Next Article in Journal
A New Quinazolinone Alkaloid along with Known Compounds with Seed-Germination-Promoting Activity from Rhodiola tibetica Endophytic Fungus Penicillium sp. HJT-A-6
Previous Article in Journal
Metabolic Rewiring in Cancer: Small Molecule Inhibitors in Colorectal Cancer Therapy
Previous Article in Special Issue
UHPLC-HRMS/MS Chemical Fingerprinting of the Bioactive Partition from Cultivated Piper aduncum L.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Strategies, Achievements, and Potential Challenges of Plant and Microbial Chassis in the Biosynthesis of Plant Secondary Metabolites

by
Taotao Han
1 and
Guopeng Miao
1,2,*
1
Department of Bioengineering, Huainan Normal University, Huainan 232038, China
2
Key Laboratory of Bioresource and Environmental Biotechnology of Anhui Higher Education Institutes, Huainan Normal University, Huainan 232038, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(9), 2106; https://doi.org/10.3390/molecules29092106
Submission received: 8 March 2024 / Revised: 27 April 2024 / Accepted: 27 April 2024 / Published: 2 May 2024
(This article belongs to the Special Issue Biomanufacturing of Natural Bioactive Compounds)

Abstract

:
Diverse secondary metabolites in plants, with their rich biological activities, have long been important sources for human medicine, food additives, pesticides, etc. However, the large-scale cultivation of host plants consumes land resources and is susceptible to pest and disease problems. Additionally, the multi-step and demanding nature of chemical synthesis adds to production costs, limiting their widespread application. In vitro cultivation and the metabolic engineering of plants have significantly enhanced the synthesis of secondary metabolites with successful industrial production cases. As synthetic biology advances, more research is focusing on heterologous synthesis using microorganisms. This review provides a comprehensive comparison between these two chassis, evaluating their performance in the synthesis of various types of secondary metabolites from the perspectives of yield and strategies. It also discusses the challenges they face and offers insights into future efforts and directions.

Graphical Abstract

1. Introduction

The origin and early diversification of land plants represent a groundbreaking period in the history of plant life. To adapt to the transition from aquatic to terrestrial environments, plants underwent various physiological adaptations, leading to the emergence of numerous secondary metabolites and metabolic pathways controlling their biosynthesis [1]. Around 1850, pharmaceutical organic chemists began to notice the functions of these secondary metabolites and estimated their structures through extraction, isolation, and spectral identification. They also attempted to reconstruct them through chemical synthesis [2]. After years of scientific exploration, secondary metabolites have become essential sources for human medicine, pesticides, fragrances, dyes, and food additives [3].
Despite the vast diversity and quantity of bioactive substances in the plant kingdom, the majority of them are present in plants at very low concentrations. Large-scale cultivation requires extensive land use, faces challenges from pests and diseases, and entails high expenses. Chemical synthesis often involves numerous steps and harsh conditions, making it a less than optimal choice. In vitro cultivation and the metabolic engineering of plants have been used for many years to enhance the synthesis of plant secondary metabolites [4]. However, with the advancements in synthetic biology, an increasing focus has shifted toward heterologous synthesis using microorganisms [5].
This review investigates the strategies, achievements, and potential challenges in the synthesis of plant secondary metabolites using both plant-based and microbial-based chassis. In the case of plant-based chassis, the emphasis is placed on the in vitro cultivation and genetic modification of natural hosts. For model plant hosts, such as Nicotiana benthamiana, the review of Liu et al. [6] provides guidance on their engineering and utilization. In terms of microorganisms, we primarily focus on the key hosts of synthetic biology, E. coli and yeast. Other non-model microorganisms can be found in the review by Liu et al. [7].

2. Types and Biosynthesis of Plant Secondary Metabolites

Plant secondary metabolites are diverse in type and originate from various biosynthetic pathways. Based on their biosynthetic origins, plant natural products can be categorized into three main classes: terpenoids, alkaloids, and polyphenols [8].
Terpenoids are a group of compounds that include all isoprene polymers and their derivatives. They are based on the five-carbon isoprene unit and are also known as isoprenoids. Terpenoids are categorized into acyclic (linear) and cyclic terpenoids based on the way the isoprene units are connected [9]. These compounds are synthesized by the condensation of isopentenyl diphosphate (IPP) and dimethylallyl diphosphate (DMAPP) (Figure 1A) [10]. IPP is converted into DMAPP by the action of IPP isomerase, and both IPP and DMAPP are further converted into geranyl pyrophosphate (GPP) by the action of isoprenyl transferases. Farnesyl pyrophosphate (FPP) and geranylgeranyl pyrophosphate (GGPP) serve as non-cyclic intermediates and are further transformed into various classes of terpenoid compounds through the action of terpenoid synthases [11,12]. Terpenoids are the most diverse class of secondary metabolites in plants, with highly variable chemical structures. They play crucial roles in various aspects of plant growth and development, such as in the biosynthesis of plant hormones like gibberellins, abscisic acid, strigolactones, and cytokinins. Terpenoids also contribute to the adaptation to environmental stresses, as they include compounds like allelopathic substances, defensive compounds against herbivores, and chemicals involved in allelopathy [13]. Moreover, plant terpenoids have significant medicinal and economic value, including well-known drugs like artemisinin for malaria treatment and paclitaxel as an anti-cancer agent [10].
Alkaloids are the largest class of nitrogen-containing organic compounds among secondary metabolites. They mainly include terpenoid indole alkaloids, isoquinoline alkaloids, tropane alkaloids, and pyridine alkaloids [14]. Most alkaloids are synthesized from L-amino acids (such as tryptophan, tyrosine, phenylalanine, lysine, and arginine) individually or in combination with steroids, secoiridoid (e.g., strychnoside), or other terpene ligands (Figure 1B). Terpenoid indole alkaloids consist of an indole ring and a secologanin. Among the key enzymes in the metabolic pathway of this class is the enzyme strictosidine synthase, and its product, strictosidine, is an important branching point in this pathway. It can be further transformed into various related alkaloids like ajmalicine, quinine, and strychnine [15]. In the biosynthesis pathway of tetrahydrophenylisoquinoline alkaloids, the branching point is (S)-reticuline. Under the action of specific synthases, it can be further converted into alkaloids such as berberine, sanguinarine, and morphine [16]. Precursors for alkaloids like nicotine and scopolamine are derived from ornithine or arginine. Key enzymes involved in the biosynthesis of these compounds include ornithine decarboxylase, tropinone reductase, and hyoscyamine hydroxylase [17]. Since the isolation of the first alkaloid, morphine, from opium in 1816, a variety of alkaloids such as piperine, atropine, caffeine, and quinine have been sequentially isolated. Currently, there are over 27,000 reported alkaloids with diverse chemical structures. Among these, 88 of them have various pharmacological properties, including analgesic (morphine), stimulant (caffeine and ephedrine), psychoactive (mescaline and cocaine), antimicrobial (quinine), and anticancer (vinblastine and vincristine) activities [18].
The plant kingdom comprises more than 8000 to 10,000 phenolic compounds, which exhibit excellent antioxidant activity due to one or more aromatic rings with hydroxyl groups. Additionally, they serve as defense compounds in plants against bacteria, fungi, and insects [19]. Phenolic compounds can be further divided into four classes: flavonoids, phenolic acids that derived from hydroxybenzoic acids or hydroxycinnamic acids, stilbenes, and lignans [20]. Flavonoids originate from the phenylpropanoid metabolic pathway, with their basic structure composed of a C6-C3-C6 unit formed by connecting two phenolic ring moieties (A and B rings) with three carbon atoms. The C3 part can be a linear chain or form a five-membered or six-membered oxygen heterocycle with the C6 portion. Depending on differences in whether the C3 portion is cyclized, the degree of oxidation, and the substitution patterns, flavonoids can be further classified into various subclasses, including chalcones, flavones, anthocyanidins, flavanones, flavonols, and isoflavones [21]. Phenolic compounds in plants have various biosynthetic pathways, with the biosynthesis of flavonoids being one of the more extensively studied and understood processes. First, phenylalanine ammonia lyase, as the first crucial enzyme in plant phenolic compound production, catalyzes the deamination of phenylalanine to form trans-cinnamic acid. Subsequently, under the action of cinnamate 4-hydroxylase, cinnamic acid is hydroxylated, forming p-coumaric acid. Then, p-coumaric acid, in conjunction with coenzyme A, is converted to p-coumaroyl-CoA by the action of 4-coumaroyl:CoA ligase. This p-coumaroyl-CoA is further catalyzed to form chalcone when combined with malonyl-CoA, and this product serves as a precursor for various flavonoid compounds (Figure 1C). Many natural phenolic compounds are beneficial for human health. Some extensively researched examples include resveratrol, a stilbene found in grapes and red wine; epigallocatechin-3-gallate, a flavanol found in tea; and quercetin and rutin, flavonols present in various plants such as tea, onions, and Ginkgo biloba [21].

3. Production of Plant Secondary Metabolites Using Natural Hosts

Heterologous synthesis in hosts like N. benthamiana, tomatoes, and rice is primarily aimed at elucidating and confirming the synthetic pathways or enhancing crop nutrition, and the yields are generally low [6]. Natural hosts for secondary metabolites inherently possess complete metabolic pathways, and their cells may be more tolerant of the accumulation of their own secondary metabolites due to the presence of compartmentalization, metabolic enzymes, and transport systems. These allow for direct in vitro cultivation to produce secondary metabolites.

3.1. Comparison of Different In Vitro Plant Culture Systems

Currently, many different types of plant secondary metabolites have been successfully produced through in vitro cultivation, including suspension cells, adventitious roots, and hairy roots (Table 1). In comparison to traditional cultivation and natural plant growth, in vitro plant culture systems offer several advantages: not influenced by geographical or seasonal changes; provides a controlled system, ensuring the continuity, uniformity, and yield of products; allows for the isolation and purification of target compounds from the culture medium, reducing downstream costs; permits control over cell growth and regulation of metabolic processes.
Suspension cells, similar to microorganisms, grow as single cells or small cell clusters. Therefore, when it comes to scaling up production, they can directly inherit the experience and equipment used in microbial fermentation, making them easier to work with compared to tissue and organ cultures. Historically, several alkaloids, such as berberine (13.2% dry weight), polyphenols like shikonin (12% dry weight), and diterpenes like Taxol (75,000 L), have been successfully produced on an industrial scale using suspension cell cultures [22,23,24,25]. In addition, triterpenoid ginsenosides, as well as other polyphenolic compounds like resveratrol, shikonin, and rosmarinic acid, can also be efficiently synthesized in suspension cells, achieving production levels of up to grams per liter (Table 1).
Table 1. Typical examples of plant secondary metabolites production in plant chassis.
Table 1. Typical examples of plant secondary metabolites production in plant chassis.
CompoundProductHostStrategy or ObservationsPlatformTiterVesselRef.
Terpenoids
Monoterpenes Molecules 29 02106 i001
Geraniol
Nicotiana tabacumStable transgene of Valeriana officinalis geraniol synthase;
Culture condition optimization.
Suspension cells5.2 mg/LFlask[26]
Nicotiana tabacumStable transgene of Valeriana officinalis geraniol synthaseHairy roots0.11 mg/LWave-mixed bag bioreactor[27]
Molecules 29 02106 i002
Limonene
Mentha piperitaOverexpression of limonene synthase and co-suppression of limonene-3-hydroxylaseLeaves3.87 mg/g FW (79% of total essential oil)Flask[28]
Nicotiana tabacumOverexpression of plastid targeting amorpha-4,11-diene synthase and FPP synthaseLeaves0.5 μg/g FWFlask[29]
Calotropis proceraFe3O4 nanoparticles and salicylic acid reduced limonene contentsHairy roots7.9% of total essential oilFlask[30]
Molecules 29 02106 i003
α-pinene
Levisticum officinaleOptimization of culture mediumHairy roots1.3% of total essential oilFlask[31]
Molecules 29 02106 i004
Linalool
Bursera linanoeScreened 6-benzylaminopurine concentration for callus inductionSuspension cells3.02 mg/g DWStirred tank bioreactor[32]
SesquiterpenesMolecules 29 02106 i005
Patchoulol
Pogostemon cablinTransient overexpression of a MYB transcription factorLeaves2.8 mg/g FW-[33]
Physcomitrella patensOverexpression of patchoulol synthase and HMGR geneMoss tissues1.34 mg/g DW-[34]
Molecules 29 02106 i006
β-eudesmol
Atractylodes lanceaElicitation with extracts from endophytic fungiSuspension cells63 μg/g FWFlask[35]
Atractylodes lancea-Hairy roots5 μg/g FWFlask[36]
Molecules 29 02106 i007
Artemisinin
Artemisia annuaElicitation by oligosaccharide from Fusarium oxysporum mycelium combined with NO donor sodium nitroprussideHairy roots28.5 mg/LFlask[37]
Artemisia annuaDesign of bioreactorShoot culture48.2 mg/LNutrient mist bioreactor[38]
Artemisia annuaCoronatine pretreatment and addition of 30 g/L sorbitolSuspension cells9.33 mg/LFlask[39]
Essential oil Anethum graveolens-Hairy roots0.02% v/w FWFlask[40]
Achillea millefoliumAddition of Miglyol 812Suspension cells0.002%, w/w FWFlask[41]
Achillea millefolium-Hairy roots0.05% v/w FWFlask[42]
Levisticum officinaleOptimization of culture mediumHairy roots0.018 v/w FWFlask[31]
DiterpenesMolecules 29 02106 i008
Sclareol
Physcomitrella patensHeterologous overexpression of two sclareol synthase genesMoss tissues2.28 mg/L-[43]
Molecules 29 02106 i009
Aethiopinone
Salvia sclareaOverexpression of CPPSHairy roots208.98 mg/LFlask[44]
Molecules 29 02106 i010
Ferruginol
230.4 mg/L
Molecules 29 02106 i011
Paclitaxel
Taxus chinensisElicitation and optimization of culture conditionsSuspension cells902 mg/LStirred tank bioreactor[45]
Taxus MediaElicitation by MeJA, sodium nitroprusside, and L-phenylalanineHairy roots3.18 mg/g DWFlask[46]
Molecules 29 02106 i012
Tanshinone
Salvia miltiorrhizaOverexpression of SmGGPPS and SmDXSIIHairy roots12.93 mg/g DWFlask[47]
TriterpenesMolecules 29 02106 i013
Ginsenosides
Panax ginsengOverexpression of PgFPSHairy roots36.42 mg/g DWFlask[48]
Panax ginsengElicitation by nitrogen-fixing bacteriaAdventitious roots105.58 mg/g DWFlask[49]
Panax sikkimensisElicitation with culture filtrates from bacteriaSuspension cells222.2 mg/LFlask[50]
Panax notoginsengAddition of conditioned mediumSuspension cells3.12 g/LAir-lift bioreactor[51]
Molecules 29 02106 i014
α-spinasterol
Platycodon grandifoliumOverexpression of PgHMGRHairy roots1.78 mg/g DWFlask[52]
Molecules 29 02106 i015
Oleanolic acid
Calendula officinalisElicitation by jasmonic acidHairy roots52.52 mg/g DWFlask[53]
Calendula officinalisElicitation by jasmonic acidSuspension cells0.84 mg/g DWFlask[54]
Molecules 29 02106 i016
Celastrol
Tripterygium wilfordiiOverexpression of TwSQS2Hairy roots2.41 mg/g DWFlask[55]
TetraterpenesMolecules 29 02106 i017
β-carotene
Morus indicaOverexpression of β-carotene hydroxylase 1 and treated with high lightPlants256 mg/g FWFlask[56]
Daucus carotaOptimization of culture mediumSuspension cells2.19 mg/g DWFlask[57]
Molecules 29 02106 i018
Astaxanthin
Solanum lycopersicumCo-expression of the algal β-carotene ketolase from Chlamydomonas reinhardtii and β-carotene hydroxylase from Haematococcus pluvialisPlant fruits16.1 mg/g DW [58]
Phenols
AnthocyaninsMolecules 29 02106 i019
Anthocyanin
Perilla frutescensOptimization of sugar supply and culture modesSuspension cells5.8 g/LFlask[59]
Nicotiana tabacumCo-expression of the MYB and bHLH transcription factors from Antirrhinum majusSuspension cells25 mg/g DWFlask and small stirred tank[60]
Aralia cordataAdministration of CO2Suspension cells1.09 g/LJar fermentor[61]
Proanthocyanidin
Molecules 29 02106 i020
Proanthocyanidin
Litchi chinensisOverexpression of LcMYB1Hairy roots15 mg/g FWFlask[62]
Molecules 29 02106 i021
(-)-Epicatechin 3-O-gallate
Fagopyrum esculentum-Hairy roots10 mg/g DWFlask[63]
Molecules 29 02106 i022
(+)-catechin
Taxus cuspidataElicitation with MeJASuspension cells34 mg/g DWFlask[64]
Molecules 29 02106 i023
(-)-epicatechin
52 mg/g DW
FlavanolsMolecules 29 02106 i024
Rutin
Fagopyrum tataricumOverexpression of UGT73BE5Hairy roots80 mg/g DWFlask[65]
Molecules 29 02106 i025
Quercetin
Polygonum multiflorumElicitation by MeJAHairy roots17.58 mg/g DWFlask[66]
Polygonum multiflorumElicitation by jasmonic acidSuspension cells0.16 mg/g DWFlask[67]
Polygonum multiflorumOptimization of culture parametersAdventitious roots3.5 mg/g DWBalloon-type bubble (air-lift) bioreactor[68]
RutinFicus deltoideaMedium optimizationSuspension cells39.13 mg/g DWFlask[69]
Quercetin3.92 mg/g DW
IsoflavonesMolecules 29 02106 i026
Genistein
Trifolium pratenseOptimization of light conditionsHairy roots2.45 mg/g DWFlask[70]
Psoralea corylifoliaElicitation with spermidineSuspension cells4.75 mg/g DWFlask[71]
StilbenoidsMolecules 29 02106 i027
Resveratrol
Arachis hypogaea L.Elicited with MeJA and cyclodextrinHairy roots5.3 mg/g DWFlask[72]
Vitis viniferaElicitation by MeJASuspension cells209 mg/LStirred bioreactor[73]
Vitis viniferaTreated with modified β-cyclodextrinSuspension cells5 g/LFlask[74]
NaphthoquinoneMolecules 29 02106 i028
Shikonin
Lithospermum erythrorhizonOptimization of culture parametersSuspension cells4 g/LStirred tank bioreactor[24]
Hydroxycinnamic acidsMolecules 29 02106 i029
Rosmarinic acid
Anchusa officinalisHigh-density culture with optimized culture parametersSuspension cells3.7 g/LStirred tank bioreactor[75]
Molecules 29 02106 i030
Chicoric acid
Echinacea purpureaAddition of 2 mg/L indole butyric acidAdventitious roots22 mg/g DWAir-lift bioreactor[76]
Molecules 29 02106 i031
Chlorogenic acid
5 mg/g DW
Alkaloids
Tetrahydroisoquinoline alkaloidsMolecules 29 02106 i032
Berberine
Thalictrum minusCells immobilized in calcium alginate beadsSuspension cells875 mg/LModified bioreactor[77]
Coptis japonicaOptimization of culture parametersSuspension cells3.5 g/LStirred bioreactor[23]
Tropane alkaloidsMolecules 29 02106 i033
Scopolamine
Duboisia myoporoidesRepeated selection of transformed root linesHairy roots32 mg/g DWFlask[78]
Pyridine alkaloidsMolecules 29 02106 i034
Nicotine
Nicotiana tabacumOptimization of culture parametersHairy roots83.3 mg/LStirred bioreactor[79]
Nicotiana tabacumOptimization of culture parametersSuspension cells780 μg/g FWFlask[80]
Quinoline alkaloidsMolecules 29 02106 i035
Camptothecin
Ophiorrhiza pumilaIn situ absorption by polystyrene resinHairy roots3.2 mg/LFlask[81]
Ophiorrhiza pumilaOverexpression of OpG10H and OpSLSHairy roots3.5mg/g FWFlask[82]
BenzylIsoquinoline alkaloidsMolecules 29 02106 i036
Morphine
Papaver somniferumRemoval of hormones from the mediumSuspension cells2.5 mg/g DWFlask[83]
Papaver somniferum-Hairy roots2.6 mg/g DWFlask[84]
Papaver bracteatumElicitation by hormone indole-3-acetic acidSuspension cells243.2 mg/g FWFlask[85]
Papaver bracteatumOverexpression of codeinone reductaseHairy roots2.8 mg/g DWFlask[86]
Papaver orientaleElicitation with salicylic acidHairy roots2.9 mg/g DWFlask[87]
Terpenoid indole alkaloidTotal terpenoid indole alkaloids
Molecules 29 02106 i037
Vincristine
Catharanthus roseusOverexpression of ORCA3 with precursors feedingHairy roots5 mg/g DWFlask[88]
Catharanthus roseusCo-overexpression of ORCA3 and SGD induced with 3 μM dexamethasoneHairy roots10 mg/g DWFlask[89]
The synthesis of plant secondary metabolites is often subject to complex spatial (subcellular structures, tissues, organs, and entire plants) and temporal (different developmental stages) regulation. From an evolutionary perspective, this synthesis and regulation in plants are significant because they help avoid self-intoxication, enable better control of metabolic flux (by adding additional regulation and reducing feedback inhibition), and provide enzymes with more suitable catalytic environments. This complexity makes the synthesis of metabolites with complex biosynthetic pathways within a single cell a significant challenge. Selecting a high-yielding and fast-growing suspension cell line is a crucial step. Taking Taxol production as an example, among the eight species of Taxus and two hybrids, T. baccata and T. chinensis are the main high-yield producers [45,90]. Similarly, for morphine production, the suspension cell line from Papaver bracteatum can achieve an approximately 10 times higher synthesis of morphine compared to P. somniferum, 243.2 mg/g FW and 2.5 mg/g DW, respectively [83,85]. Aside from the influence of plant species, as suspension cell lines are sub-cultured, they often undergo varying degrees of differentiation, which can also affect yield stability. Additionally, the reduction in yield after long-term cultivation is also related to epigenetics [91]. In a study on the production of indole alkaloids by Catharanthus roseus suspension cells, it was found that, after 8 years of continuous cultivation, the total yield gradually decreased. However, researchers were still able to identify some cells with highly productive phenotypes, demonstrating that the initial callus tissue is the source of subsequent heterogeneity [92]. Lee et al. [93] partially addressed this issue by isolating and culturing undifferentiated cambial meristematic cells from T. cuspidata, achieving a high content of 263 mg/kg FW. Nevertheless, plant stem cell isolation technology, particularly for herbaceous plants, still requires improvement, and therefore, this technology has not seen more widespread application.
Plant organs have specialized tissue structures and do not experience the yield instability caused by the redifferentiation of suspension cells [94]. Particularly, roots have been found in recent years to be capable of achieving higher yields compared to their corresponding suspension cell lines. Industrial productions of ginsenosides (20,000 L) using adventitious roots of Panax ginseng has already been achieved [95]. Moreover, various compounds such as artemisinin in Artemisia annua, oleanolic acid in Calendula officinalis, quercetin in Polygonum multiflorum, genistein in Trifolium pratense, oleanolic acid in Calendula officinalis, celastrol in Tripterygium wilfordii, tropane alkaloids in Duboisia [96], pyridine alkaloids in Nicotiana tabacum, as well as terpenoid indole alkaloids and quinoline alkaloids, all show higher yields in hairy roots (Table 1).
However, roots are not beneficial for the production of all secondary metabolites. Some secondary metabolites show no significant difference in content between hairy and natural roots, such as tanshinone in Salvia miltiorrhiza [97], and benzyl isoquinoline alkaloids in P. somniferum (opium poppy) and Eschscholzia californica [98]. Compared to suspension cells, the content of the sesquiterpene β-eudesmol in Atractylodes lancea adventitious roots is lower than that in suspension cells (5 and 18 μg/g FW, respectively) [35,36]. The content of benzyl isoquinoline alkaloids, such as morphine, in P. bracteatum is also higher in suspension cells compared to hairy roots. Berberine, shikonin, and Taxol, as mentioned earlier, also have significantly higher content in suspension cells compared to roots. Adventitious roots and hairy roots exhibit comparable abilities in the synthesis of many secondary metabolites, but the latter often demonstrate a more stable morphology, possibly due to the integration of plant hormone biosynthetic genes from Agrobacterium rhizogenes into the genomes. For instance, in our study, the adventitious roots of Tripterygium wilfordii tend to aggregate and undergo partial dedifferentiation [99], while hairy roots do not [100]. Additionally, unlike adventitious roots, hairy roots appear to have lower capability for secreting secondary metabolites into the culture medium, such as the indole alkaloids of Catharanthus roseus [101], tropane alkaloids of Duboisia leichhardtii [102], sesquiterpene pyridine alkaloids of Tripterygium wilfordii [103], terpenoids of Salvia miltiorrhiza [104], and red beet pigments [105]. This difference may also be attributed to the genetic manipulation of hairy roots.
Comparing suspension cells and root cultures, it appears that structurally complex compounds often require cell organization with differentiation, as found in root or organ forms, to be effectively synthesized. This is particularly true for various classes of alkaloids and triterpenes. However, some structurally simple secondary metabolites, like essential oils, do not consistently demonstrate high production in suspension cells [31,40]. Due to the diverse sources of suspension cells and the significant variation in metabolite production across different tissues, it is not easy to establish a universal principle based on existing data. Therefore, for specific secondary metabolites, screening in different in vitro culture systems remains necessary.

