5,953
Views
18
CrossRef citations to date
0
Altmetric
Review Articles

Engineering microalgae: transition from empirical design to programmable cells

ORCID Icon, , , &
Pages 1233-1256 | Received 17 Sep 2020, Accepted 12 Feb 2021, Published online: 15 Jun 2021

Abstract

Domesticated microalgae hold great promise for the sustainable provision of various bioresources for human domestic and industrial consumption. Efforts to exploit their potential are far from being fully realized due to limitations in the know-how of microalgal engineering. The associated technologies are not as well developed as those for heterotrophic microbes, cyanobacteria, and plants. However, recent studies on microalgal metabolic engineering, genome editing, and synthetic biology have immensely helped to enhance transformation efficiencies and are bringing new insights into this field. Therefore, this article, summarizes recent developments in microalgal biotechnology and examines the prospects for generating specialty and commodity products through the processes of metabolic engineering and synthetic biology. After a brief examination of empirical engineering methods and vector design, this article focuses on quantitative transformation cassette design, elaborates on target editing methods and emerging digital design of algal cellular metabolism to arrive at high yields of valuable products. These advances have enabled a transition of manners in microalgal engineering from single-gene and enzyme-based metabolic engineering to systems-level precision engineering, from cells created with genetically modified (GM) tags to that without GM tags, and ultimately from proof of concept to tangible industrial applications. Finally, future trends are proposed in microalgal engineering, aiming to establish individualized transformation systems in newly identified species for strain-specific specialty and commodity products, while developing sophisticated universal toolkits in model algal species.

Graphical Abstract

Introduction

Microalgae, with an estimate of 72,500 species [Citation1], are of great ecological importance, as they contribute almost half of the global organic carbon fixation [Citation2]. They provide a variety of natural products to support ecosystems (e.g. coral reef ecosystems [Citation3]) by photosynthesis with efficiencies approximately three times greater than those achieved by land plants [Citation4]. The microalgal lifecycle of fast cell division and continuous biomass accumulation highlights the advantages gained from their sustainable cultivation [Citation5]. Accordingly, various attempts have been made to exploit microalgae for the production of commodity and specialty chemicals to meet human domestic, and industrial demands. To identify and improve species that naturally produce valuable compounds, mutation breeding and genetic modifications have been utilized for a long period, but are largely via empirical approaches [Citation6].

Admittedly, extensive reviews cover the bioengineering of cyanobacteria [Citation7], the genomic context underpinning microalgal diversity [Citation8], and the ecology, evolution, and applications of microalgae [Citation9,Citation10]. Therefore, this article primarily focuses on eukaryotic microalgae with a brief comparative study between microalgae and cyanobacteria. Two primary problems are preventing microalgal biotechnology from further development and industrialization. One is that only a few model microalgal species can be routinely transformed and ready for downstream industrialization. The exploitation of all but a handful of algal species is thus severely impeded by a limitation in molecular tools for competent engineering. The other problem resides in the extremely low transformation efficiencies (even compared to those achieved with plants) that are prevailing when working with microalgae, clearly demonstrating a need for novel and improved transformation techniques.

Recent studies on microalgal metabolic engineering, genome editing, and synthetic biology are facilitating the improvement of transformation efficiencies and are bringing new insights into cellular metabolic processes [Citation11]. There exists a demand to jointly consider these developments through a synthetic perspective of microalgal breeding. Therefore, after briefly reviewing empirical methods for vector design and engineering, specifically, this article focuses on (i) quantitative transformation cassette designs, (ii) elaborate target editing methods, (iii) digital design of algal cellular metabolism for the high yield of valuable products, and (iv) problems and countermeasures of industrial application. Drawing lessons from the broader history of the field and emerging advances, it is anticipated that a new era of rational design of digital microalgal cells is coming. This era would substantially benefit human society on food supply, energy consumption, and environmental sustainability.

Relevance of microalgae as photosynthetic cell factories

Broadly speaking, strategies for engineering microalgal metabolic pathways can be divided into two categories: those based on endogenous pathways, and those derived from non-native pathways by involving heterologous genes.

Natural compounds

Long-term adaptation to a wide range of ecotypes has engendered a diverse phenotype and genotype of microalgae, as well as helping algae evolve robust acclimation plasticity [Citation8], enabling them to adapt to various niches and produce a vast array of compounds. Microalgae are highly efficient at sequestrating CO2, accumulating biomass, and many secondary metabolites including pharmaceutically and nutritionally active compounds for human beings (or precursors for such compounds) (). It has therefore been suggested that genetically-tailored microalgae could serve as “platform strains” to convert CO2 into diverse and useful compounds from metabolic intermediates [Citation26]. However, current microalgal product yields are generally too low to meet the cost of commercial exploitation. Thus gene engineering approaches have been adopted in an attempt to overcome this drawback by modulating the activity of endogenous rate-limiting enzymes.

