined-,
LED (nm): R (621–710), B (450–475), R:B (1:1) (400–700 nm
PPFD (μmol m–2 s–1): 50
Shoot diameter and length, root length and no., FW, DWR and B LEDs produced high-quality in vitro plants
[68]
Moluccella laevis L.LED (nm): W 8000–10,000 K (400–700) and B (430):B (460):R (610):R (630):fR (730) (5:10:10:35:35:5)
(LED lamps Commled Solutions)
PPFD (μmol m–2 s–1): 180
Portable LightMeter HD 2302.0 equipped with LP 471 PAR and LP471 UVA detectors (DeltaOhm)
Shoots and buds proliferation, axillary shoots no. and length, callus diameter, CHLPGRs effect on shoot growth and development was stronger than the light infuence[69]
Two tomato cvs: House Momotaro and Mini CarolEight-peak LED [C] (nm)
R (625 and 660), B (420 and 450), G (520), fR (730), UV (390) and W (400–700 nm) -undefined-; DARK
PPFD (μmol m–2 s–1): 226–249ASA, DHA, antioxidant enzymes, H2O2, oxidative parametersHigh light intensity enhanced ASA content. Differences with B and R LEDs were observed[70]
Musa spp. CV. Dwarf CavendishLED R:B (18:2)
FL: W -undefined-
Iluminance (lux): 1000Shoots no., length, FW, DW, CHO, CHL, CNOCR:B LEDs improved in vitro propagation[71]
Drosera burmannii Vahl and D. indica L.LED (nm): W (400–700), B (425–500) and R (600–700) -undefined-
DARK
Irradiance (W·m2): 61.5
Distance lamps and plants: 30 cm
Plumbagin contentB LED enhanced plumbagin level, being higher in aerial parts[72]
Gerbera jamesoniiFL: W (C) -undefined-
LED (nm); R:B (7:3) (670 and 430)
PPFD LED (µmol m−2 s−1): 40, 80 and 120
PPFD WFL (µmol m−2 s−1): 40 [C]
Multiplication, leaf no., morphometry, axillary shoots, height, FW, DW, CHLR:B and BA improved growth. High radiation enhanced leaf features[73]
Solanum tuberosum L.LED (nm): B (450), G (530) and R (660)
LED: WW [C] –undefined-
PPFD (µmol m−2 s−1): 100
Distance lamps and shelf: 60 cm
Height, diameter, FW, DW, leaf area and no., health indexB increased health index. Light quality induced DEGs patterns[74]
ChickpeasR LED -undefined-
FL: (EN 12464-1 FLUORA)
PPFD (µmol m−2 s−1): 70Transgenic shoots, graftingLEDs and micro-injury improved transformation[75]
Lippia rotundifolia ChamLED: W, R, B, R:B (1:1; 2.5:1;1:2.5)
FL: CW -undefined-
PPFD FL CW (µmol m−2 s−1): 20, 54, 78, 88 and 110
PPFD (µmol m−2 s−1): 42
Shoot length, leaf no., DW (shoot, leaf, root and total), CHL, CNOCR:B and low intensity promoted high growth and pigments[76]
Lycium barbarum L. (Goji Berry)FL: W (400–700 nm) –undefined-
R:B LED (630 and 460 nm)
PPFD W FL (µmol m−2 s−1): 36
PPFD LED (μmol m−2 s−1): 86
Buds %, shoot length and no., leaf nºRB stimulated shoots length and the multiplication[77]
Hygrophila polyspermaLED: W, R, B, R:B, R:W, B:W, R:B:W -undefined- No. of shootsW:R:B incremented the shoots nº[78]
Bixa orellana L.FL: CW 4200 K
LED W -undefined-
B:R -undefined-
PPFD (µmol m−2 s−1): 50, 150 and 200Leaf stomatal density, no. and area, bixin, MDA, CNOC, CHLB:R and FL enhanced bixin and pigments depending on the cv.[79]
Curculigo orchioides GaertnLED (nm): B (470), R (630), B:R (1:1)
FL: CW (300–700 nm)
-undefined- [C].