3.2. Enhancing the Yield of Plant in Vitro Culture Systems by Process Optimization

Due to the close connection between primary and secondary metabolism in plants, any factor that affects primary metabolism may also influence secondary metabolism. Therefore, adjusting the growth environment can further enhance the yield of secondary metabolites. This can include optimizing the types and concentrations of nutrients and plant hormones in the culture medium, as well as optimizing aeration rate, pH, and inoculum size [4]. Additionally, the addition of precursors, elicitors, and immobilization techniques may also further increase production [106]. In particular, methods that mimic biotic stress have been widely applied [107,108]. For example, jasmonic acid and its methyl derivative, methyl jasmonate (MeJA), which are important plant stress signaling molecules, can induce the biosynthesis of defensive compounds [109]. Secondary metabolites secreted into the culture medium may be degraded by metabolic enzymes in the medium, and the accumulated secondary metabolites themselves could be harmful to plant cells and/or to the feedback inhibition of biosynthesis. Therefore, various methods, such as in situ adsorption and two-phase culture, should be employed to remove the products from the culture medium promptly [106,110,111].
The successful commercial production of plant-based secondary metabolites depends on the scaling up of the process from shake flasks to bioreactors. Stirred-tank reactors and airlift reactors are widely used in cell suspension cultures [112]. The former are known for their simple scalability, strong oxygen supply, and uniform mixing, but they have the disadvantage of high shear forces [113]. The latter are easy to scale up, have low operating costs and contamination rates, and generate low shear forces. However, they suffer from poor oxygen mass transfer, non-uniform mixing, and a propensity for foaming [114]. The cultivation of root tissues primarily involves the use of mist bioreactors [115] and balloon-type bubble bioreactors [116]. The research group led by Kee-Yoeup Paek in South Korea successfully used a balloon-type bubble bioreactor to scale up the cultivation of Panax ginseng hairy roots to volumes ranging from 10,000 to 20,000 L [117]. Temporary immersion bioreactors and root zone immersion bioreactors are suitable for explant culture. Rotating drum bioreactors can be used for the culture of suspension cells, root tissues, and explants, but they face challenges when scaling up, and their heat transfer performance is low. Other reactor types, such as disposable reactors, membrane bioreactors, and wave bioreactors, are challenging to scale up due to their specific characteristics and are primarily suitable for small-scale cultivation [112,118]. More information on the structure and design of plant bioreactors can be found in the article by Murthy et al. [119].

3.3. Plant Metabolic Engineering

The complexity of plant secondary metabolite biosynthetic pathways results in a diverse range of strategies for manipulating metabolic pathways [120], including utilizing gene editing tools, such as Clustered, Regularly Interspaced, Short Palindromic Repeats (CRISPR)/dead CRISPR associated protein 9 (dCAS9), to activate single or multiple-target genes; reducing the expression of genes in competitive pathways through techniques such as RNA interference (RNAi) and CRISPR/CAS9; modifying the expression of transcription factors that control multiple biosynthetic genes; and enhancing the compartmentalization and/or release of the desired product by transporter engineering.
Enhancing the activity of key enzymes controlling the biosynthesis of a specific secondary metabolite through genetic engineering can increase the content of the target secondary metabolite in transgenic plants. For example, overexpressing FPP synthase, squalene synthase, or the upstream 3-hydroxy-3-methylglutaryl-coenzyme A reductase (HMGR) in the mevalonate pathway (MVA), for example, can increase the production of downstream terpenoid end-products. However, due to the presence of downstream branch pathways, increasing the expression of upstream pathway enzyme genes alone may not achieve the accumulation of the target compound. For instance, overexpressing the first two genes, tabersonine 16-hydroxylase (T16H) and 16-O-methyl transferase (16OMT), in C. roseus hairy roots failed to produce the target vindoline [121]. Using RNAi to disrupt undesired metabolic branches can redirect synthetic metabolism toward the target. Allen et al. [122] used this technique to silence the gene family of codeinone reductase (COR) in poppies, blocking the synthesis of opioid substances while leading to the significant accumulation of the precursor to codeine, (S)-reticuline. Overexpressing limonene synthase while inhibiting the expression of by-product synthetic gene limonene-3-hydroxylase substantially increased the content of limonene and its proportion in the total essential oil (up to 79%) [28].
The activity and expression of these biosynthetic enzymes are regulated by corresponding transcription factors and other regulatory genes, making metabolic engineering of transcription factors potentially highly effective [123]. Currently, transcription factors involved in plant secondary metabolite biosynthetic pathways mainly include WRKY, myeloblastosis (MYB), basic helix–loop–helix (bHLH), APETALA2/Ethylene-Responsive-Factor (AP2/ERF), jasmonate-responsive ERF (JRE), basic Leucine Zipper (bZIP), SQUAMOSA promoter-binding protein-like (SPL), and YABBY classes, among others [11,124]. Taking the indole alkaloid biosynthetic pathway as an example, the overexpression of ORCA3 together with the addition of the terpene precursor [88] or overexpression of downstream key enzyme genes [89] can increase the final yield of the target compound.
The excessive accumulation of secondary metabolites within cells can have two main effects: on one hand, it can increase cellular toxicity and hinder cell growth; on the other hand, it can result in the feedback inhibition of synthetic enzyme activity, reducing overall biosynthesis [125,126]. Metabolic flux balance analysis indicates that the extracellular secretion and transfer of secondary metabolites are among the key limiting steps for improving yield [127]. Additionally, due to the simplicity of the culture medium, the economic aspects of purifying secondary metabolites through medium separation are favorable. Avoiding cell disruption can further save on cultivation and production costs [128]. Therefore, enhancing the extracellular secretion of secondary metabolites through genetic engineering is of significant importance in plant or microbial biosynthesis research [127,129]. Two main types of proteins have been identified to be involved in the transport of plant secondary metabolites: ABC (ATP-binding cassette) and MATE (Multidrug and toxic compound extrusion) [130,131]. Among these, ABC transport proteins exhibit higher substrate specificity compared to MATE, and they are more frequently associated with plant disease resistance [132], especially the pleiotropic drug resistance (PDR) transport proteins [133,134]. Pomahačová et al. [135] heterologously expressed the transporter gene CjMDR1, which mediates the intracellular transport of berberine, in C. roseus. This led to an increase in the intracellular content of ajmalicine and tetrahydroalstonine in suspension cell lines. Our research group successfully enhanced the extracellular secretion of sesquiterpene pyridine alkaloids and triptolide in Tripterygium wilfordii by regulating influx and efflux transporter proteins, respectively [103,136]. However, the application of transporter proteins in metabolic engineering is still limited, as only a small number of plant secondary metabolite transporter proteins have been functionally characterized, and research on the post-translational regulation of transporter protein activity is lacking [137]. Our recent study elucidated the phosphorylation regulation mechanism of the AtPDR6 transporter protein, which mediates the extracellular secretion of camalexin in Arabidopsis thaliana [138], laying a foundation for subsequent transporter engineering.

3.4. Multigene Transformation of Plants

Plant secondary metabolism involves the expression of multiple genes, and enhancing the co-expression of multiple genes is essential for further increasing the yield of secondary metabolites. The invention of large DNA fragment assembly technologies has made the construction of large fragment vectors more convenient. Some widely used techniques include Zinc Finger Nuclease and Homing Endonuclease-Mediated Assembly [139], ligation cycling reaction [140], Golden Gate cloning [141], Cre/loxP recombination [142], MultiRound Gateway recombination [143], Gibson Assembly [144], and yeast assembly [145]. However, due to the possibility of multi-gene silencing, rearrangement, or internal interactions in plants [146], successfully transforming vectors containing hundreds of thousands of base pairs into plant cells and achieving stable expression in plants, especially non-model plants, is one of the major challenges in plant metabolic engineering [147,148]. The mainstream methods of introducing large DNA into plants include particle bombardment, polyethylene glycol-mediated protoplast transformation, and Agrobacterium-mediated transformation. The first two methods involve the induction of callus tissue or regeneration of the cell wall, which can be challenging for many plants. For the last method, Agrobacterium cannot infect many medicinal plants or has very low efficiency [149]. Additionally, when co-expressing multiple genes, it is important to adjust the strength of gene promoters and carefully choose the insertion sites to prevent gene silencing [150]. Placing multiple pathway enzyme genes into artificial plant chromosomes can help avoid position effects associated with genome insertion, potentially achieving stable expression [151,152]. However, this technology still faces challenges, such as the potential inability to form active centromeres and the risk of abnormal plant growth or death [153].

3.5. Elucidation of Biosynthetic Pathway

All of the aforementioned strategies for metabolic regulation require an in-depth understanding of the biosynthetic pathways of secondary metabolites. The redundancy of pathway enzyme genes and challenges in substrate accumulation or chemical synthesis hinder the specific identification of enzyme genes for certain steps [8,154], and only widely studied natural products have been fully elucidated (Figure 1). Comparing gene expression levels between different biotic/abiotic stresses, elicitors, and tissues is an effective strategy for identifying biosynthetic genes. These co-expression data are typically analyzed through linear correlation analyses, such as linear regression and/or hierarchical clustering based on Euclidean distance, to extract potential genes with similar functions to the “bait” gene [155].
In recent years, the functional elucidation of candidate genes in a biosynthetic pathway can be achieved by expressing them constitutively in N. benthamiana, without the need to consider synthesis order or prepare large amounts of intermediates. For example, Caputi et al. [156] identified the enzymes responsible for synthesizing tabersonine 1 and catharanthine 3 by overexpressing three enzymes, precondylocarpine acetate synthase, dihydroprecondylocarpine synthase, and tabersonine synthase/catharanthine synthase, in N. benthamiana. By rerouting diterpene biosynthesis from the chloroplast to the cytosol, De La Peña and Sattely [157] achieved the complete elucidation of the momilactone biosynthetic pathway in tobacco. However, due to the influence of endogenous metabolic pathways on the “orthogonality” of the transgenic pathway in N. benthamiana and its lack of accumulation of some necessary precursors, it often requires the simultaneous infiltration of precursor molecules during transient transformation or the use of other hosts. Metabolic engineering operations in yeast can help construct synthesis pathways for pathway intermediates or test the function of specific enzymes by adding substrates. For example, using this approach, the P450 enzyme CYP71AV1 involved in the biosynthesis of artemisinin was functionally identified, and the precursor, artemisinic acid, was abundantly produced [11].