Table 1. Targeted compounds produced in microalgal cell factories.

Heterologous compounds

Beyond producing endogenous compounds, microalgae could potentially be recruited as cell factories to produce many different non-native compounds, ranging from small organic molecules to large recombinant proteins. Examples of the small molecules that could be produced using microalgae are provided in the report entitled “Top Value-Added Chemicals from Biomass” [Citation30], which identified twelve platform chemicals, that is, small organic compounds that can be produced from sugars by microorganisms and subsequently converted into industrially relevant molecules.

The attractiveness of using microalgae as hosts to produce larger biomacromolecules (e.g. recombinant proteins) can be understood by comparing the properties with that of alternative biological production systems, such as bacteria, yeasts, mammals, insects, or plants (). These desirable qualities, together with the rising demand for recombinant proteins, have driven this pursuit to introduce druggability into transgenic microalgae [Citation26]. Consequently, increasing numbers of recombinant proteins, including antibodies, immunotoxins, vaccine antigens, and mammary-associated serum amyloid, have been produced from metabolic intermediates in microalgae (). Yet, for all that, it still remains challenging to the sustainable production of foreign chemicals by introducing a de novo engineered pathway, necessitating the development of new approaches and advanced engineering strategies.

Table 2. Characteristics and advantages of representative expression systems.

As putative cell factories, cyanobacteria offer distinct advantages but also usually have some drawbacks, when compared with microalgae. Among the advantages are the ease of transformation and the absence of epigenetic regulation or suppressor mutations to counter the effects of transformation. As a result, cyanobacteria have been successfully engineered to make a variety of heterologous fuels and useful chemicals [Citation41–43]. A breakthrough has been achieved with the design of oligonucleotide fusion constructs (target genes are fused to the highly-expressed endogenous [Citation44] or exogenous genes [Citation45], as protein over-expression vectors that have been used in cyanobacteria to produce plant and human genes that are otherwise difficult to express [Citation44–46]. The “fusion constructs” could facilitate the heterologous proteins being accumulated as dominant cyanobacterial proteins, accounting for 20-25% of the total cell proteins [Citation44]. However, pertinent in this respect is that genes from eukaryotic organisms, for example, plants, animals, yeasts, and humans, are consistently expressed at low levels, in both microalgae and cyanobacteria, in spite of the use of strong promoters designed to confer “over-expression” of transgenes [Citation44]. Compared to microalgae, another drawback of the cyanobacterial system is that despite intensive industrial cultivation of cyanobacteria Spirulina (Arthrospira) species, the productivity of most cyanobacteria is lower under mass culture and bright sunlight conditions [Citation4]. There are reports of very-rapid-growth of unicellular cyanobacteria [Citation47,Citation48], however, they have not yet been tested on the industrial scale.

Engineering vectors: transition from empirical to quantitative designs

To deliver transgenes into microalgae, several methods have been developed, such as the glass-bead method, Agrobacterium-mediated transformation, electroporation, and particle bombardment. The comparison of these transformation methods and the characteristics of transformed microalgae have been listed in . They have been intensively reviewed elsewhere [Citation75] and thus are not discussed in detail in this review. It should be noted that despite the pros and cons, electroporation-based transformation outweighs other methods in terms of the wide applicable range in microalgal species. Another promising and potentially species-independent method is nanoparticle-mediated DNA delivery, yet to be applied in microalgae. It was first developed by using bacterial magnetic particles (50–100 nm in diameter) in the 1990s [Citation76] and was restricted due to difficulties in particle preparation [Citation77] but recently refined in plants [Citation73,Citation74]. Moreover, despite many challenges, particularly for microalgae holding a tough cell wall, when combined with a droplet microfluidics platform, electroporation [Citation78,Citation79] and nanoparticles [Citation80] are promising methods to convert the current “population transformation” of microalgae into high throughput “single-cell engineering” [Citation81]. One of the necessary jobs that remain for practical application is to decrease the cost and simplify the manipulation of microfluidics.

Table 3. DNA delivery methods of modern breeding strategies for microalgae.

Notwithstanding, the progress achieved in developing tools for delivering exogenous DNA, routine transformation is available only for a restricted number of microalgae [Citation82]. Moreover, even in established species, obstacles pertaining to gene delivery efficiency, transgene stability, or heritability are preventing the transformation systems from practice. Therefore, it necessitates a methodological transition from the tools used to manipulate metabolism, relying on experience-dependent strategies and constrained in particular species or strains [Citation83], to methods of quantitative and mathematic design.

Empirical designs

Critical to the creation of transgenic microalgae is the ability to transform cells with specific DNA sequences using vector constructs. To ensure the proper transcription, marker genes and/or genes of interest are typically expressed in individual cassettes harboring a 5′ promoter and a 3′ terminator. Conventional protocols for selecting marker genes for microalgal transformation have been extensively reviewed [Citation84]. Promoters are critical components that work in concert with other genetic elements (enhancers, silencers, transcription factors, and boundary elements/insulators) to direct the transcription of marker genes and other sequences. Promoter availability and selection thus profoundly influence the success of constructing robust genetic transformation systems. However, until now, few promoters are available for algal vectors, partially due to the shortcomings of the respective genetic toolkits (particularly the limited number of known regulatory elements).