PPFD (µmol m−2 s−1): 50Germination, shoots, roots, FW, DW, leaf area and density, CHL, CNOC, othersB and BR improved the synseed growth. R reduced growth and germination[80]
Arnebia euchromaFL CW (420 nm
LED (nm): UV (410–416)
B (450–455), R (650–660), R:B:W (25:25:50) and W//DARK
PPFD (µmol m−2 s−1): 60Germination rate, shoot no., FW, DW, naphthoquinones contentRBW, R enhanced growth. R increaded roots and dark the naphthoquinones content[81]
Solanum tuberosum L.LED (nm): R (660), B (450), R:B (80:20; 70:30;50:50)
FL: CW -undefined-
PPFD (µmol m−2 s−1): 40FW, height, CHL, soluble sugar contentR:B (70:30) was the best[82]
Fragaria x ananassa cv. FestivalLED (nm): fR (735):R (640):G (510):B (450):WW (3000 K) (15:55:15:10:5)
FL: CW (C) -undefined-
PPFD LED (µmol m−2 s−1): 25, 50, 75 and 100
PPFD FL (µmol m−2 s−1): 45
Survival rate, shoots no. and lenght, leaf area, height, root no., rooting %, FW, DWPPFD of 75μmol m–2 s–1, R:fR:B:G:WW was suitable[83]
Fagonia indicaLED (nm): W (380–780), B(380–560), G (480–670), Y (530–780), R (610–715) -undefined-
FL: W -undefined-; DARK
PPFD (µmol m−2 s−1): 40–50FW, DW, flavonoid and phenolic contents, callus, antioxidative enzyme activitiesW LED enhanced phenolics and flavonoids production. Under B LED, SOD and POD were best[84]
Cariniana legalis (Martius) O. KuntzeLED W: -undefined-
LED: low B and DR, W:low B, DR: fR, W: medium B: DR, W:medium B:DR FR -undefined-
FL: (C) -undefined-
PPFD (µmol m−2 s−1): 55Length and no. of shoots, shoot inductionB:R:fR light induced shoot elongation. W:B:R LED affected the endogenous contents of Pas[85]
Acacia melanoxylonW LED (400–700 nm)
R:B LED (1:1; 1:4; 1:4) -undefined-
FL: (C) -undefined-
PPFD (µmol m−2 s−1):
W LED: 45, 90 and 135
LED and FL: 135
Bud proliferation, growth, rooting rate, length and no.R:B promoted growth. High photoperiod and intensity enhanced growth and reduced the proliferation[86]
Pyrus communis L.LED (nm): B (454), R (660), fR (745), R:B (1:1), fR:B (1:1), R:fR (1:1)
FL: WW 3000 K (C)
PPFD (µmol m−2 s−1): 40No. and length of shoots, callus weight, leaf area, CHL, CNOCR promoted shoots length; fR stimulated the nº, but reduced the shoot quality. B enhanced the callus growth[87]
Tulipa tarda StapfFL: W (390–760 nm), R (647–770 nm), B (400–492 nm)
DARK [C]
Frequency of differentiation, no. of adventitious bulbs, FWDark enhanced the adventitious bulbs nº. Light spectra did not produces differences[88]
Phalaenopsis ‘Fmk02010’LED: R, G, B, W, R:G, R:B, R:W
G:B, G:W, B:W, R:G:B, R:B:W, R:G:W, G:B:W -all undefined-
FL: W (C) -undefined-
PPFD (µmol m−2 s−1): 54No. and FW of PLBs, shoots, and roots, length of shootsR, B or R:B used first and then W enhanced the regeneration and specifics CHO content[89]
Gerbera jamesoniiLED (nm): R (657), B (450), R:B (8:2; 7:3; 6:4; 5:5)
FL: W (C) -undefined-
PPDF (µmol m−2 s−1): 40Height; leaf length and no.; root length and no.; DW, CNOC, CHLR:B (7:3) improved growth and photosynthetic activity[90]
Chrysanthemum × grandiflorum, Gerbera jamesonii, Heuchera × hybrida, Ficus benjamina, and Lamprocapnos spectabilisLED: B:G:R:FR (14:16:53:17; 12:19:61:8; 8:2:65:25)
UV:B:G:R:FR (1:20:39:35:5)
-undefined-
FL: CW 6200 K (C)
PPFD FL (µmol m−2 s−1): 62–65Micropropagation efficiency, shoot length, leaf and root no., root length, FW, DW, CHLB:G:R:fR LED was the best for plant quality, micropropagation and cost reduction[91]
Ocimum basilicumLED (nm): R (660), B (460), G (510), Y (570), W (400–700)
FL W (400–700 nm) (C)-undef-
DARK
PPFD (µmol m−2 s−1): 40–50Phenolic and flavonoid contentLEDs light is a potent elicitor for in vitro metabolites production[92]
Stevia rebaudiana (Bertoni)Solar Box (SB) LED
LED: R:G:B, R:B
LED WW (3000 K)