4. Production of Plant Secondary Metabolites Using Microbial Chassis

Compared to plant hosts, microorganisms have gained popularity in recent years due to their faster growth rates, simpler and clearer metabolic networks, ease of genetic transformation, and scalability (Table 2). Among microorganisms, Escherichia coli and yeast are more widely used than others, mainly due to their well-defined metabolic pathways, more stable gene expression systems, and efficient, abundant biosynthetic elements [7].

4.1. Engineering Enzymes and Cofactors

Once the biosynthetic pathway of plant secondary metabolites is elucidated, the genes encoding pathway enzymes can theoretically be introduced into microorganisms to produce these secondary metabolites. However, enzymes from the native host may have no activity or poor activity in microorganisms. Through genome and transcriptome mining, homologous enzymes or isoenzymes with potential catalytic activity can be identified and screened. For example, Cheng et al. [162] identified the most active limonene synthase from nine candidate enzymes obtained from the genera Citrus, Solanum habrochaites, and Agastache rugosa. Jiang et al. [159] discovered an enzyme with higher activity from C. roseus through the screening of nine geraniol synthase genes from diverse species. Additionally, there is a wealth of metagenomic resources in unculturable environmental microorganisms that can be explored for genes encoding enzymes with potential activities [208].
After obtaining candidate pathway enzyme genes, combinatorial optimization can be used to find the optimal combination. For example, Clomburg et al. [209] combined genes from Staphylococcus aureus, Myxococcus xanthus, Clostridium beijerinckii, and E. coli to construct an isoprenoid alcohol pathway using a retrobiosynthetic approach to provide a high accumulation of precursors for downstream terpene synthesis. By expressing four biosynthetic enzymes from three different plants, Kim [187] achieved 18.6 mg/L of genistein from p-coumaric acid in E. coli. For synthesizing lycopene, Shi et al. [181] initially screened the activities of nine GGPP synthase, five CrtB, and seven CrtI enzymes from various hosts, followed by further screening of over thirty combinations. This process led to a significant increase in lycopene production, from 4.5 mg/L to 118 mg/L. However, this combinatorial optimization approach often involves a considerable amount of trial-and-error efforts. Fortunately, gene mining aided by bioinformatics [210] and the convenience of component assembly in synthetic biology [211] can reduce the initial screening and vector construction work.
If the enzymes discovered in microorganisms exhibit issues such as insufficient substrate specificity, known as substrate promiscuity, low activity, and poor stability, they can be modified using enzyme engineering techniques such as random mutagenesis, directed evolution, and computer-aided semi-rational design, to alter active pockets, replace amino acids in loop regions, or make other modifications. In yeast, ERG20 catalyzes the condensation of IPP and DMAPP into both GPP and FPP. Ignea et al. [212,213] introduced mutations into ERG20 using rational design, changing it into sole GPP synthase or GGPP synthase. Sometimes, the activity and (or) stability of enzymes can also be improved by fusing them with tag proteins or other enzymes. For instance, terpene synthase and lavandulyl diphosphate synthase fused with maltose-binding protein exhibited improved stability [169]. In their study, it was also found that the fused enzymes showed higher activity compared to individual expression, which was mainly due to the closely located enzymes which allowed for a more efficient substrate transfer. Zhang et al. [172] found that GGPPS and taxadiene synthase exhibit different subcellular localization in yeast and that fusion expression increases taxadiene production by 54%.
When expressing plant-derived enzymes in microorganisms, especially P450 enzymes, it is common to encounter reduced or lost activity. Approximately 40% of plant-derived P450 enzymes are expressed poorly in yeast [214]. Additionally, P450 enzymes typically require characteristic redox partners, such as cytochrome P450 reductase (CPR), to transfer electrons. However, microorganisms may lack a sufficient amount of CPR or the available CPR may not effectively couple with exogenous P450 enzymes. Possible strategies to address this issue include adjusting the expression level of P450 enzyme genes and engineering them [176], screening for better CPRs [177], and taking inspiration from the natural P450/CPR fusion observed in BM3 (CYP102A1) from Bacillus megaterium by fusing P450 enzymes with CPR [171,187].
As essential compounds for the activity of various exogenous enzymes, cofactors often face competition with endogenous enzymes, which often leads to inadequate cofactor supply and subsequent growth retardant. Due to the involvement of approximately 1600 enzyme-catalyzed reactions, the NAD(P)H/NAD(P)+ and ATP/ADP systems have attracted widespread attention [215]. Adjusting the expression of three genes in the central metabolic module of E. coli: α-ketoglutarate dehydrogenase, succinate dehydrogenase, and transaldolase B enhanced carbon flux toward the TCA cycle and the pentose phosphate module, thereby increasing the supply of NADPH and ATP [182]. Similarly, in yeast, NADPH supply can be increased by reducing the glycolytic pathway and enhancing the pentose phosphate pathway [216]. Using this strategy, Cao et al. [169] significantly increased the synthesis of the diterpene sclareol. Also in yeast, the NADH kinase POS5 can convert NADH to NADPH, and the overexpression of this gene can increase NADPH supply [181]. In addition to supplementing cofactors, it is also possible to block endogenous cofactor synthesis pathways in the host and then use the modified host for high-throughput screening of target enzymes. This approach may yield enzyme variants that allow for cofactor regeneration without affecting growth [217]. It is noteworthy that cofactors are typically synthesized and stored in specific organelles in the native host, which further complicates the challenge of expressing active P450 enzymes in microorganisms. For example, heme, primarily found in mitochondria, can hinder the construction of P450 enzyme synthesis pathways in the cytoplasm [218].

4.2. Pathway Establishment and Optimization

Microbial metabolic networks have developed tight regulation to maintain metabolic homeostasis, and introducing exogenous synthetic pathways may lead to disruptions in endogenous metabolism, increase cellular metabolic load, and the introduction of new substances that are toxic to cells. Therefore, it is essential to carefully design synthetic pathways to balance the overall metabolic flux and cofactor levels [219].
Before constructing a pathway, it is important to consider the form in which exogenous DNA exists in microorganisms. While integrating exogenous DNA into the genome is ideal due to the instability of plasmids and variations in copy number, many laboratory-scale studies still use plasmids due to their convenience. Traditional homologous recombination systems using counter-selection markers are time-consuming, so site-specific recombination and transposon-mediated random insertion are often used. The former is widely used in yeast, utilizing sites such as nutrient and amino acid synthesis gene loci, delta sites, and rDNA sites. The CRISPR/Cas system can induce site-specific breaks, allowing for knockouts, in-del mutations, and the insertion of exogenous DNA. Coupling specific sites with CRISPR/Cas9 can enable a higher insertion copy number, potentially increasing production further [220].
Manipulating individual or combinations of genes remains labor-intensive and iterative. Each change that leads to a change in phenotype needs to be considered in the next step. The ideal approach involves multiplex genome engineering, such as oligo-based, recombinase-based, and CRISPR-based methods, along with biosensors and high-throughput screening tools [221]. However, this also increases uncertainty and operational complexity, so it is not widely applied in the successful cases we have investigated.
Similar to plant metabolic engineering, microbial strategies mainly include the regulation of pathway enzyme gene expression, inhibition of branches, and control of transport. The expression levels of pathway enzyme genes are one of the crucial factors influencing product synthesis. Many studies choose to overexpress the enzyme genes responsible for precursor supply pathways to increase the capacity of precursor pools. For example, to increase upstream products of terpenoids, IDI1 and HMGR are often overexpressed [159,160]. However, it is worth noting that, during the early stages of cultivation, the overaccumulation of precursors or end products may have adverse effects on cell growth, such as precursors of terpenoids and polyphenols [160,222]. At the same time, the early expression of exogenous pathways can slow down cell growth due to increased metabolic burden. In such cases, the dynamic control of gene expression can be chosen. In the study of Xie et al. [222], the expression of genes involved in the MVA pathway and squalene synthesis pathway initially remained at normal levels. It was only after the cells reached high density that these genes were overexpressed, achieving a form of sequential control. Dynamic control is often achieved through condition- or inducible-type promoters that regulate gene expression, such as GAL promoters [223] or hexose transporter promoters [162].
Dealing with nonlinear diverging–converging metabolic pathways often involves a more precise control of the supply of two or more precursors, and adjusting the metabolic flux in one branch may affect the others. This requires laborious work to optimize gene copy numbers, expression levels, ribosome binding sites, and other factors. Separating diverging pathways and building them separately in different strains can reduce metabolic stress, create distinct cellular environments, and minimize interference from unwanted pathway modules. For example, the synthesis of rosmarinic acid was divided into three E. coli strains to provide two precursors and downstream pathways. By optimizing the balance of these three strains and the composition of the culture medium, the rosmarinic acid content was increased from 4.5 mg/L in a mono-culture to 172 mg/L [199]. Sometimes, the low substrate specificity of upstream enzymes may lead to unwanted conversions into downstream products [194]. In such cases, enzyme engineering can be used to modify the enzymes, or the pathway can be split and expressed in different strains. Horinouchi et al. [224] separated the genistein synthesis pathways and placed them in E. coli and yeast, achieving a higher production of 100 mg/L when mixed.
Similar to plant chassis, microbial heterologous synthesis faces challenges in compound transport, such as precursor uptake, material exchange between subcellular compartments, intermediate loss, and feedback inhibition due to insufficient final product secretion [225]. Zhao et al. [190] discovered that overexpressing OmpF, among the ten native transport proteins screened, significantly increased the production of the stilbene resveratrol, indicating that the screening of native transport proteins can be valuable, even though most of them may not be conducive to improving yield or cell growth when overexpressed. Anthocyanins are often stored in the central vacuole in plants, where the pH is lower than in the cytoplasm, which is more favorable for the stability of the products. Lowering the pH of the culture medium and maximizing the secretion of anthocyanins into the extracellular space can boost production. In a study that aimed to synthesize anthocyanin in E. coli, it was found that simultaneously knocking out the catechin efflux transporter tolC and overexpressing the product C3G efflux transporter yadH significantly promoted product synthesis, with tolC knockout having a greater impact [195].
Alkaloids, due to their high pKa, carry a positive charge during microbial fermentation, making them unable to cross the membrane by diffusion. Overexpressing a MATE transporter protein, AtDTX1, from Arabidopsis in the yeast cell membrane, increases the secretion of (S)-reticuline, an intermediate in benzylisoquinoline alkaloids (BIAs) biosynthesis, and enhances its production 11-fold [226]. Dastmalchi et al. [227] systematically screened and functionally identified a series of active transport proteins, BIA uptake permeases (BUPs), in opium poppy, which are involved in the intracellular transport of BIAs. Combining AtDTX1 and BUPs to regulate intermediate transport may make strategies involving the co-culture of different strains more promising for BIA production [201].

4.3. Fermentation Condition and Its Optimization

Once optimal recombinant strains have been obtained, the fermentation conditions become a critical last step in determining whether the production can be scaled up successfully. Specific conditions that affect strain performance include the composition and pH of the culture medium, environmental conditions, and the fermentation process itself. The composition of the culture medium can directly impact the yield of the recombinant strain. For instance, Liu et al. [207] found that peptones from different sources can affect the yield when synthesizing ajmalicine and sanguinarine in yeast. Additionally, the combination of components in the culture medium needs to be optimized in conjunction with the promoters used for pathway construction. For example, when using the Gal promoter, it is essential to pay attention to the ratio of galactose to other carbon sources and the timing of their addition. When using the HXT1 promoter, the concentration of glucose should be carefully managed. The pH of the culture medium plays a role in regulating aspects of the bioprocess, such as compound solubility, stability, and transport. Environmental conditions, such as aeration rate and temperature, can impact cell metabolism, growth, and product yield.
Similar to plant chassis, the timely removal or transfer of products is often used to enhance yield. For compounds with low polarity, like terpenes, two-phase culture methods that facilitate in situ product removal have been proven to significantly increase yield. For instance, adding diisononylphthalate to extract limonene has been shown to be effective [161].

5. Comparison and Prospects of Plant and Microbial Chassis

Assuming a standard of g/L as a high-yield chassis, a systematic analysis of cases for the production of plant-derived bioactive compounds in both plant and microbial chassis (Table 1 and Table 2) reveals the following:
  • In instances wherein plant chassis have succeeded and microbial chassis have not, they include the production of the diterpene paclitaxel, polyphenols such as anthocyanin, shikonin, rosmarinic acid, and the isoquinoline alkaloid berberine.
  • Microbial chassis have succeeded while plant chassis achieve lower yields for various monoterpenes, the sesquiterpene srtemisinic acid, the simple diterpene sclareol, and diterpene precursors like taxadiene and miltiradiene, the simple triterpene amyrin, linear tetraterpenes lycopene and β-carotene with little downstream tailoring enzyme modifications, polyphenolic upstream products like quercetin and (+)-catechin, and the isoquinoline alkaloid precursor (S)-reticuline.
  • Triterpenes like ginsenosides and polyphenolic compounds such as resveratrol can be produced at high yields in both plant and microbial chassis.
It is evident that microbial chassis are more efficient at producing structurally simple compounds or intermediates, while they face challenges in achieving high yields for complex bioactive compounds due to issues like interference with endogenous metabolism and low enzyme activity. For instance, even though the crucial isoquinoline alkaloid precursor (S)-reticuline can be produced at high yields in yeast (4.6 g/L) [203], it can only yield nM levels of downstream product morphine, even using (R,S)-norlaudanosoline as an externally added substrate [228]. Similarly, terpenoid indole alkaloids with complex synthesis pathways, such as vinblastine, face issues like promiscuous reactions, interactions between hosts and foreign substances, as well as challenges in expressing active enzymes. Plant chassis, especially those using natural hosts, excel in overcoming these challenges, often requiring simple fermentation condition optimizations to achieve high yields for certain bioactive components. However, due to the strict metabolic regulation and complex feedback mechanisms of plants, they may struggle to achieve high yields for even structurally simple products (Table 3).
From a strategic perspective, both plant and microbial platforms employ similar approaches, involving pathway construction or enhancement, overexpression of key enzyme genes, reduction in or elimination of competitive metabolic flux, enzyme engineering, and transport engineering (Figure 2). However, plants have complex, mostly unexplored metabolic pathways, lower genetic transformation efficiency, and a limited promoter library, making them less successful in various strategy aspects compared to microbes, such as dynamic control and precise expression regulation. Also, strategies for eliminating bottleneck steps have shown more excellent results in microbial platforms. Specifically, if the product yield significantly increases upon the addition of a specific intermediate, it indicates that the enzyme catalyzing the synthesis of that intermediate may have insufficient activity or the intermediate is being used by the endogenous pathway [144]. If no intermediates are available for use, one can consider using LC-MS to measure the abundance of metabolites throughout the entire pathway [203].
Although microbial chassis have the advantages and achievements mentioned above, they also face their own challenges. Since endogenous metabolic pathways are interconnected, optimal regulation strategies cannot be found without globally controlling the distribution of metabolic flux. For example, in the synthesis of tropine alkaloids using yeast, it was surprisingly found that removing leucine auxotrophy by expressing 3-isopropylmalate dehydrogenase significantly increased the conversion of hyoscyamine (85% decrease) to scopolamine (more than 3-fold increase) [229]. Current analysis of global metabolic flux can only be performed through somewhat unsatisfactory model construction and simulation [230], or more promisingly, using artificial intelligence [231,232]. However, they all rely on repeatable and reliable datasets, while standardized strains and operations are difficult to unify in the academic community. As mentioned above, transport engineering of secondary metabolites is a highly promising strategy. However, we currently have many gaps in the understanding of the functions and activity regulation mechanisms of plant secondary metabolite transporters and even microbial endogenous transporters. Therefore, this strategy has not yet reached its maximum potential.
In summary, in the current research progress, both microbial and plant platforms have seen successful cases. Due to the clearer metabolic pathways of microbes and the ease of genetic transformation and scale-up cultivation, they have achieved a series of remarkable accomplishments, especially eukaryotic yeasts, in the synthesis of various structurally uncomplicated compounds. In the future, if microbial platforms can successfully address global metabolic regulation, maintenance of heterologous enzyme activity, and issues related to metabolite transport, it is believed that even more success can be achieved. As our understanding of plants endogenous metabolism becomes clearer and more convenient genetic transformation methods and transcriptional regulatory elements are developed, the successful experiences on microbial platforms may guide plants to become a more versatile host.

Author Contributions

The manuscript was drafted by T.H. and G.M. and was revised by T.H. and G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China [31800256] and the Natural Science Foundation of Anhui Province [2108085MC84]. The funders had no role in the study design, data collection, analysis, decision to publish, or preparation of the manuscript.

Data Availability Statement

All data are contained within the article.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Kessler, A.; Kalske, A. Plant secondary metabolite diversity and species interactions. Annu. Rev. Ecol. Evol. Syst. 2018, 49, 115–138. [Google Scholar] [CrossRef]
  2. Hussain, M.S.; Fareed, S.; Ansari, S.; Rahman, M.A.; Ahmad, I.Z.; Saeed, M. Current approaches toward production of secondary plant metabolites. J. Pharm. Bioallied Sci. 2012, 4, 10. [Google Scholar] [CrossRef]
  3. Sreenivasulu, N.; Fernie, A.R. Diversity: Current and prospective secondary metabolites for nutrition and medicine. Curr. Opin. Biotechnol. 2022, 74, 164–170. [Google Scholar] [CrossRef] [PubMed]
  4. Marchev, A.S.; Yordanova, Z.P.; Georgiev, M.I. Green (cell) factories for advanced production of plant secondary metabolites. Crit. Rev. Biotechnol. 2020, 40, 443–458. [Google Scholar] [CrossRef]
  5. Pyne, M.E.; Narcross, L.; Martin, V.J. Engineering plant secondary metabolism in microbial systems. Plant Physiol. 2019, 179, 844–861. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, X.; Zhang, P.; Zhao, Q.; Huang, A.C. Making small molecules in plants: A chassis for synthetic biology-based production of plant natural products. J. Integr. Plant Biol. 2023, 65, 417–443. [Google Scholar] [CrossRef]
  7. Liu, J.; Wang, X.; Dai, G.; Zhang, Y.; Bian, X. Microbial chassis engineering drives heterologous production of complex secondary metabolites. Biotechnol. Adv. 2022, 59, 107966. [Google Scholar] [CrossRef]
  8. Twaij, B.M.; Hasan, M.N. Bioactive secondary metabolites from plant sources: Types, synthesis, and their therapeutic uses. Int. J. Plant Biol. 2022, 13, 4–14. [Google Scholar] [CrossRef]
  9. Wang, Z.; Nelson, D.R.; Zhang, J.; Wan, X.; Peters, R.J. Plant (di) terpenoid evolution: From pigments to hormones and beyond. Nat. Prod. Rep. 2023, 40, 452–469. [Google Scholar] [CrossRef]
  10. Bergman, M.E.; Davis, B.; Phillips, M.A. Medically useful plant terpenoids: Biosynthesis, occurrence, and mechanism of action. Molecules 2019, 24, 3961. [Google Scholar] [CrossRef]
  11. Nagegowda, D.A.; Gupta, P. Advances in biosynthesis, regulation, and metabolic engineering of plant specialized terpenoids. Plant Sci. 2020, 294, 110457. [Google Scholar] [CrossRef] [PubMed]
  12. Karunanithi, P.S.; Zerbe, P. Terpene synthases as metabolic gatekeepers in the evolution of plant terpenoid chemical diversity. Front. Plant Sci. 2019, 10, 1166. [Google Scholar] [CrossRef] [PubMed]
  13. Boncan, D.A.T.; Tsang, S.S.; Li, C.; Lee, I.H.; Lam, H.-M.; Chan, T.-F.; Hui, J.H. Terpenes and terpenoids in plants: Interactions with environment and insects. Int. J. Mol. Sci. 2020, 21, 7382. [Google Scholar] [CrossRef] [PubMed]
  14. Bhambhani, S.; Kondhare, K.R.; Giri, A.P. Diversity in chemical structures and biological properties of plant alkaloids. Molecules 2021, 26, 3374. [Google Scholar] [CrossRef] [PubMed]
  15. Omar, F.; Tareq, A.M.; Alqahtani, A.M.; Dhama, K.; Sayeed, M.A.; Emran, T.B.; Simal-Gandara, J. Plant-based indole alkaloids: A comprehensive overview from a pharmacological perspective. Molecules 2021, 26, 2297. [Google Scholar] [CrossRef] [PubMed]
  16. Singh, S.; Pathak, N.; Fatima, E.; Negi, A.S. Plant isoquinoline alkaloids: Advances in the chemistry and biology of berberine. Eur. J. Med. Chem. 2021, 226, 113839. [Google Scholar] [CrossRef] [PubMed]
  17. Huang, J.-P.; Wang, Y.-J.; Tian, T.; Wang, L.; Yan, Y.; Huang, S.-X. Tropane alkaloid biosynthesis: A centennial review. Nat. Prod. Rep. 2021, 38, 1634–1658. [Google Scholar] [CrossRef] [PubMed]
  18. Pereira, A.G.; Cassani, L.; Garcia-Oliveira, P.; Otero, P.; Mansoor, S.; Echave, J.; Xiao, J.; Simal-Gándara, J.; Prieto, M. Plant Alkaloids: Production, Extraction, and Potential Therapeutic Properties. In Natural Secondary Metabolites: From Nature, through Science, to Industry; Springer: Berlin/Heidelberg, Germany, 2023; pp. 157–200. [Google Scholar]
  19. Hano, C.; Tungmunnithum, D. Plant polyphenols, more than just simple natural antioxidants: Oxidative stress, aging and age-related diseases. Medicines 2020, 7, 26. [Google Scholar] [CrossRef] [PubMed]
  20. Truzzi, F.; Tibaldi, C.; Zhang, Y.; Dinelli, G.; D′Amen, E. An overview on dietary polyphenols and their biopharmaceutical classification system (BCS). Int. J. Mol. Sci. 2021, 22, 5514. [Google Scholar] [CrossRef]
  21. Leri, M.; Scuto, M.; Ontario, M.L.; Calabrese, V.; Calabrese, E.J.; Bucciantini, M.; Stefani, M. Healthy Effects of Plant Polyphenols: Molecular Mechanisms. Int. J. Mol. Sci. 2020, 21, 1250. [Google Scholar] [CrossRef]
  22. Sato, F.; Yamada, Y. High berberine-producing cultures of Coptis japonica cells. Phytochemistry 1984, 23, 281–285. [Google Scholar] [CrossRef]
  23. Matsubara, K.; Kitani, S.; Yoshioka, T.; Morimoto, T.; Fujita, Y.; Yamada, Y. High density culture of Coptis japonica cells increases berberine production. J. Chem. Technol. Biotechnol. 1989, 46, 61–69. [Google Scholar] [CrossRef]
  24. Fujita, Y. Shikonin: Production by plant (Lithospermum erythrorhizon) cell cultures. In Medicinal and Aromatic Plants I; Springer: Berlin/Heidelberg, Germany, 1988; pp. 225–236. [Google Scholar]
  25. Expósito, O.; Bonfill, M.; Moyano, E.; Onrubia, M.; Mirjalili, M.; Cusido, R.; Palazon, J. Biotechnological production of taxol and related taxoids: Current state and prospects. Anti-Cancer Agents Med. Chem. (Former. Curr. Med. Chem.-Anti-Cancer Agents) 2009, 9, 109–121. [Google Scholar] [CrossRef] [PubMed]
  26. Vasilev, N.; Schmitz, C.; Grömping, U.; Fischer, R.; Schillberg, S. Assessment of cultivation factors that affect biomass and geraniol production in transgenic tobacco cell suspension cultures. PloS one 2014, 9, e104620. [Google Scholar] [CrossRef] [PubMed]
  27. Ritala, A.; Dong, L.; Imseng, N.; Seppänen-Laakso, T.; Vasilev, N.; van der Krol, S.; Rischer, H.; Maaheimo, H.; Virkki, A.; Brändli, J. Evaluation of tobacco (Nicotiana tabacum L. cv. Petit Havana SR1) hairy roots for the production of geraniol, the first committed step in terpenoid indole alkaloid pathway. J. Biotechnol. 2014, 176, 20–28. [Google Scholar] [CrossRef] [PubMed]
  28. Mahmoud, S.S.; Williams, M.; Croteau, R. Cosuppression of limonene-3-hydroxylase in peppermint promotes accumulation of limonene in the essential oil. Phytochemistry 2004, 65, 547–554. [Google Scholar] [CrossRef] [PubMed]
  29. Wu, S.; Schalk, M.; Clark, A.; Miles, R.B.; Coates, R.; Chappell, J. Redirection of cytosolic or plastidic isoprenoid precursors elevates terpene production in plants. Nat. Biotechnol. 2006, 24, 1441–1447. [Google Scholar] [CrossRef] [PubMed]
  30. Adabavazeh, F.; Pourseyedi, S.; Nadernejad, N.; Razavizadeh, R.; Mozafari, H. Evaluation of synthesized magnetic nanoparticles and salicylic acid effects on improvement of antioxidant properties and essential oils of Calotropis procera hairy roots and seedlings. Plant Cell Tissue Organ Cult. 2022, 151, 133–148. [Google Scholar] [CrossRef]
  31. Santos, P.A.; Figueiredo, A.C.; Oliveira, M.M.; Barroso, J.G.; Pedro, L.G.; Deans, S.G.; Scheffer, J.J. Growth and essential oil composition of hairy root cultures of Levisticum officinale WDJ Koch (lovage). Plant Sci. 2005, 168, 1089–1096. [Google Scholar] [CrossRef]
  32. Pavón-Reyes, L.; Evangelista-Lozano, S.; Sepúlveda-Jiménez, G.; Ávila, V.C.; Rodríguez-Monroy, M. Cell culture of Bursera linanoe in a stirred tank bioreactor for production of linalool and linalyl acetate. Nat. Prod. Commun. 2017, 12, 1934578X1701200301. [Google Scholar]
  33. Chen, X.; Li, J.; Liu, Y.; Wu, D.; Huang, H.; Zhan, R.; Chen, W.; Chen, L. PatSWC4, a methyl jasmonate-responsive MYB (v-myb avian myeloblastosis viral oncogene homolog)-related transcription factor, positively regulates patchoulol biosynthesis in Pogostemon cablin. Ind. Crops Prod. 2020, 154, 112672. [Google Scholar] [CrossRef]
  34. Zhan, X.; Zhang, Y.-H.; Chen, D.-F.; Simonsen, H.T. Metabolic engineering of the moss Physcomitrella patens to produce the sesquiterpenoids patchoulol and α/β-santalene. Front. Plant Sci. 2014, 5, 636. [Google Scholar] [CrossRef] [PubMed]
  35. Fang, F.; Dai, C.; Wang, Y. Role of nitric oxide and hydrogen peroxide in the essential oil increasing of suspension cells from Atractylodes lancea induced by endophytic fungal Cunninghamella sp. AL4 elicitor. Sheng Wu Gong Cheng Xue Bao = Chin. J. Biotechnol. 2009, 25, 1490–1496. [Google Scholar]
  36. Zhang, C.; Guo, X.; Wang, H.; Dai, X.; Yan, B.; Wang, S.; Guo, L. Induction and metabolomic analysis of hairy roots of Atractylodes lancea. Appl. Microbiol. Biotechnol. 2023, 107, 6655–6670. [Google Scholar] [CrossRef] [PubMed]
  37. Zheng, L.P.; Guo, Y.T.; Wang, J.W.; Tan, R.X. Nitric oxide potentiates oligosaccharide-induced artemisinin production in Artemisia annua hairy roots. J. Integr. Plant Biol. 2008, 50, 49–55. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, C.-Z.; Guo, C.; Wang, Y.-C.; Ouyang, F. Comparison of various bioreactors on growth and artemisinin biosynthesis of Artemisia annua L. shoot cultures. Process Biochem. 2003, 39, 45–49. [Google Scholar] [CrossRef]
  39. Salehi, M.; Karimzadeh, G.; Naghavi, M.R. Synergistic effect of coronatine and sorbitol on artemisinin production in cell suspension culture of Artemisia annua L. cv. Anamed. Plant Cell Tissue Organ Cult. 2019, 137, 587–597. [Google Scholar] [CrossRef]
  40. Santos, P.A.G.; Figueiredo, A.C.; Lourenço, P.M.L.; Barroso, J.G.; Pedro, L.G.; Oliveira, M.M.; Schripsema, J.; Deans, S.G.; Scheffer, J.J.C. Hairy root cultures of Anethum graveolens (dill): Establishment, growth, time-course study of their essential oil and its comparison with parent plant oils. Biotechnol. Lett. 2002, 24, 1031–1036. [Google Scholar] [CrossRef]
  41. Figueiredo, A.C.; Salomé, M.; Pais, S.; Scheffer, J.J.C. Composition of the essential oil from cell suspension cultures of Achillea millefolium ssp. millefolium. Plant Cell Tissue Organ Cult. 1995, 40, 113–118. [Google Scholar] [CrossRef]
  42. Lourenço, P.M.L.; Figueiredo, A.C.; Barroso, J.G.; Pedro, L.G.; Oliveira, M.M.; Deans, S.G.; Scheffer, J.J.C. Essential oils from hairy root cultures and from plant roots of Achillea millefolium. Phytochemistry 1999, 51, 637–642. [Google Scholar] [CrossRef]
  43. Pan, X.-W.; Han, L.; Zhang, Y.-H.; Chen, D.-F.; Simonsen, H.T. Sclareol production in the moss Physcomitrella patens and observations on growth and terpenoid biosynthesis. Plant Biotechnol. Rep. 2015, 9, 149–159. [Google Scholar] [CrossRef]
  44. Vaccaro, M.C.; Alfieri, M.; De Tommasi, N.; Moses, T.; Goossens, A.; Leone, A. Boosting the synthesis of pharmaceutically active abietane diterpenes in S. sclarea hairy roots by engineering the GGPPS and CPPS genes. Front. Plant Sci. 2020, 11, 924. [Google Scholar] [CrossRef] [PubMed]
  45. Bringi, V.; Kadkade, P.; Prince, C.; Roach, B. Enhanced production of paclitaxel and taxanes by cell cultures of Taxus species. U.S. Patent 8,338,143, 25 December 2012. [Google Scholar]
  46. Sykłowska-Baranek, K.; Sygitowicz, G.; Maciejak-Jastrzębska, A.; Pietrosiuk, A.; Szakiel, A. Application of Priming Strategy for Enhanced Paclitaxel Biosynthesis in Taxus× Media Hairy Root Cultures. Cells 2022, 11, 2062. [Google Scholar] [CrossRef] [PubMed]
  47. Shi, M.; Luo, X.; Ju, G.; Li, L.; Huang, S.; Zhang, T.; Wang, H.; Kai, G. Enhanced diterpene tanshinone accumulation and bioactivity of transgenic Salvia miltiorrhiza hairy roots by pathway engineering. J. Agric. Food Chem. 2016, 64, 2523–2530. [Google Scholar] [CrossRef] [PubMed]
  48. Kim, Y.-K.; Kim, Y.B.; Uddin, M.R.; Lee, S.; Kim, S.-U.; Park, S.U. Enhanced triterpene accumulation in Panax ginseng hairy roots overexpressing mevalonate-5-pyrophosphate decarboxylase and farnesyl pyrophosphate synthase. ACS Synth. Biol. 2014, 3, 773–779. [Google Scholar] [CrossRef] [PubMed]
  49. Le, K.-C.; Im, W.-T.; Paek, K.-Y.; Park, S.-Y. Biotic elicitation of ginsenoside metabolism of mutant adventitious root culture in Panax ginseng. Appl. Microbiol. Biotechnol. 2018, 102, 1687–1697. [Google Scholar] [CrossRef] [PubMed]
  50. Biswas, T.; Pandey, S.S.; Maji, D.; Gupta, V.; Kalra, A.; Singh, M.; Mathur, A.; Mathur, A. Enhanced expression of ginsenoside biosynthetic genes and in vitro ginsenoside production in elicited Panax sikkimensis (Ban) cell suspensions. Protoplasma 2018, 255, 1147–1160. [Google Scholar] [CrossRef] [PubMed]
  51. Woragidbumrung, K.-o.; Sae-Tang, P.; Yao, H.; Han, J.; Chauvatcharin, S.; Zhong, J.-J. Impact of conditioned medium on cell cultures of Panax notoginseng in an airlift bioreactor. Process Biochem. 2001, 37, 209–213. [Google Scholar] [CrossRef]
  52. Kim, Y.-K.; Kim, J.K.; Kim, Y.B.; Lee, S.; Kim, S.-U.; Park, S.U. Enhanced Accumulation of Phytosterol and Triterpene in Hairy Root Cultures of Platycodon grandiflorum by Overexpression of Panax ginseng 3-Hydroxy-3-methylglutaryl-coenzyme A Reductase. J. Agric. Food Chem. 2013, 61, 1928–1934. [Google Scholar] [CrossRef]
  53. Alsoufi, A.S.M.; Pączkowski, C.; Szakiel, A.; Długosz, M. Effect of jasmonic acid and chitosan on triterpenoid production in Calendula officinalis hairy root cultures. Phytochem. Lett. 2019, 31, 5–11. [Google Scholar] [CrossRef]
  54. Wiktorowska, E.; Długosz, M.; Janiszowska, W. Significant enhancement of oleanolic acid accumulation by biotic elicitors in cell suspension cultures of Calendula officinalis L. Enzym. Microb. Technol. 2010, 46, 14–20. [Google Scholar] [CrossRef]
  55. Zhang, B.; Chen, L.; Huo, Y.; Feng, J.; Ma, Z.; Zhang, X.; Zhu, C. Enhanced production of celastrol in Tripterygium wilfordii hairy root cultures by overexpression of TwSQS2. Biochem. Eng. J. 2020, 161, 107681. [Google Scholar] [CrossRef]
  56. Saeed, B.; Das, M.; Khurana, P. Overexpression of β-carotene hydroxylase1 (BCH1) in Indian mulberry, Morus indica cv. K2, confers tolerance against UV, high temperature and high irradiance stress induced oxidative damage. Plant Cell Tissue Organ Cult. 2015, 120, 1003–1014. [Google Scholar] [CrossRef]
  57. Oleszkiewicz, T.; Kruczek, M.; Baranski, R. Repression of carotenoid accumulation by nitrogen and NH4+ supply in carrot callus cells in vitro. Plants 2021, 10, 1813. [Google Scholar] [CrossRef] [PubMed]
  58. Huang, J.-C.; Zhong, Y.-J.; Liu, J.; Sandmann, G.; Chen, F. Metabolic engineering of tomato for high-yield production of astaxanthin. Metab. Eng. 2013, 17, 59–67. [Google Scholar] [CrossRef] [PubMed]
  59. Zhong, J.-J.; Yoshida, T. High-density cultivation of Perilla frutescens cell suspensions for anthocyanin production: Effects of sucrose concentration and inoculum size. Enzym. Microb. Technol. 1995, 17, 1073–1079. [Google Scholar] [CrossRef]
  60. Appelhagen, I.; Wulff-Vester, A.K.; Wendell, M.; Hvoslef-Eide, A.-K.; Russell, J.; Oertel, A.; Martens, S.; Mock, H.-P.; Martin, C.; Matros, A. Colour bio-factories: Towards scale-up production of anthocyanins in plant cell cultures. Metab. Eng. 2018, 48, 218–232. [Google Scholar] [CrossRef]