FL: W 6500 K [C]
LED (nm): R (650), G (520), B (450)
PPDF FL (µmol m−2 s−1): 49
PPDF LED (µmol m−2 s−1): 75, 135, 230 and 382
Height, internode and leaf length, leaf width, FW and DW of roots and shootsLight intensity at 75 to 230 μmol m−2 s−1 improved the plant development[93]
Alpinia purpurataLED: -undefined-
FL -undefined-
PPFD (µmol m−2 s−1): 72Shoot and leaf no., shoot lengthThe light effect varied according to the cv.[94]
Paeonia ostii FengdanLED: R, B, R:B –undefined-
DARK (C)
PPFD (µmol m−2 s−1): 50Hypocotyl and epicotyl dormancy breaking %Dormancy was broken in dark, R or B depending on the case[95]
Bacopa monnieri (L.) WettstLED (nm): B (425–500), R (600–700)
LED: W (400–700) (C); DARK
Triterpenoid saponin glycosides contentB light was the most suitable for bioactive compound content[96]
Cunninghamia (C.) lanceolataR:B (4:1; 8:1), R:B:P (8:1:1), R:B:P:G (6:1:1:1; 8:1:1:1)
W LED (C) -undefined-
LEDs (nm): Red (620–630), B (460–470), P (410–420), G (520–530)
Illuminance (lux): 600–700 on the surface of the bottles
Rooting rate, root no., surface area and activityR:B:P:G enhanced root growth, and R:B:P:G (8:1:1:1) was the best[97]
Agastache rugosaLED W: WL 2700 K; WL 3000 K, NW 4000 K
LED (nm):RB (450), B (470), B (500), G (525), R (660), fR (720)
FL: CW 4000 K
PPFD FL and WLED (μmol m−2 s−1): 40Axillary bud breaking %, axillary shoots no., length of shootThe age of cultures, light and amino acids affected the phenolic compounds content[98]
Punica granatum L.LED (nm): UV (<400 nm), B (400–500), G (500–600), R (600–700), fR (700–800)PPFD (μmol m−2 s−1): 200 ± 20Leaf no., leaf area, root length, shoot height, FW and DW of shoots, leaves and rootsFLs reduced the roots. B, R, high G, enhanced the morphological features[99]
Myrtus comutis L.LED (nm):B (430), R (670), R:B (70:30)
FL: W 6200 K (C)
PPFD (μmol m−2 s−1): 35Multiplication, height, leaf no., FW, DW, CHL, CNOC, othersR stimulated growth and polyphenols. BA enhanced growth[100]
Salvia miltiorrhiza12 light treatments. Combination of LED (nm): R (660), G (525), B (450), fR (730) and UV (380) Tanshinone IIA (TSIIA) contentLEDs affected secondary metabolite production through gene regulation[101]
Solanum tuberosumLED (nm): R (660), B (440) and G (525)
LED WW -undefined-
PPFD (μmol m−2 s−1): 75Stem diameter, height, nodes no., leaf area, FW, DWR:B stimulated the micropropagation and microtuber production[102]
Ajuga multiflora BungeLED (nm): B (660), R (450)
FL: CW -undefined-
PPFD (μmol m−2 s−1): 45Shoots no. and length, tocopherols, CNOCWFL and 2% sucrose increased shoots number. B, R enhanced the micropropagation[103]
Handroanthus ochraceusLED: WW (peaks 475, 550 nm)
-undefined-
FL: W (peaks 400, 440, 490, 550, 615, 710 nm) -undefined-
PPFD (μmol m−2 s−1): 15, 20, 30, 40, 50 and 60Multiplication rate, hyper hydricity; roots %; shoots and roots FW, DW and length; leaf numberHigh-power LED irradiation increased the shoot growth
[104]
Lippia gracilisLED -undefined-: R, B, R:B (2.5:1; 1:2.5)
FL: CW –undefined-
PPFD (μmol m−2 s−1): 26, 51, 69, 94 and 130Shoot and root length, shoot and leaf no., DW, CHL, CNOCR and 94 µmol m−2 s−1 stimulated the growth. B, the photosynthetic pigments[105]
Solanum tuberosum L.LED (nm): R (630), B (445–465), Y (590), G (520), R:B:Y [6:2:1], R:B:G (6:2:1), R:Y:G (6:2:1), R:B:G (6:2:1)
FL: (400–700 nm) (C) -undef-
PPFD (μmol m−2 s−1): 72 ± 2Stem and root length, stem diameter, health index, leaf area, CHL, FW, DW, starch, othersR:B:Y LED increased the vigor in in vitro plants[106]
Brachypodium distachyonLED: CW -undefined-
LED: B:R (450 and 660 nm)
FL: CW -undefined-
PPFD FL (μmol m−2 s−1): 150
PPFD LED (μmol m−2 s−1): 56
Height, DW, shoot no., panicles, roots, histochemical analysesLight quality regulates cell wall deposition and lignification patterns[107]
Dianches caryophyllus L.LED: R and B -undefined-
FL: CW -undefined-
PPFD (μmol m−2 s−1): 50Adventitious shoot, proteins levelsFLs increased hyperhydricity R and B LEDs reduced it[108]
Dianthus caryophyllus L.W LED [C] -undefined-
LED: B, R, R:B -undefined-
Shoots no. and length, hyperhydricity, plant quality, CLHB or R LEDs and ventilation improved the quality of plants[109]
Phalaenopsis y CymbidiumLED: R, B, R:B (90:10), R:B LED (80:20), R:B LED (70:30) -undef-
FL: Growlux [C]
PPDF FL (μmol m−2 s−1): 45
PPDF LED (μmol m−2 s−1): 60–75
Leaf no, height, shoot and root FW, DW, root length and nº, CLHR and B enhanced the in vitro propagation. R produced weak plants with thin stems[110]
Gerbera jamesoniiLED (nm): R (430), B (670), R:B (50:50; 70:30), R:B:W (430–730) (40:40:20) and R:B:fR (730) (49:49:2) FL: CW 6200 K [C]PPFD (μmol m−2 s−1): 40Shoot no., plant length and heigh, rooting %, roots length and no., leaf no., DW, CLH, CNOCR:B (70:30) incremented the multiplication. R was optimal in rhizogenesis[111]
Lilium regaleLED (nm): R (670), B (430) and R:B (70:30).
FL: CW 6200 K (C)
DARK
PPFD (μmol m−2 s−1): 35Regeneration %, bulbs, shoots and roots no.B and B:R enhanced the organogenesis, dark and FL reduced it. R promoted roots, but reduced bulb growth[112]
Heliconia Champneiana cv. Splash.LED (nm): R (620–630), B (455–475) and R:B [70:30]
FL: W (380–780 nm) [C] -undef-
PPFD (μmol m−2 s−1): 25Height, FW, no. and length of roots, no. of leavesB reduced the growth, improved the quality and the survival in acclimatization[113]
Campomanesia rufaFL: W (20 W) -undefined-
LED: R:B (7:3) -undefined-
PPDF FL (μmol m−2 s−1): 44
PPDF LED (μmol m−2 s−1): 98
Shoots and buds no., shoots length, leaves no.BAP and FL W FLs improved shoots and buds growth[114]
Corymbia. torelliana x C. citriodora, C. citriodora x C. torellianaLED (nm): B (450) and R (660)
FL: Grolux
DARK
PPFD (μmol m−2 s−1): 80Shoots length and no., contamination, oxidationR, B LEDs produced the best results in in vitro propagation[115]
Chrysanthemum morifolium Ramat cv. JimbaLED (nm): G (565), B (450), R (660), Y (590), B:R (10:90; 20:80; 30:70; 40:60; 50:50; 60:40)
FL: (C) -undefined-
PPFD (μmol m−2 s−1): 40–45FW, DW, leaf size, no. and stomata density, photosynthesis rateMicroponic system, R:B (70:30) and AgNPs enhanced the development[116]
Camillia sinensisLED: B, R, FR, W -undefined-
FL: W (C) -undefined-
PPDF FL (μmol m−2 s−1): 30–40
PPDF LED (μmol m−2 s−1): 50
Leaf area and FW, stem length and diameter, CLH, CNOCR increased growth. B stimulated CsLHY expression and fR inhibited[117]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Barceló-Muñoz, A.; Barceló-Muñoz, M.; Gago-Calderon, A. Effect of LED Lighting on Physical Environment and Microenvironment on In Vitro Plant Growth and Morphogenesis: The Need to Standardize Lighting Conditions and Their Description. Plants 2022, 11, 60. https://doi.org/10.3390/plants11010060

AMA Style

Barceló-Muñoz A, Barceló-Muñoz M, Gago-Calderon A. Effect of LED Lighting on Physical Environment and Microenvironment on In Vitro Plant Growth and Morphogenesis: The Need to Standardize Lighting Conditions and Their Description. Plants. 2022; 11(1):60. https://doi.org/10.3390/plants11010060

Chicago/Turabian Style

Barceló-Muñoz, Araceli, Marta Barceló-Muñoz, and Alfonso Gago-Calderon. 2022. "Effect of LED Lighting on Physical Environment and Microenvironment on In Vitro Plant Growth and Morphogenesis: The Need to Standardize Lighting Conditions and Their Description" Plants 11, no. 1: 60. https://doi.org/10.3390/plants11010060

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