  61. Kobayashi, Y.; Akita, M.; Sakamoto, K.; Liu, H.; Shigeoka, T.; Koyano, T.; Kawamura, M.; Furuya, T. Large-scale production of anthocyanin by Aralia cordata cell suspension cultures. Appl. Microbiol. Biotechnol. 1993, 40, 215–218. [Google Scholar] [CrossRef]
  62. Qin, Y.; Wang, D.; Fu, J.; Zhang, Z.; Qin, Y.; Hu, G.; Zhao, J. Agrobacterium rhizogenes-mediated hairy root transformation as an efficient system for gene function analysis in Litchi chinensis. Plant Methods 2021, 17, 103. [Google Scholar] [CrossRef]
  63. Trotin, F.; Moumou, Y.; Vasseur, J. Flavanol production by Fagopyrum esculentum hairy and normal root cultures. Phytochemistry 1993, 32, 929–931. [Google Scholar] [CrossRef]
  64. Bulgakov, V.P.; Tchernoded, G.; Veselova, M.; Fedoreyev, S.; Muzarok, T.; Zhuravlev, Y. Catechin production in cultured cells of Taxus cuspidata and Taxus baccata. Biotechnol. Lett. 2011, 33, 1879–1883. [Google Scholar] [CrossRef]
  65. Yin, Q.; Han, X.; Han, Z.; Chen, Q.; Shi, Y.; Gao, H.; Zhang, T.; Dong, G.; Xiong, C.; Song, C. Genome-wide analyses reveals a glucosyltransferase involved in rutin and emodin glucoside biosynthesis in tartary buckwheat. Food Chem. 2020, 318, 126478. [Google Scholar] [CrossRef] [PubMed]
  66. Ho, T.-T.; Lee, J.-D.; Ahn, M.-S.; Kim, S.-W.; Park, S.-Y. Enhanced production of phenolic compounds in hairy root cultures of Polygonum multiflorum and its metabolite discrimination using HPLC and FT-IR methods. Appl. Microbiol. Biotechnol. 2018, 102, 9563–9575. [Google Scholar] [CrossRef] [PubMed]
  67. Thiruvengadam, M.; Rekha, K.; Rajakumar, G.; Lee, T.-J.; Kim, S.-H.; Chung, I.-M. Enhanced production of anthraquinones and phenolic compounds and biological activities in the cell suspension cultures of Polygonum multiflorum. Int. J. Mol. Sci. 2016, 17, 1912. [Google Scholar] [CrossRef] [PubMed]
  68. Ho, T.-T.; Lee, K.-J.; Lee, J.-D.; Bhushan, S.; Paek, K.-Y.; Park, S.-Y. Adventitious root culture of Polygonum multiflorum for phenolic compounds and its pilot-scale production in 500 L-tank. Plant Cell Tissue Organ Cult. 2017, 130, 167–181. [Google Scholar] [CrossRef]
  69. Ong, S.; Ling, A.; Poospooragi, R.; Moosa, S. Production of Flavonoid compounds in cell cultures of Ficus deltoidea as influenced by medium composition. Int. J. Med. Aromat. Plants 2011, 1, 62–74. [Google Scholar]
  70. Anil Kumar, M.; Sravanthi Pammi, S.; Sukanya, M.; Giri, A. Enhanced production of pharmaceutically important isoflavones from hairy root rhizoclones of Trifolium pratense L. Vitr. Cell. Dev.-Pl. 2018, 54, 94–103. [Google Scholar] [CrossRef]
  71. Shinde, A.N.; Malpathak, N.; Fulzele, D.P. Optimized production of isoflavones in cell cultures of Psoralea corylifolia L. using elicitation and precursor feeding. Biotechnol. Bioprocess Eng. 2009, 14, 612–618. [Google Scholar] [CrossRef]
  72. Yang, T.; Fang, L.; Nopo-Olazabal, C.; Condori, J.; Nopo-Olazabal, L.; Balmaceda, C.; Medina-Bolivar, F. Enhanced production of resveratrol, piceatannol, arachidin-1, and arachidin-3 in hairy root cultures of peanut co-treated with methyl jasmonate and cyclodextrin. J. Agric. Food Chem. 2015, 63, 3942–3950. [Google Scholar] [CrossRef]
  73. Donnez, D.; Kim, K.-H.; Antoine, S.; Conreux, A.; De Luca, V.; Jeandet, P.; Clément, C.; Courot, E. Bioproduction of resveratrol and viniferins by an elicited grapevine cell culture in a 2 L stirred bioreactor. Process Biochem. 2011, 46, 1056–1062. [Google Scholar] [CrossRef]
  74. Garcia, B.M.P. Method for the production of resveratrol in cell cultures. U.S. Patent 7309591B2, 18 December 2007. [Google Scholar]
  75. Su, W.; Lei, F.; Kao, N. High density cultivation of Anchusa officinalis in a stirred-tank bioreactor with in situ filtration. Appl. Microbiol. Biotechnol. 1995, 44, 293–299. [Google Scholar] [CrossRef]
  76. Wu, C.-H.; Murthy, H.N.; Hahn, E.-J.; Paek, K.-Y. Large-scale cultivation of adventitious roots of Echinacea purpurea in airlift bioreactors for the production of chichoric acid, chlorogenic acid and caftaric acid. Biotechnol. Lett. 2007, 29, 1179–1182. [Google Scholar] [CrossRef] [PubMed]
  77. Kobayashi, Y.; Fukui, H.; Tabata, M. Berberine production by batch and semi-continuous cultures of immobilized Thalictrum cells in an improved bioreactor. Plant Cell Rep. 1988, 7, 249–252. [Google Scholar] [CrossRef] [PubMed]
  78. Yukimune, Y.; Hara, Y.; Yamada, Y. Tropane alkaloid production in root cultures of Duboisia myoporoides obtained by repeated selection. Biosci. Biotechnol. Biochem. 1994, 58, 1443–1446. [Google Scholar] [CrossRef]
  79. Zhao, B.; Agblevor, F.A.; KC, R.; Jelesko, J.G. Enhanced production of the alkaloid nicotine in hairy root cultures of Nicotiana tabacum L. Plant Cell Tissue Organ Cult. 2013, 113, 121–129. [Google Scholar] [CrossRef]
  80. Mantell, S.; Pearson, D.; Hazell, L.; Smith, H. The effect of initial phosphate and sucrose levels on nicotine accumulation in batch suspension cultures of Nicotiana tabacum L. Plant Cell Rep. 1983, 2, 73–77. [Google Scholar] [CrossRef] [PubMed]
  81. Saito, K.; Sudo, H.; Yamazaki, M.; Koseki-Nakamura, M.; Kitajima, M.; Takayama, H.; Aimi, N. Feasible production of camptothecin by hairy root culture of Ophiorrhiza pumila. Plant Cell Rep. 2001, 20, 267–271. [Google Scholar] [CrossRef]
  82. Shi, M.; Gong, H.; Cui, L.; Wang, Q.; Wang, C.; Wang, Y.; Kai, G. Targeted metabolic engineering of committed steps improves anti-cancer drug camptothecin production in Ophiorrhiza pumila hairy roots. Ind. Crops Prod. 2020, 148, 112277. [Google Scholar] [CrossRef]
  83. Siah, C.L.; Doran, P.M. Enhanced codeine and morphine production in suspended Papaver somniferum cultures after removal of exogenous hormones. Plant Cell Rep. 1991, 10, 349–353. [Google Scholar] [CrossRef]
  84. Le Flem-Bonhomme, V.; Laurain-Mattar, D.; Fliniaux, M. Hairy root induction of Papaver somniferum var. album, a difficult-to-transform plant, by A. rhizogenes LBA 9402. Planta 2004, 218, 890–893. [Google Scholar] [CrossRef]
  85. Behzadirad, M.; Naghavi, M.; Shahnejat Bushehri, A. Importance of hormonal elicitors in inducing morphine biosynthesis in the cell culture of Papaver bracteatum Lindl. J. Agric. Sci. Technol. 2020, 22, 261–270. [Google Scholar]
  86. Sharafi, A.; Sohi, H.H.; Mousavi, A.; Azadi, P.; Khalifani, B.H.; Razavi, K. Metabolic engineering of morphinan alkaloids by over-expression of codeinone reductase in transgenic hairy roots of Papaver bracteatum, the Iranian poppy. Biotechnol. Lett. 2013, 35, 445–453. [Google Scholar] [CrossRef] [PubMed]
  87. Hashemi, S.M.; Naghavi, M.R. Production and gene expression of morphinan alkaloids in hairy root culture of Papaver orientale L. using abiotic elicitors. Plant Cell Tissue Organ Cult. 2016, 125, 31–41. [Google Scholar] [CrossRef]
  88. Peebles, C.A.; Hughes, E.H.; Shanks, J.V.; San, K.-Y. Transcriptional response of the terpenoid indole alkaloid pathway to the overexpression of ORCA3 along with jasmonic acid elicitation of Catharanthus roseus hairy roots over time. Metab. Eng. 2009, 11, 76–86. [Google Scholar] [CrossRef] [PubMed]
  89. Sun, J.; Peebles, C.A.M. Engineering overexpression of ORCA3 and strictosidine glucosidase in Catharanthus roseus hairy roots increases alkaloid production. Protoplasma 2016, 253, 1255–1264. [Google Scholar] [CrossRef] [PubMed]
  90. Malik, S.; Cusidó, R.M.; Mirjalili, M.H.; Moyano, E.; Palazón, J.; Bonfill, M. Production of the anticancer drug taxol in Taxus baccata suspension cultures: A review. Process Biochem. 2011, 46, 23–34. [Google Scholar] [CrossRef]
  91. Sanchez-Muñoz, R.; Moyano, E.; Khojasteh, A.; Bonfill, M.; Cusido, R.M.; Palazon, J. Genomic methylation in plant cell cultures: A barrier to the development of commercial long-term biofactories. Eng. Life Sci. 2019, 19, 872–879. [Google Scholar] [CrossRef] [PubMed]
  92. Deus-Neumann, B.; Zenk, M. Instability of indole alkaloid production in Catharanthus roseus cell suspension cultures. Planta Medica 1984, 50, 427–431. [Google Scholar] [CrossRef]
  93. Lee, E.-K.; Jin, Y.-W.; Park, J.H.; Yoo, Y.M.; Hong, S.M.; Amir, R.; Yan, Z.; Kwon, E.; Elfick, A.; Tomlinson, S.; et al. Cultured cambial meristematic cells as a source of plant natural products. Nat. Biotechnol. 2010, 28, 1213–1217. [Google Scholar] [CrossRef]
  94. Gantait, S.; Mukherjee, E. Hairy root culture technology: Applications, constraints and prospect. Appl. Microbiol. Biotechnol. 2021, 105, 35–53. [Google Scholar] [CrossRef]
  95. Paek, K.Y.; Murthy, H.N.; Hahn, E.J.; Zhong, J.J. Large scale culture of ginseng adventitious roots for production of ginsenosides. Adv. Biochem. Eng./Biotechnol. 2009, 113, 151–176. [Google Scholar] [PubMed]
  96. Muranaka, T.; Kitamura, Y.; Ikenaga, T. Genetic transformation of Duboisia species. Transgenic Medicinal Plants 1999, 117–132. [Google Scholar]
  97. Wang, J.W.; Wu, J.Y. Tanshinone biosynthesis in Salvia miltiorrhiza and production in plant tissue cultures. Appl. Microbiol. Biotechnol. 2010, 88, 437–449. [Google Scholar] [CrossRef] [PubMed]
  98. Park, S.U.; Facchini, P.J. Agrobacterium rhizogenes-mediated transformation of opium poppy, Papaver somniferum L. and California poppy, Eschscholzia californica Cham. root cultures. J. Exp. Bot. 2000, 51, 1005–1016. [Google Scholar] [CrossRef]
  99. Miao, G.P.; Zhu, C.S.; Yang, Y.Q.; Feng, M.X.; Ma, Z.Q.; Feng, J.T.; Zhang, X. Elicitation and in situ adsorption enhanced secondary metabolites production of Tripterygium wilfordii Hook. f. adventitious root fragment liquid cultures in shake flask and a modified bubble column bioreactor. Bioprocess Biosyst. Eng. 2014, 37, 641–650. [Google Scholar] [CrossRef] [PubMed]
  100. Zhu, C.; Miao, G.; Guo, J.; Huo, Y.; Zhang, X.; Xie, J.; Feng, J. Establishment of Tripterygium wilfordii Hook. f. Hairy root culture and optimization of its culture conditions for the production of triptolide and wilforine. J. Microbiol. Biotechnol. 2014, 24, 823–834. [Google Scholar] [CrossRef] [PubMed]
  101. Ruiz-May, E.; Galaz-Avalos, R.M.; Loyola-Vargas, V.M. Differential secretion and accumulation of terpene indole alkaloids in hairy roots of Catharanthus roseus treated with methyl jasmonate. Mol. Biotechnol. 2009, 41, 278–285. [Google Scholar] [CrossRef] [PubMed]
  102. Mano, Y.; Ohkawa, H.; Yamada, Y. Production of tropane alkaloids by hairy root cultures of Duboisia leichhardtii transformed by Agrobacterium rhizogenes. Plant Sci. 1989, 59, 191–201. [Google Scholar] [CrossRef]
  103. Miao, G.P.; Han, J.; Zhang, J.F.; Zhu, C.S.; Zhang, X. A MDR transporter contributes to the different extracellular production of sesquiterpene pyridine alkaloids between adventitious root and hairy root liquid cultures of Tripterygium wilfordii Hook. f. Plant Mol. Biol. 2017, 95, 51–62. [Google Scholar] [CrossRef]
  104. Zhi, B.H.; Alfermann, A. Diterpenoid production in hairy root cultures of Salvia miltiorrhiza. Phytochemistry 1993, 32, 699–703. [Google Scholar] [CrossRef]
  105. Taya, M.; Mine, K.; Kino-Oka, M.; Tone, S.; Ichi, T. Production and release of pigments by culture of transformed hairy root of red beet. J. Ferment. Bioeng. 1992, 73, 31–36. [Google Scholar] [CrossRef]
  106. Selwal, N.; Rahayu, F.; Herwati, A.; Latifah, E.; Supriyono; Suhara, C.; Kade Suastika, I.B.; Mahayu, W.M.; Wani, A.K. Enhancing secondary metabolite production in plants: Exploring traditional and modern strategies. J. Agric. Food Res. 2023, 14, 100702. [Google Scholar] [CrossRef]
  107. Halder, M.; Sarkar, S.; Jha, S. Elicitation: A biotechnological tool for enhanced production of secondary metabolites in hairy root cultures. Eng. Life Sci. 2019, 19, 880–895. [Google Scholar] [CrossRef] [PubMed]
  108. Thakur, M.; Bhattacharya, S.; Khosla, P.K.; Puri, S. Improving production of plant secondary metabolites through biotic and abiotic elicitation. J. Appl. Res. Med. Aromat. Plants 2019, 12, 1–12. [Google Scholar] [CrossRef]
  109. Sohn, S.-I.; Pandian, S.; Rakkammal, K.; Largia, M.J.V.; Thamilarasan, S.K.; Balaji, S.; Zoclanclounon, Y.A.B.; Shilpha, J.; Ramesh, M. Jasmonates in plant growth and development and elicitation of secondary metabolites: An updated overview. Front. Plant Sci. 2022, 13, 942789. [Google Scholar] [CrossRef] [PubMed]
  110. Murthy, H.N.; Lee, E.J.; Paek, K.Y. Production of secondary metabolites from cell and organ cultures: Strategies and approaches for biomass improvement and metabolite accumulation. Plant Cell Tissue Org. 2014, 118, 1–16. [Google Scholar] [CrossRef]
  111. Scarpa, G.M.; Prota, V.; Schianchi, N.; Manunta, F. In vitro cultures for the production of secondary metabolites. In Secondary Metabolites-Trends and Reviews; IntechOpen: London, UK, 2022. [Google Scholar]
  112. Weathers, P.J.; Towler, M.J.; Xu, J. Bench to batch: Advances in plant cell culture for producing useful products. Appl. Microbiol. Biotechnol. 2010, 85, 1339–1351. [Google Scholar] [CrossRef]
  113. Su, W.W. Bioreactor engineering for recombinant protein production using plant cell suspension culture. Plant Tissue Cult. Eng. 2008, 6, 135–159. [Google Scholar]
  114. Huang, T.-K.; McDonald, K.A. Bioreactor engineering for recombinant protein production in plant cell suspension cultures. Biochem. Eng. J. 2009, 45, 168–184. [Google Scholar] [CrossRef]
  115. Weathers, P.; Liu, C.; Towler, M.; Wyslouzil, B. Mist reactors: Principles, comparison of various systems, and case studies. Electron J. Integr. Biosci. 2008, 3, 29–37. [Google Scholar]
  116. Choi, S.M.; Ho Son, S.; Rho Yun, S.; Woung Kwon, O.; Hoon Seon, J.; Yoeup Paek, K. Pilot-scale culture of adventitious roots of ginseng in a bioreactor system. Plant Cell Tissue Organ Cult. 2000, 62, 187–193. [Google Scholar] [CrossRef]
  117. Choi, Y.-E.; Kim, Y.-S.; Paek, K.-Y. Types and designs of bioreactors for hairy root culture. Plant Tissue Cult. Eng. 2008, 161–172. [Google Scholar]
  118. Verma, P.; Khan, S.A.; Alhandhali, A.J.A.; Parasharami, V.A. Bioreactor upscaling of different tissue of medicinal herbs for extraction of active phytomolecules: A step towards industrialization and enhanced production of phytochemicals. In Plant Growth Regulators: Signalling under Stress Conditions; Springer: Berlin/Heidelberg, Germany, 2021; pp. 455–481. [Google Scholar] [CrossRef]
  119. Murthy, H.N.; Joseph, K.S.; Paek, K.Y.; Park, S.Y. Bioreactor systems for micropropagation of plants: Present scenario and future prospects. Front. Plant Sci. 2023, 14, 1159588. [Google Scholar] [CrossRef] [PubMed]
  120. Farré, G.; Blancquaert, D.; Capell, T.; Van Der Straeten, D.; Christou, P.; Zhu, C. Engineering complex metabolic pathways in plants. Annu. Rev. Plant Biol. 2014, 65, 187–223. [Google Scholar] [CrossRef]
  121. Sun, J.; Zhao, L.; Shao, Z.; Shanks, J.; Peebles, C.A.M. Expression of tabersonine 16-hydroxylase and 16-hydroxytabersonine-O-methyltransferase in Catharanthus roseus hairy roots. Biotechnol. Bioeng. 2018, 115, 673–683. [Google Scholar] [CrossRef] [PubMed]
  122. Allen, R.S.; Millgate, A.G.; Chitty, J.A.; Thisleton, J.; Miller, J.A.; Fist, A.J.; Gerlach, W.L.; Larkin, P.J. RNAi-mediated replacement of morphine with the nonnarcotic alkaloid reticuline in opium poppy. Nat. Biotechnol. 2004, 22, 1559–1566. [Google Scholar] [CrossRef] [PubMed]
  123. Deng, C.; Wu, Y.; Lv, X.; Li, J.; Liu, Y.; Du, G.; Chen, J.; Liu, L. Refactoring transcription factors for metabolic engineering. Biotechnol. Adv. 2022, 57, 107935. [Google Scholar] [CrossRef] [PubMed]
  124. Mora-Vásquez, S.; Wells-Abascal, G.G.; Espinosa-Leal, C.; Cardineau, G.A.; García-Lara, S. Application of metabolic engineering to enhance the content of alkaloids in medicinal plants. Metab. Eng. Commun. 2022, 14, e00194. [Google Scholar] [CrossRef] [PubMed]
  125. van der Hoek, S.A.; Borodina, I. Transporter engineering in microbial cell factories: The ins, the outs, and the in-betweens. Curr. Opin. Biotechnol. 2020, 66, 186–194. [Google Scholar] [CrossRef]
  126. Yue, W.; Ming, Q.-l.; Lin, B.; Rahman, K.; Zheng, C.-J.; Han, T.; Qin, L.-p. Medicinal plant cell suspension cultures: Pharmaceutical applications and high-yielding strategies for the desired secondary metabolites. Crit. Rev. Biotechnol. 2016, 36, 215–232. [Google Scholar] [CrossRef]
  127. Wu, X.; Liu, J.; Liu, Z.; Gong, G.; Zha, J. Microbial cell surface engineering for high-level synthesis of bio-products. Biotechnol. Adv. 2022, 55, 107912. [Google Scholar] [CrossRef]
  128. Ahmed, M.S.; Lauersen, K.J.; Ikram, S.; Li, C. Efflux transporters’ engineering and their application in microbial production of heterologous metabolites. ACS Synth. Biol. 2021, 10, 646–669. [Google Scholar] [CrossRef] [PubMed]
  129. Nogia, P.; Pati, P.K. Plant secondary metabolite transporters: Diversity, functionality, and their modulation. Front. Plant Sci. 2021, 12, 758202. [Google Scholar] [CrossRef] [PubMed]
  130. Lefèvre, F.; Boutry, M. Towards identification of the substrates of ATP-binding cassette transporters. Plant Physiol. 2018, 178, 18–39. [Google Scholar] [CrossRef] [PubMed]
  131. Upadhyay, N.; Kar, D.; Deepak Mahajan, B.; Nanda, S.; Rahiman, R.; Panchakshari, N.; Bhagavatula, L.; Datta, S. The multitasking abilities of MATE transporters in plants. J. Exp. Bot. 2019, 70, 4643–4656. [Google Scholar] [CrossRef] [PubMed]
  132. Shoji, T. ATP-binding cassette and multidrug and toxic compound extrusion transporters in plants: A common theme among diverse detoxification mechanisms. Int. Rev. Cell Mol. Biol. 2014, 309, 303–346. [Google Scholar] [PubMed]
  133. Gani, U.; Vishwakarma, R.A.; Misra, P. Membrane transporters: The key drivers of transport of secondary metabolites in plants. Plant Cell Rep. 2021, 40, 1–18. [Google Scholar] [CrossRef]
  134. Dahuja, A.; Kumar, R.R.; Sakhare, A.; Watts, A.; Singh, B.; Goswami, S.; Sachdev, A.; Praveen, S. Role of ATP-binding cassette transporters in maintaining plant homeostasis under abiotic and biotic stresses. Physiol Plant. 2021, 171, 785–801. [Google Scholar] [CrossRef]
  135. Pomahačová, B.; Dušek, J.; Dušková, J.; Yazaki, K.; Roytrakul, S.; Verpoorte, R. Improved accumulation of ajmalicine and tetrahydroalstonine in Catharanthus cells expressing an ABC transporter. J. Plant Physiol. 2009, 166, 1405–1412. [Google Scholar] [CrossRef]
  136. Miao, G.; Han, J.; Huo, Y.B.; Wang, C.R.; Wang, S.C. Identification and functional characterization of a PDR transporter in Tripterygium wilfordii Hook. f. that mediates the efflux of triptolide. Plant Mol. Biol. 2021, 106, 145–156. [Google Scholar] [CrossRef]
  137. Aryal, B.; Laurent, C.; Geisler, M. Learning from each other: ABC transporter regulation by protein phosphorylation in plant and mammalian systems. Biochem. Soc. Trans. 2015, 43, 966–974. [Google Scholar] [CrossRef]
  138. Han, J.; Liu, C.-X.; Liu, J.; Wang, C.-R.; Wang, S.-C.; Miao, G. AGC kinases OXI1 and AGC2-2 regulate camalexin secretion and disease resistance by phosphorylating transporter PDR6. Plant Physiol. 2024, kiae186. [Google Scholar] [CrossRef]
  139. Zeevi, V.; Liang, Z.; Arieli, U.; Tzfira, T. Zinc finger nuclease and homing endonuclease-mediated assembly of multigene plant transformation vectors. Plant Physiol. 2012, 158, 132–144. [Google Scholar] [CrossRef]
  140. Kok, S.D.; Stanton, L.H.; Slaby, T.; Durot, M.; Holmes, V.F.; Patel, K.G.; Platt, D.; Shapland, E.B.; Serber, Z.; Dean, J.; et al. Rapid and reliable dna assembly via ligase cycling reaction. ACS Synth. Biol. 2014, 3, 97–106. [Google Scholar] [CrossRef]
  141. Engler, C.; Kandzia, R.; Marillonnet, S. A one pot, one step, precision cloning method with high throughput capability. PloS ONE 2008, 3, e3647. [Google Scholar] [CrossRef]
  142. Lin, L.; Liu, Y.-G.; Xu, X.; Li, B. Efficient linking and transfer of multiple genes by a multigene assembly and transformation vector system. Proc. Natl. Acad. Sci. USA 2003, 100, 5962–5967. [Google Scholar] [CrossRef] [PubMed]
  143. Buntru, M.; Gärtner, S.; Staib, L.; Kreuzaler, F.; Schlaich, N. Delivery of multiple transgenes to plant cells by an improved version of MultiRound Gateway technology. Transgenic Res. 2013, 22, 153–167. [Google Scholar] [CrossRef] [PubMed]
  144. Gibson, D.G.; Young, L.; Chuang, R.-Y.; Venter, J.C.; Hutchison III, C.A.; Smith, H.O. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 2009, 6, 343–345. [Google Scholar] [CrossRef]
  145. Gibson, D.G.; Benders, G.A.; Axelrod, K.C.; Zaveri, J.; Algire, M.A.; Moodie, M.; Montague, M.G.; Venter, J.C.; Smith, H.O.; Hutchison, C.A. One-step assembly in yeast of 25 overlapping DNA fragments to form a complete synthetic Mycoplasma genitalium genome. Proc. Natl. Acad. Sci. USA 2008, 105, 20404–20409. [Google Scholar] [CrossRef] [PubMed]
  146. Zorrilla López, U.; Masip Vilà, G.; Arjó Pont, G.; Bai, C.; Banakar, R.; Bassie Rene, L.; Berman Quintana, J.; Farré Martinez, G.; Miralpeix i Anglada, B.; Pérez Massot, E. Engineering metabolic pathways in plants by multigene transformation. Int. J. Dev. Biol. 2013, 57, 565–576. [Google Scholar] [CrossRef]
  147. Zhu, Q.; Wang, B.; Tan, J.; Liu, T.; Li, L.; Liu, Y.-G. Plant Synthetic Metabolic Engineering for Enhancing Crop Nutritional Quality. Plant Commun. 2020, 1, 100017. [Google Scholar] [CrossRef]
  148. Rizzo, P.; Chavez, B.G.; Leite Dias, S.; D’Auria, J.C. Plant synthetic biology: From inspiration to augmentation. Curr. Opin. Biotechnol. 2023, 79, 102857. [Google Scholar] [CrossRef] [PubMed]
  149. Azizi-Dargahlou, S.; Pouresmaeil, M. Agrobacterium tumefaciens-mediated plant transformation: A Review. Mol. Biotechnol. 2023, 1–18. [Google Scholar] [CrossRef]
  150. Gomide, M.d.S.; Leitão, M.d.C.; Coelho, C.M. Biocircuits in plants and eukaryotic algae. Front. Plant Sci. 2022, 13, 982959. [Google Scholar] [CrossRef]
  151. Matzke, A.J.; Matzke, M.A. Position effects and epigenetic silencing of plant transgenes. Curr. Opin. Plant Biol. 1998, 1, 142–148. [Google Scholar] [CrossRef]
  152. Birchler, J.A.; Swyers, N.C. Engineered minichromosomes in plants. Exp. Cell Res. 2020, 388, 111852. [Google Scholar] [CrossRef] [PubMed]
  153. Kan, M.; Huang, T.; Zhao, P. Artificial chromosome technology and its potential application in plants. Front. Plant Sci. 2022, 13, 970943. [Google Scholar] [CrossRef]
  154. Kwan, B.D.; Seligmann, B.; Nguyen, T.D.; Franke, J.; Dang, T.T. Leveraging synthetic biology and metabolic engineering to overcome obstacles in plant pathway elucidation. Curr. Opin. Plant Biol. 2023, 71, 102330. [Google Scholar] [CrossRef] [PubMed]
  155. Dugé de Bernonville, T.; Papon, N.; Clastre, M.; O’Connor, S.E.; Courdavault, V. Identifying missing biosynthesis enzymes of plant natural products. Trends Pharmacol. Sci. 2020, 41, 142–146. [Google Scholar] [CrossRef]
  156. Caputi, L.; Franke, J.; Farrow, S.C.; Chung, K.; Payne, R.M.E.; Nguyen, T.-D.; Dang, T.-T.T.; Soares Teto Carqueijeiro, I.; Koudounas, K.; Dugé de Bernonville, T.; et al. Missing enzymes in the biosynthesis of the anticancer drug vinblastine in Madagascar periwinkle. Science 2018, 360, 1235–1239. [Google Scholar] [CrossRef]
  157. De La Peña, R.; Sattely, E.S. Rerouting plant terpene biosynthesis enables momilactone pathway elucidation. Nat. Chem. Biol. 2021, 17, 205–212. [Google Scholar] [CrossRef]
  158. Liu, W.; Xu, X.; Zhang, R.; Cheng, T.; Cao, Y.; Li, X.; Guo, J.; Liu, H.; Xian, M. Engineering Escherichia coli for high-yield geraniol production with biotransformation of geranyl acetate to geraniol under fed-batch culture. Biotechnol. Biofuels 2016, 9, 58. [Google Scholar]
  159. Jiang, G.-Z.; Yao, M.-D.; Wang, Y.; Zhou, L.; Song, T.-Q.; Liu, H.; Xiao, W.-H.; Yuan, Y.-J. Manipulation of GES and ERG20 for geraniol overproduction in Saccharomyces cerevisiae. Metab. Eng. 2017, 41, 57–66. [Google Scholar] [CrossRef]
  160. Zhao, J.; Li, C.; Zhang, Y.; Shen, Y.; Hou, J.; Bao, X. Dynamic control of ERG20 expression combined with minimized endogenous downstream metabolism contributes to the improvement of geraniol production in Saccharomyces cerevisiae. Microb. Cell Factories 2017, 16, 17. [Google Scholar] [CrossRef]
  161. Willrodt, C.; David, C.; Cornelissen, S.; Bühler, B.; Julsing, M.K.; Schmid, A. Engineering the productivity of recombinantEscherichia colifor limonene formation from glycerol in minimal media. Biotechnol. J. 2014, 9, 1000–1012. [Google Scholar] [CrossRef]
  162. Cheng, S.; Liu, X.; Jiang, G.; Wu, J.; Zhang, J.-l.; Lei, D.; Yuan, Y.-J.; Qiao, J.; Zhao, G.-R. Orthogonal engineering of biosynthetic pathway for efficient production of limonene in Saccharomyces cerevisiae. ACS Synth. Biol. 2019, 8, 968–975. [Google Scholar] [CrossRef] [PubMed]
  163. Yang, J.; Nie, Q.; Ren, M.; Feng, H.; Jiang, X.; Zheng, Y.; Liu, M.; Zhang, H.; Xian, M. Metabolic engineering of Escherichia coli for the biosynthesis of alpha-pinene. Biotechnol. Biofuels 2013, 6, 60. [Google Scholar] [CrossRef] [PubMed]
  164. Zhou, P.; Du, Y.; Xu, N.; Yue, C.; Ye, L. Improved linalool production in Saccharomyces cerevisiae by combining directed evolution of linalool synthase and overexpression of the complete mevalonate pathway. Biochem. Eng. J. 2020, 161, 107655. [Google Scholar] [CrossRef]
  165. Ma, B.; Liu, M.; Li, Z.-H.; Tao, X.; Wei, D.-Z.; Wang, F.-Q. Significantly enhanced production of patchoulol in metabolically engineered Saccharomyces cerevisiae. J. Agric. Food Chem. 2019, 67, 8590–8598. [Google Scholar] [CrossRef] [PubMed]
  166. Mitsui, R.; Nishikawa, R.; Yamada, R.; Matsumoto, T.; Ogino, H. Construction of yeast producing patchoulol by global metabolic engineering strategy. Biotechnol. Bioeng. 2020, 117, 1348–1356. [Google Scholar] [CrossRef]
  167. Paddon, C.J.; Westfall, P.J.; Pitera, D.J.; Benjamin, K.; Fisher, K.; McPhee, D.; Leavell, M.D.; Tai, A.; Main, A.; Eng, D.; et al. High-level semi-synthetic production of the potent antimalarial artemisinin. Nature 2013, 496, 528–532. [Google Scholar] [CrossRef]
  168. Lenihan, J.R.; Tsuruta, H.; Diola, D.; Renninger, N.S.; Regentin, R. Developing an industrial artemisinic acid fermentation process to support the cost-effective production of antimalarial artemisinin-based combination therapies. Biotechnol. Prog. 2008, 24, 1026–1032. [Google Scholar] [CrossRef]
  169. Cao, X.; Yu, W.; Chen, Y.; Yang, S.; Zhao, Z.K.; Nielsen, J.; Luan, H.; Zhou, Y.J. Engineering yeast for high-level production of diterpenoid sclareol. Metab. Eng. 2023, 75, 19–28. [Google Scholar] [CrossRef] [PubMed]
  170. Schalk, M.; Pastore, L.; Mirata, M.A.; Khim, S.; Schouwey, M.; Deguerry, F.; Pineda, V.; Rocci, L.; Daviet, L. Toward a biosynthetic route to sclareol and amber odorants. J. Am. Chem. Soc. 2012, 134, 18900–18903. [Google Scholar] [CrossRef] [PubMed]
  171. Ajikumar, P.K.; Xiao, W.-H.; Tyo, K.E.J.; Wang, Y.; Simeon, F.; Leonard, E.; Mucha, O.; Phon, T.H.; Pfeifer, B.; Stephanopoulos, G. Isoprenoid pathway optimization for taxol precursor overproduction in Escherichia coli. Science 2010, 330, 70–74. [Google Scholar] [CrossRef] [PubMed]
  172. Zhang, C.; Chen, W.; Dong, T.; Wang, Y.; Yao, M.; Xiao, W.; Li, B. Elimination of enzymes catalysis compartmentalization enhancing taxadiene production in Saccharomyces cerevisiae. Front. Bioeng. Biotechnol. 2023, 11, 1141272. [Google Scholar] [CrossRef] [PubMed]
  173. Hu, T.; Zhou, J.; Tong, Y.; Su, P.; Li, X.; Liu, Y.; Liu, N.; Wu, X.; Zhang, Y.; Wang, J.; et al. Engineering chimeric diterpene synthases and isoprenoid biosynthetic pathways enables high-level production of miltiradiene in yeast. Metab. Eng. 2020, 60, 87–96. [Google Scholar] [CrossRef] [PubMed]
  174. Guo, J.; Zhou, Y.J.; Hillwig, M.L.; Shen, Y.; Yang, L.; Wang, Y.; Zhang, X.; Liu, W.; Peters, R.J.; Chen, X. CYP76AH1 catalyzes turnover of miltiradiene in tanshinones biosynthesis and enables heterologous production of ferruginol in yeasts. Proc. Natl. Acad. Sci. USA 2013, 110, 12108–12113. [Google Scholar] [CrossRef] [PubMed]
  175. Zhuang, Y.; Yang, G.-Y.; Chen, X.; Liu, Q.; Zhang, X.; Deng, Z.; Feng, Y. Biosynthesis of plant-derived ginsenoside Rh2 in yeast via repurposing a key promiscuous microbial enzyme. Metab. Eng. 2017, 42, 25–32. [Google Scholar] [CrossRef] [PubMed]
  176. Wang, P.; Wei, W.; Ye, W.; Li, X.; Zhao, W.; Yang, C.; Li, C.; Yan, X.; Zhou, Z. Synthesizing ginsenoside Rh2 in Saccharomyces cerevisiae cell factory at high-efficiency. Cell Discov. 2019, 5, 5. [Google Scholar] [CrossRef]
  177. Zhu, M.; Wang, C.; Sun, W.; Zhou, A.; Wang, Y.; Zhang, G.; Zhou, X.; Huo, Y.; Li, C. Boosting 11-oxo-β-amyrin and glycyrrhetinic acid synthesis in Saccharomyces cerevisiae via pairing novel oxidation and reduction system from legume plants. Metab. Eng. 2018, 45, 43–50. [Google Scholar] [CrossRef]
  178. Yu, Y.; Rasool, A.; Liu, H.; Lv, B.; Chang, P.; Song, H.; Wang, Y.; Li, C. Engineering Saccharomyces cerevisiae for high yield production of α-amyrin via synergistic remodeling of α-amyrin synthase and expanding the storage pool. Metab. Eng. 2020, 62, 72–83. [Google Scholar] [CrossRef]
  179. Zhao, Y.; Fan, J.; Wang, C.; Feng, X.; Li, C. Enhancing oleanolic acid production in engineered Saccharomyces cerevisiae. Bioresour. Technol. 2018, 257, 339–343. [Google Scholar] [CrossRef] [PubMed]
  180. Zhao, Y.; Hansen, N.L.; Duan, Y.-T.; Prasad, M.; Motawia, M.S.; Møller, B.L.; Pateraki, I.; Staerk, D.; Bak, S.; Miettinen, K.; et al. Biosynthesis and biotechnological production of the anti-obesity agent celastrol. Nat. Chem. 2023, 15, 1236–1246. [Google Scholar] [CrossRef] [PubMed]
  181. Shi, B.; Ma, T.; Ye, Z.; Li, X.; Huang, Y.; Zhou, Z.; Ding, Y.; Deng, Z.; Liu, T. Systematic metabolic engineering of Saccharomyces cerevisiae for lycopene overproduction. J. Agric. Food Chem. 2019, 67, 11148–11157. [Google Scholar] [CrossRef] [PubMed]
  182. Sun, T.; Miao, L.; Li, Q.; Dai, G.; Lu, F.; Liu, T.; Zhang, X.; Ma, Y. Production of lycopene by metabolically-engineered Escherichia coli. Biotechnol. Lett. 2014, 36, 1515–1522. [Google Scholar] [CrossRef] [PubMed]
  183. Yan, G.-l.; Wen, K.-r.; Duan, C.-q. Enhancement of β-carotene production by over-expression of hmg-coa reductase coupled with addition of ergosterol biosynthesis inhibitors in recombinant Saccharomyces cerevisiae. Curr. Microbiol. 2011, 64, 159–163. [Google Scholar] [CrossRef] [PubMed]
  184. Xie, W.; Liu, M.; Lv, X.; Lu, W.; Gu, J.; Yu, H. Construction of a controllable β-carotene biosynthetic pathway by decentralized assembly strategy in Saccharomyces cerevisiae. Biotechnol. Bioeng. 2014, 111, 125–133. [Google Scholar] [CrossRef]
  185. Zhou, P.; Xie, W.; Yao, Z.; Zhu, Y.; Ye, L.; Yu, H. Development of a temperature-responsive yeast cell factory using engineered Gal4 as a protein switch. Biotechnol. Bioeng. 2018, 115, 1321–1330. [Google Scholar] [CrossRef] [PubMed]
  186. Jiang, G.; Yang, Z.; Wang, Y.; Yao, M.; Chen, Y.; Xiao, W.; Yuan, Y. Enhanced astaxanthin production in yeast via combined mutagenesis and evolution. Biochem. Eng. J. 2020, 156, 107519. [Google Scholar] [CrossRef]
  187. Kim, B.-G. Biological synthesis of genistein in Escherichia coli. J. Microbiol. Biotechnol. 2020, 30, 770–776. [Google Scholar] [CrossRef]
  188. Meng, Y.; Liu, X.; Zhang, L.; Zhao, G.-R. Modular Engineering of Saccharomyces cerevisiae for de novo biosynthesis of genistein. Microorganisms 2022, 10, 1402. [Google Scholar] [CrossRef]
  189. Liu, Q.; Liu, Y.; Li, G.; Savolainen, O.; Chen, Y.; Nielsen, J. De novo biosynthesis of bioactive isoflavonoids by engineered yeast cell factories. Nat. Commun. 2021, 12, 6085. [Google Scholar] [CrossRef] [PubMed]
  190. Zhao, Y.; Wu, B.-H.; Liu, Z.-N.; Qiao, J.; Zhao, G.-R. Combinatorial optimization of resveratrol production in engineered E. coli. J. Agric. Food Chem. 2018, 66, 13444–13453. [Google Scholar] [CrossRef] [PubMed]
  191. Meng, L.; Diao, M.; Wang, Q.; Peng, L.; Li, J.; Xie, N. Efficient biosynthesis of resveratrol via combining phenylalanine and tyrosine pathways in Saccharomyces cerevisiae. Microb. Cell Factories 2023, 22, 46. [Google Scholar] [CrossRef] [PubMed]
  192. Du, Y.; Yang, B.; Yi, Z.; Hu, L.; Li, M. Engineering Saccharomyces cerevisiae coculture platform for the production of flavonoids. J. Agric. Food Chem. 2020, 68, 2146–2154. [Google Scholar] [CrossRef] [PubMed]
  193. Tartik, M.; Liu, J.; Mohedano, M.T.; Mao, J.; Chen, Y. Optimizing yeast for high-level production of kaempferol and quercetin. Microb. Cell Factories 2023, 22, 74. [Google Scholar] [CrossRef] [PubMed]
  194. Eichenberger, M.; Hansson, A.; Fischer, D.; Dürr, L.; Naesby, M. De novo biosynthesis of anthocyanins in Saccharomyces cerevisiae. FEMS Yeast Res. 2018, 18, foy046. [Google Scholar] [CrossRef]
  195. Lim, C.G.; Wong, L.; Bhan, N.; Dvora, H.; Xu, P.; Venkiteswaran, S.; Koffas, M.A.G.; Kelly, R.M. Development of a recombinant Escherichia coli strain for overproduction of the plant pigment anthocyanin. Appl. Environ. Microbiol. 2015, 81, 6276–6284. [Google Scholar] [CrossRef] [PubMed]
  196. Zhao, S.; Jones, J.A.; Lachance, D.M.; Bhan, N.; Khalidi, O.; Venkataraman, S.; Wang, Z.; Koffas, M.A.G. Improvement of catechin production in Escherichia coli through combinatorial metabolic engineering. Metab. Eng. 2015, 28, 43–53. [Google Scholar] [CrossRef] [PubMed]
  197. Wang, S.; Wang, R.; Liu, T.; Zhan, Z.; Kang, L.; Wang, Y.; Lv, C.; Werck-Reichhart, D.; Guo, L.; Huang, L. Production of 3-geranyl-4-hydroxybenzoate acid in yeast, an important intermediate of shikonin biosynthesis pathway. FEMS Yeast Res. 2017, 17, fox065. [Google Scholar] [CrossRef]
  198. Rodriguez, A.; Kildegaard, K.R.; Li, M.; Borodina, I.; Nielsen, J. Establishment of a yeast platform strain for production of p-coumaric acid through metabolic engineering of aromatic amino acid biosynthesis. Metab. Eng. 2015, 31, 181–188. [Google Scholar] [CrossRef]
  199. Li, Z.; Wang, X.; Zhang, H. Balancing the non-linear rosmarinic acid biosynthetic pathway by modular co-culture engineering. Metab. Eng. 2019, 54, 1–11. [Google Scholar] [CrossRef] [PubMed]
  200. Srinivasan, P.; Smolke, C.D. Engineering cellular metabolite transport for biosynthesis of computationally predicted tropane alkaloid derivatives in yeast. Proc. Natl. Acad. Sci. USA 2021, 118, e2104460118. [Google Scholar] [CrossRef]
  201. Urui, M.; Yamada, Y.; Ikeda, Y.; Nakagawa, A.; Sato, F.; Minami, H.; Shitan, N. Establishment of a co-culture system using Escherichia coli and Pichia pastoris (Komagataella phaffii) for valuable alkaloid production. Microb. Cell Factories 2021, 20, 200. [Google Scholar] [CrossRef] [PubMed]
  202. Guo, D.; Kong, S.; Sun, Y.; Li, X.; Pan, H. Development of an artificial biosynthetic pathway for biosynthesis of (S)-reticuline based on HpaBC in engineered Escherichia coli. Biotechnol. Bioeng. 2021, 118, 4635–4642. [Google Scholar] [CrossRef] [PubMed]
  203. Pyne, M.E.; Kevvai, K.; Grewal, P.S.; Narcross, L.; Choi, B.; Bourgeois, L.; Dueber, J.E.; Martin, V.J.J. A yeast platform for high-level synthesis of tetrahydroisoquinoline alkaloids. Nat. Commun. 2020, 11, 3337. [Google Scholar] [CrossRef] [PubMed]
  204. Han, J.; Li, S. De novo biosynthesis of berberine and halogenated benzylisoquinoline alkaloids in Saccharomyces cerevisiae. Commun. Chem. 2023, 6, 27. [Google Scholar] [CrossRef]
  205. Zhang, J.; Hansen, L.G.; Gudich, O.; Viehrig, K.; Lassen, L.M.M.; Schrübbers, L.; Adhikari, K.B.; Rubaszka, P.; Carrasquer-Alvarez, E.; Chen, L.; et al. A microbial supply chain for production of the anti-cancer drug vinblastine. Nature 2022, 609, 341–347. [Google Scholar] [CrossRef] [PubMed]
  206. Gao, J.; Zuo, Y.; Xiao, F.; Wang, Y.; Li, D.; Xu, J.; Ye, C.; Feng, L.; Jiang, L.; Liu, T.; et al. Biosynthesis of catharanthine in engineered Pichia pastoris. Nat. Synth. 2023, 2, 231–242. [Google Scholar] [CrossRef]
  207. Liu, T.; Gou, Y.; Zhang, B.; Gao, R.; Dong, C.; Qi, M.; Jiang, L.; Ding, X.; Li, C.; Lian, J. Construction of ajmalicine and sanguinarine de novo biosynthetic pathways using stable integration sites in yeast. Biotechnol. Bioeng. 2022, 119, 1314–1326. [Google Scholar] [CrossRef]
  208. Guazzaroni, M.-E.; Silva-Rocha, R.; Ward, R.J. Synthetic biology approaches to improve biocatalyst identification in metagenomic library screening. Microb. Biotechnol. 2015, 8, 52–64. [Google Scholar] [CrossRef]
  209. Clomburg, J.M.; Qian, S.; Tan, Z.; Cheong, S.; Gonzalez, R. The isoprenoid alcohol pathway, a synthetic route for isoprenoid biosynthesis. Proc. Natl. Acad. Sci. USA 2019, 116, 12810–12815. [Google Scholar] [CrossRef] [PubMed]
  210. Ziemert, N.; Alanjary, M.; Weber, T. The evolution of genome mining in microbes—A review. Nat. Prod. Rep. 2016, 33, 988–1005. [Google Scholar] [CrossRef] [PubMed]
  211. Hirschi, S.; Ward, T.R.; Meier, W.P.; Müller, D.J.; Fotiadis, D. Synthetic biology: Bottom-up assembly of molecular systems. Chem. Rev. 2022, 122, 16294–16328. [Google Scholar] [CrossRef] [PubMed]
  212. Ignea, C.; Pontini, M.; Maffei, M.E.; Makris, A.M.; Kampranis, S.C. Engineering monoterpene production in yeast using a synthetic dominant negative geranyl diphosphate synthase. ACS Synth. Biol. 2014, 3, 298–306. [Google Scholar] [CrossRef] [PubMed]
  213. Ignea, C.; Trikka, F.A.; Nikolaidis, A.K.; Georgantea, P.; Ioannou, E.; Loupassaki, S.; Kefalas, P.; Kanellis, A.K.; Roussis, V.; Makris, A.M. Efficient diterpene production in yeast by engineering Erg20p into a geranylgeranyl diphosphate synthase. Metab. Eng. 2015, 27, 65–75. [Google Scholar] [CrossRef] [PubMed]
  214. Renault, H.; Bassard, J.-E.; Hamberger, B.; Werck-Reichhart, D. Cytochrome P450-mediated metabolic engineering: Current progress and future challenges. Curr. Opin. Plant Biol. 2014, 19, 27–34. [Google Scholar] [CrossRef] [PubMed]
  215. Liu, J.; Li, H.; Zhao, G.; Caiyin, Q.; Qiao, J. Redox cofactor engineering in industrial microorganisms: Strategies, recent applications and future directions. J. Ind. Microbiol. Biotechnol. 2018, 45, 313–327. [Google Scholar] [CrossRef] [PubMed]
  216. Yu, T.; Zhou, Y.J.; Huang, M.; Liu, Q.; Pereira, R.; David, F.; Nielsen, J. Reprogramming yeast metabolism from alcoholic fermentation to lipogenesis. Cell 2018, 174, 1549–1558. [Google Scholar] [CrossRef] [PubMed]
  217. Nielsen, J.R.; Weusthuis, R.A.; Huang, W.E. Growth-coupled enzyme engineering through manipulation of redox cofactor regeneration. Biotechnol Adv. 2023, 63, 108102. [Google Scholar] [CrossRef]
  218. Michener, J.K.; Nielsen, J.; Smolke, C.D. Identification and treatment of heme depletion attributed to overexpression of a lineage of evolved P450 monooxygenases. Proc. Natl. Acad. Sci. USA 2012, 109, 19504–19509. [Google Scholar] [CrossRef]
  219. Madhavan, A.; Arun, K.B.; Sindhu, R.; Binod, P.; Kim, S.H.; Pandey, A. Tailoring of microbes for the production of high value plant-derived compounds: From pathway engineering to fermentative production. Biochim. Biophys. Acta (BBA)-Proteins Proteom. 2019, 1867, 140262. [Google Scholar] [CrossRef] [PubMed]
  220. Malcı, K.; Walls, L.E.; Rios-Solis, L. Multiplex genome engineering methods for yeast cell factory development. Front. Bioeng. Biotechnol. 2020, 8, 1264. [Google Scholar] [CrossRef] [PubMed]
  221. Auxillos, J.Y.; Garcia-Ruiz, E.; Jones, S.; Li, T.; Jiang, S.; Dai, J.; Cai, Y. Multiplex Genome engineering for optimizing bioproduction in Saccharomyces cerevisiae. Biochemistry 2019, 58, 1492–1500. [Google Scholar] [CrossRef] [PubMed]
  222. Xie, W.; Ye, L.; Lv, X.; Xu, H.; Yu, H. Sequential control of biosynthetic pathways for balanced utilization of metabolic intermediates in Saccharomyces cerevisiae. Metab. Eng. 2015, 28, 8–18. [Google Scholar] [CrossRef]
  223. Peng, B.; Wood, R.J.; Nielsen, L.K.; Vickers, C.E. An Expanded Heterologous GAL Promoter collection for diauxie-inducible expression in Saccharomyces cerevisiae. ACS Synth. Biol. 2018, 7, 748–751. [Google Scholar] [CrossRef] [PubMed]
  224. Horinouchi, S. Combinatorial biosynthesis of plant medicinal polyketides by microorganisms. Curr. Opin. Chem. Biol. 2009, 13, 197–204. [Google Scholar] [CrossRef] [PubMed]
  225. Belew, Z.M.; Poborsky, M.; Nour-Eldin, H.H.; Halkier, B.A. Transport engineering in microbial cell factories producing plant-specialized metabolites. Curr. Opin. Green Sustain. Chemistry 2022, 33, 100576. [Google Scholar] [CrossRef]
  226. Yamada, Y.; Urui, M.; Oki, H.; Inoue, K.; Matsui, H.; Ikeda, Y.; Nakagawa, A.; Sato, F.; Minami, H.; Shitan, N. Transport engineering for improving the production and secretion of valuable alkaloids in Escherichia coli. Metab. Eng. Commun. 2021, 13, e00184. [Google Scholar] [CrossRef] [PubMed]
  227. Dastmalchi, M.; Chang, L.; Chen, R.; Yu, L.; Chen, X.; Hagel, J.M.; Facchini, P.J. Purine permease-type benzylisoquinoline alkaloid transporters in opium poppy. Plant Physiol. 2019, 181, 916–933. [Google Scholar] [CrossRef]
  228. Fossati, E.; Narcross, L.; Ekins, A.; Falgueyret, J.-P.; Martin, V.J.J. Synthesis of morphinan alkaloids in Saccharomyces cerevisiae. PLoS ONE 2015, 10, e0124459. [Google Scholar] [CrossRef]
  229. Srinivasan, P.; Smolke, C.D. Biosynthesis of medicinal tropane alkaloids in yeast. Nature 2020, 585, 614–619. [Google Scholar] [CrossRef] [PubMed]
  230. Khanijou, J.K.; Kulyk, H.; Bergès, C.; Khoo, L.W.; Ng, P.; Yeo, H.C.; Helmy, M.; Bellvert, F.; Chew, W.; Selvarajoo, K. Metabolomics and modelling approaches for systems metabolic engineering. Metab. Eng. Commun. 2022, 15, e00209. [Google Scholar] [CrossRef] [PubMed]
  231. Lawson, C.E.; Martí, J.M.; Radivojevic, T.; Jonnalagadda, S.V.R.; Gentz, R.; Hillson, N.J.; Peisert, S.; Kim, J.; Simmons, B.A.; Petzold, C.J.; et al. Machine learning for metabolic engineering: A review. Metab. Eng. 2021, 63, 34–60. [Google Scholar] [CrossRef]
  232. Helmy, M.; Smith, D.; Selvarajoo, K. Systems biology approaches integrated with artificial intelligence for optimized metabolic engineering. Metab. Eng. Commun. 2020, 11, e00149. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The biosynthetic pathway of terpenoids (A), alkaloids (B), and phenylpropanoids (C) in plants. The orange shaded area represents endogenous metabolic pathways present in microorganisms. The catalytic enzymes are shown in blue, intermediates in black. Solid arrow represents one step reaction, and dashed arrow represents multiple steps. The structures of typical natural products with underlines are displayed. Green circles indicate compounds for which the synthesis pathways have not been fully elucidated. Abbreviations: DXS,1-deoxy-D-xylulose 5-phosphate synthase; DXR, 1deoxy-D-xylulose 5-phosphate reductoisomerase; MCT, 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase; CMK,4-(cytidine 5‘diphospho)-2-Cmethyl-D-erythritol kinase; MECPS:ME-CDP synthase; HDS, 4-hydroxy-3-methylbut-2-enyl-diphosphate synthase; IDS: isoprenyl diphosphate synthase; IDI, isopentenyl diphosphate isomerase; GPP, geranyl diphosphate; GPPS, GPP synthase; GGPP, geranylgeranyl diphosphate; GGPPS, GGPP synthase; TPS: terpene synthase; AACT, acetyl-CoA acetyltransferase; HMGS, hydroxymethylglutaryl-CoA synthase; HMGR, 3-hydroxy-3-methylglutaryl-CoA reductase; MVK, mevalonate kinase; PMK, phosphomevalonate kinase; MPDC: mevalonate diphosphate decarboxylase; FPP, farnesyl diphosphate; FPPS, FPP synthase; STR: strictosidine synthase; PAL: phenylalanine ammonia lyase; C4H, cinnamate 4-hydroxylase; 4CL, 4-coumarate: CoA ligase; CHS, chalcone synthese; CHI, chalcone isomerase; F3H, flavanone 3-hydroxylase; F3′H, flavonoid-3′-hydroxylase; F5′H, flavonoid-5′-hydroxylase; DFR, dihydroflavonol reductase; ANS, Anthocyanidin synthese; 3GT: flavonoid 3-O-glucosyltransferase; LAR: Leucoanthocyanidin reductase; ANR: anthocyanidin reductase.
Figure 1. The biosynthetic pathway of terpenoids (A), alkaloids (B), and phenylpropanoids (C) in plants. The orange shaded area represents endogenous metabolic pathways present in microorganisms. The catalytic enzymes are shown in blue, intermediates in black. Solid arrow represents one step reaction, and dashed arrow represents multiple steps. The structures of typical natural products with underlines are displayed. Green circles indicate compounds for which the synthesis pathways have not been fully elucidated. Abbreviations: DXS,1-deoxy-D-xylulose 5-phosphate synthase; DXR, 1deoxy-D-xylulose 5-phosphate reductoisomerase; MCT, 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase; CMK,4-(cytidine 5‘diphospho)-2-Cmethyl-D-erythritol kinase; MECPS:ME-CDP synthase; HDS, 4-hydroxy-3-methylbut-2-enyl-diphosphate synthase; IDS: isoprenyl diphosphate synthase; IDI, isopentenyl diphosphate isomerase; GPP, geranyl diphosphate; GPPS, GPP synthase; GGPP, geranylgeranyl diphosphate; GGPPS, GGPP synthase; TPS: terpene synthase; AACT, acetyl-CoA acetyltransferase; HMGS, hydroxymethylglutaryl-CoA synthase; HMGR, 3-hydroxy-3-methylglutaryl-CoA reductase; MVK, mevalonate kinase; PMK, phosphomevalonate kinase; MPDC: mevalonate diphosphate decarboxylase; FPP, farnesyl diphosphate; FPPS, FPP synthase; STR: strictosidine synthase; PAL: phenylalanine ammonia lyase; C4H, cinnamate 4-hydroxylase; 4CL, 4-coumarate: CoA ligase; CHS, chalcone synthese; CHI, chalcone isomerase; F3H, flavanone 3-hydroxylase; F3′H, flavonoid-3′-hydroxylase; F5′H, flavonoid-5′-hydroxylase; DFR, dihydroflavonol reductase; ANS, Anthocyanidin synthese; 3GT: flavonoid 3-O-glucosyltransferase; LAR: Leucoanthocyanidin reductase; ANR: anthocyanidin reductase.
Molecules 29 02106 g001
Figure 2. Schematic graph for strategies applied in plant secondary metabolites production in plant and microbial chassis. Substrate a undergoes a series of enzyme-catalyzed reactions to ultimately yield product j through intermediates bi. Strategies for enhancing the yield of the final product include: 1. Dividing the entire reaction pathway into several modules for individual optimization. 2. Eliminating bottlenecks in the reaction pathway by overexpressing key enzyme genes. 3. Knocking down or knocking out enzyme genes that catalyze intermediate products into other non-target substances. 4. Enhancing overall reaction efficiency by adjusting the levels of regulators that simultaneously control multiple enzyme genes expression. 5. Dynamically controlling enzyme gene expression using inducible promoters to reduce cellular burden or toxic side effects. 6. Screening for the optimal combination of enzyme genes obtained from different species. 7. Enhancing enzyme activity, stability, and adaptability to the host by amino acid site mutagenesis, co-expression of redox partners, cofactor engineering, etc. 8. Storing target compounds in specific compartments by manipulating transport proteins. 9. Splitting metabolic pathways into several parts, expressing them in different cell lines, and then obtaining the final product through co-cultivation.
Figure 2. Schematic graph for strategies applied in plant secondary metabolites production in plant and microbial chassis. Substrate a undergoes a series of enzyme-catalyzed reactions to ultimately yield product j through intermediates bi. Strategies for enhancing the yield of the final product include: 1. Dividing the entire reaction pathway into several modules for individual optimization. 2. Eliminating bottlenecks in the reaction pathway by overexpressing key enzyme genes. 3. Knocking down or knocking out enzyme genes that catalyze intermediate products into other non-target substances. 4. Enhancing overall reaction efficiency by adjusting the levels of regulators that simultaneously control multiple enzyme genes expression. 5. Dynamically controlling enzyme gene expression using inducible promoters to reduce cellular burden or toxic side effects. 6. Screening for the optimal combination of enzyme genes obtained from different species. 7. Enhancing enzyme activity, stability, and adaptability to the host by amino acid site mutagenesis, co-expression of redox partners, cofactor engineering, etc. 8. Storing target compounds in specific compartments by manipulating transport proteins. 9. Splitting metabolic pathways into several parts, expressing them in different cell lines, and then obtaining the final product through co-cultivation.
Molecules 29 02106 g002
Table 2. Typical examples of plant secondary metabolites production in microbial chassis.
Table 2. Typical examples of plant secondary metabolites production in microbial chassis.
CompoundProductHostEngineering StrategyTiter (mg/L)VesselRef.
Terpenoids
Monoterpenes Molecules 29 02106 i038
Geraniol
E. coliOverexpression of heterotic geraniol synthase, GPP synthase, and mevalonate (MVA) pathway2000Fermentor[158]
Two-phase culture
Hydrolyzing geranyl acetate to geraniol by overexpression of acetylesterase
S. cerevisiaeOverexpression of isopentenyl diphosphate isomerase1 (IDI1) and truncated HMGR1680Fermentor[159]
Co-expression of the reverse fusion of truncated GES-Erg20WW variant and another copy of Erg20WW variant
Carbon restriction
S. cerevisiaeDeletion of OYE2 to reduce the endogenous metabolism of geraniol1690Fermentor[160]
Dynamic control of ERG20 expression
Overexpression of IDI1 and truncated HMGR
Molecules 29 02106 i039
Limonene
E. coliThe two-liquid phase fed-batch setup2700 Fermentor[161]
Limonene synthase from Mentha spicata and GPP synthase 2 from Abies grandis
S. cerevisiaeAn orthogonal engineering by introducing truncated neryl diphosphate synthase from Solanum lycopersicum and limonene synthase from Citrus lemon917.7Flask[162]
Expression of ERG20 regulated by the glucose-sensing promoter HXT1
Molecules 29 02106 i040
α-pinene
E. coliCo-expressed IspA from E. coli and Pt30 from Pinus taeda970Fermentor[163]
Heterologous MVA pathway and GPP synthase
Molecules 29 02106 i041
Linalool
S. cerevisiaeDirected evolution of linalool synthase53.14Flask[164]
Overexpression of the complete MVA pathway
SesquiterpenesMolecules 29 02106 i042
Patchoulol
S. cerevisiaeFusion expression of FPP and patchoulol synthase466.8Flask[165]
Enhanced expression of the limiting genes of the MVA
Expression of squalene synthase driven by HXT1 promoter
Farnesol biosynthesis was inhibited
S. cerevisiaeGlobal metabolic engineering strategy42.1Flask[166]
Modulated expression of nine genes involved
Molecules 29 02106 i043
Artemisinic acid
S. cerevisiaeDown-regulation of ERG9 under CTR3 promoter in response to copper and reductase CRP125,000Flask[167]
Additional introduction of CYB5, ADH1, and ALDH1
S. cerevisiaeERG9 controlled with promoter Pmet32500Fermentor[168]
Addition of methionine
DiterpenesMolecules 29 02106 i044
Sclareol
S. cerevisiaeRewiring central metabolism for supplication of acetyl-CoA and NADPH11,400Fermentor[169]
Optimization of MVA pathway
Fusion of two diterpene synthase and maltose-binding protein
Knock-out regulatory factors
E. coliReconstruction of sclareol biosynthetic pathway1500Flask[170]
High-cell-density fermentation
Molecules 29 02106 i045
Taxadiene
E. coliOptimally balanced expressions of up and downstream pathway modules1020Fermentor[171]
Two-phase cultivation
S. cerevisiaeFusion of GGPPS and taxadiene synthase184.2Fermentor[172]
Molecules 29 02106 i046
Miltiradiene
S. cerevisiaeOverexpression of MVA pathway genes3500Fermentor[173]
Downregulation of transcription factor ROS1 and competing enzyme ERG9
Fusion of two Class I and Class II terpene synthases
Molecules 29 02106 i047
Ferruginol
S. cerevisiaeFusion of GGPP synthase and FPS; fusion of copalyl diphosphate synthase and kaurene synthase-like10.5Flask[174]
Co-expression with plant CPR
TriterpenesMolecules 29 02106 i048
Ginsenosides
S. cerevisiaeSemi-rational design of UGT51300Fermentor[175]
Preventing Rh2 degradation and increasing UDP-glucose precursor supply
S. cerevisiaeModular engineering of the MVA and optimization of P450 expression levels2250Fermentor[176]
Increasing the copy number and expression level of UGTPg45 and direct evolution by random mutation
Molecules 29 02106 i049
Amyrin
S. cerevisiaeIntroducing efficient cytochrome P450s and pairing their reduction systems108.1Fermentor[177]
By increasing the copy number of Uni25647 and pairing CPRs
S. cerevisiaeEnzyme engineering of amyrin synthase1108Fermentor[178]
Overexpression of MVA pathway and diacylglycerol acyltransferase
Molecules 29 02106 i050
Oleanolic acid
S. cerevisiaeScreening for better CPR606.9Fermentor[179]
Knock-out galactokinase (GAL1) and negative transcriptional regulator (GAL80)
Molecules 29 02106 i051
Precursors of celastrol
S. cerevisiaeIncrease endogenous cytosolic MVA pathway15.3Flask[180]
TetraterpenesMolecules 29 02106 i052
Lycopene
S. cerevisiaeAdjusting the copy number of three key genes, knocking-out endogenous bypass genes3280Flask[181]
Increasing the supply of the precursor acetyl-CoA, balancing NADPH utilization
Using GAL-inducible system
E. coliModulating three genes to increase ATP and NADPH supply3520Fermentor[182]
Modulating pathway gene expressions by screening RBS library
Molecules 29 02106 i053
β-carotene
S. cerevisiaeOver-expression of HMGR6.29 mg/g DW Flask[183]
Addition of ergosterol biosynthesis inhibitors
S. cerevisiaeReconstructing controllable multi-gene pathways by employing the GAL regulatory system20.79 mg/g DWFlask[184]
Optimized by repeatedly using GAL10-GAL1 bidirectional promoters with high efficiency
Molecules 29 02106 i054
Astaxanthin
S. cerevisiaeIntroducing the GAL regulation system235Fermentor[185]
Adopting temperature as an input signal
S. cerevisiaePhysical mutagenesis by ARTP and adaptive evolution driven by H2O2404.78Fermentor[186]
Phenols
IsoflavonesMolecules 29 02106 i055
Genistein
E. coliScreening for better enzymes35Flask[187]
Translational fusion of RcIFS and OsCPR
S. cerevisiaeScreening for CHS31.02Flask[188]
Enhancement of precursor pathway and expression of feedback-resistant enzyme genes
Optimization of enzyme subcellular localizations
Molecules 29 02106 i056
Daidzin
S. cerevisiaeScreening for pathway genes73.2Flask[189]
Screening and overexpression of CPR partners
Increase co-factors supply
Dynamic control of p-coumaric acid supply
Enzyme fusion
StilbenoidsMolecules 29 02106 i057
Resveratrol
E. coliScreening for pathway genes238.71Flask[190]
Antisense inhibition of competing pathway to increase malonyl-CoA supply
Transport engineering for resveratrol secretion
Expression of chaperones to aid enzyme folding
S. cerevisiaeEnhance precursor pathway and expression of feedback-resistant enzyme genes4100Fermentor[191]
FlavanolsMolecules 29 02106 i058
Quercetin
S. cerevisiaeEngineered a delphinidin-overproducing S.cerevisiae-S. cerevisiae co-culture154.2Fermentor[192]
QuercetinS. cerevisiaeCombinational screening for downstream enzyme genes930Fermentor[193]
Molecules 29 02106 i059
Kaempferol
956
Increase gene copy numbers in genome
Enhance precursor pathway and expression of feedback-resistant enzyme genes
AnthocyaninsMolecules 29 02106 i060
Anthocyanin
S. cerevisiaeScreening pathway enzymes from different plants5.4
(total anthocyanins)
Flask[194]
Molecules 29 02106 i061
cyanidin 3-O-glucoside
E. coliEnhancing substrate and precursor availability350
(from (+)-catechin)
[195]
Balancing gene expression
Manipulate transportation by tolC knockout and yadH overexpression
ProanthocyanidinMolecules 29 02106 i062
(+)-catechin
E. coliCombinatorial screening pathway genes911
(from eriodictyol)
Flask[196]
Improving the availability of NADPH
NaphthoquinoneMolecules 29 02106 i063
3-geranyl-4-hydroxybenzoate acid
S. cerevisiaeIncrease GPP supply by increasing HMG1 and Erg20(K197G)179.29
(from hydroxybenzoate acid)
Flask[197]
Molecules 29 02106 i064
p-Coumaric acid
S. cerevisiaeScreening for p-hydroxybenzoate:geranyltransferase1890Fermentor[198]
Overexpression of feedback inhibition-resistant DAHP synthase and chorismate mutase
Knockout aromatic product pathway genes ARO10 (phenylpyruvate decarboxylase) and PDC5 (pyruvate decarboxylase)
Hydroxycinnamic acidsMolecules 29 02106 i065
Rosmarinic acid
E. coliEstablishment of partial pathway in different strains and individual optimization and combination of strains172Flask[199]
Alkaloids
Tropane alkaloidsMolecules 29 02106 i066
Scopolamine
S. cerevisiaeTransporter engineering by disrupting TPO5 and overexpression vacuolar tropine importers and two exporters0.172Flask[200]
Overexpression NADPH regeneration-related genes
Tryptophan prototrophy
Isoquinoline alkaloidsMolecules 29 02106 i067
Stylopine
E. coli and P. pastorisStylopine and reticuline module synthesized in P. pastoris and E. coli, respectively1.615Flask[201]
Inoculation ratio and medium were optimized
Molecules 29 02106 i068
(S)-reticuline
E. coliIntroducing two feedback-resistant enzymes307Flask[202]
Establishment of an additional heterologous pathway
S. cerevisiaeKnock-out of competing oxidoreductases consuming 4-hydroxyphenylacetaldehyde4600Fermentor[203]
Increase 4-hydroxyphenylpyruvate and dopamine supply
Screening and engineering norcoclaurine synthase
Molecules 29 02106 i069
Berberine
S. cerevisiaeKnockout competing oxidoreductases consuming 4-hydroxyphenylacetaldehyde1.08Fermentor[204]
Enhancement of the supply of tyrosine
Terpenoid indole alkaloidMolecules 29 02106 i070
Vinblastine
S. cerevisiaeScreening for better enzymes0.4Flask[205]
Downregulation of competing enzyme genes
Co-expression with CPRs
Molecules 29 02106 i071
Catharanthine
Pichia pastorisThree modules integrated in stable genome integration sites2.57Fermentor[206]
Optimize copy number of pathway genes
Increase in cofactor, S-adenosyl-l-methionine supply, and co-expression with CPR genes
Induction medium optimization
Molecules 29 02106 i072
Ajmalicine
S. cerevisiaeStable integration in the genome61.4Flask[207]
Increase precursor supply by increasing gene copies and knocking-out competing pathway genes
Co-expression with CPRs
Table 3. Advantages and disadvantages of plant and microbial chassis.
Table 3. Advantages and disadvantages of plant and microbial chassis.
HostAdvantagesDisadvantages
Plant
  • The natural host has prepared cofactors and precursors;
  • Organelles and higher-order membrane structures allow for the storage of intermediates and products;
  • Rich redox partners that are better matched with P450 enzymes.
  • Complex endogenous metabolic pathways that interfere with the target pathway;
  • Strict metabolic regulation, including potential unknown regulators;
  • Transgene expressions are prone to silencing;
  • Limited gene expression elements and unclear driving force of promoters;
  • Slow growth rates;
  • Demanding fermentation conditions and bioreactors and challenging to achieve high-density fermentation.
Microbes
  • Relatively clear endogenous metabolic pathways and regulatory mechanisms;
  • Convenient and rapid genetic manipulation;
  • Abundance of standardized expression elements;
  • Fast growth rates;
  • Easy achievement of large-scale and high-density cultivation.
  • Lack of advanced membrane structures and organelles;
  • Activity of exogenous enzymes may be hindered;
  • Additional provision of redox partners matching P450 enzymes and increased cofactor supply are required.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Han, T.; Miao, G. Strategies, Achievements, and Potential Challenges of Plant and Microbial Chassis in the Biosynthesis of Plant Secondary Metabolites. Molecules 2024, 29, 2106. https://doi.org/10.3390/molecules29092106

AMA Style

Han T, Miao G. Strategies, Achievements, and Potential Challenges of Plant and Microbial Chassis in the Biosynthesis of Plant Secondary Metabolites. Molecules. 2024; 29(9):2106. https://doi.org/10.3390/molecules29092106

Chicago/Turabian Style

Han, Taotao, and Guopeng Miao. 2024. "Strategies, Achievements, and Potential Challenges of Plant and Microbial Chassis in the Biosynthesis of Plant Secondary Metabolites" Molecules 29, no. 9: 2106. https://doi.org/10.3390/molecules29092106

Article Metrics

Back to TopTop