Most attempts to engineer microalgae have been conducted empirically using a handful of repurposed tools. To achieve viable transformants, strong constitutive promoters from phylogenetically closely related algal species, viruses, or occasionally higher plants have been harnessed for transgene expression in microalgae. The viral promoters of CaMV35S (the cauliflower mosaic virus) and SV40 (Simian virus 40; an oncogenic simian polyomavirus) have been utilized for transient expression in some algal species, but heterologous promoter regions are usually inadequately recognized and regulated in microalgae [Citation85] (). In this respect, there are publications describing the successful transformation of microalgae with results that, unfortunately, could not be reproduced in other laboratories [Citation125]. For example, while the Arabidopsis thaliana U6 gene promoter was able to drive gene transcription in Chlamydomonas reinhardtii [Citation85], vectors feature in either Arabidopsis or Chlamydomonas promoters were not successfully recognized in Nannochloropsis sp. [Citation109].

Table 4. Examples of genetic engineering in a variety of microalgae.

As alternatives to exogenous promoters, vectors can be constructed using an orthologous promoter related to one that has been previously characterized. For instance, heat shock proteins [Citation86] and tubulin [Citation126] are highly expressed proteins in microalgae such as Chlamydomonas (). Their promoters are therefore regularly used to drive the constitutive nuclear expression of numerous genes in Chlamydomonas, and orthologous promoters have been used successfully in a number of other microalgae. However, microalgae tend to have numerous orthologs with quantitively unknown transcriptional and protein expression levels. In many cases, it is difficult to select suitable driving promoters for vector design. For example, Nannochloropsis oceanica strain IMET1 harbors nine orthologs for heat shock proteins and eight orthologs for tubulin. Each of these promoter regions was used separately to construct vectors and drive gene expression [Citation128]. However, success was achieved with only 50% of the assembled vectors even though the same transformation protocol was used in all cases. Similarly in N. oceanica strain 1779, the constructs harboring a C. reinhardtii α-tubulin promoter or a native lipid droplet surface protein (LDSP) promoter were employed to drive the expression of the Streptomyces hygroscopicus aph7 gene, conferring resistance to Hygromycin B. The latter achieves a more than 10-fold increase in transformation events compared with the former [Citation108]. The limited transformation efficiency of microalgal expression systems using conventional strategies has resulted in poorly reproducible transformation protocols and has constrained the scope of reverse genetic tool development. Therefore, a more rational approach for promoter dissection and the design of biological engineering systems is needed.

Quantitative methods for transforming cassette design

The dissection of regulatory elements is essential for the design of engineering systems, which can in turn facilitate an understanding of their natural counterparts [Citation129] (). However, the regulatory mechanisms remain unenlightened in microalgae, even with the model species C. reinhardtii. Therefore, characterized endogenous promoters have been the primary options when constructing customized vectors for specific microalgae. For example, a promoter for the gene encoding fucoxanthin-chlorophyll binding proteins (FCP) was isolated from diatoms, thoroughly tested, and widely used for the transformation of vector design (). It is noticeable that vectors incorporating this promoter exhibited stable and relatively high transformation frequencies in both biolistic and electroporation-induced transformations [Citation102,Citation130]. Additionally, several endogenous regulatory elements were characterized and incorporated into vectors for Chlamydomonas transformations using Agrobacterium [Citation89], electroporation [Citation89], silicon carbide whiskers [Citation91], glass beads [Citation49], and biolistic methods [Citation117], achieving comparable efficiencies. Thus, vector assembly is a key determinant of transformation efficiency. The frontiers of this field have been advanced by the development of novel promoter engineering strategies which are generally classified as (a) random mutagenesis; (b) hybrid promoter design; (c) de-novo promoter synthesis [Citation131].

Figure 1. Approaches of quantitative design for the engineering vectors. Abbreviations: RBS, ribosomal binding site; ORF, open reading frame.

Figure 1. Approaches of quantitative design for the engineering vectors. Abbreviations: RBS, ribosomal binding site; ORF, open reading frame.

Figure 2. Engineering strategies for microalgae. (a) Overexpression strategies for a single gene or multiple genes. (b) Random DNA insertional mutagenesis. (c) RNAi-mediated gene knockdown. (d) Homologous recombination. (e) Targeted genome editing by CRISPR/Cas9 technology. Abbreviations: R, resistance marker; Gene, interested gene(s); S, sense strand; AS, antisense strand; RuvC, RuvC nuclease domain; PAM, protospacer adjacent motif; HNH, HNH nuclease domain; crRNA, CRISPR RNA; tracrRNA, trans-activating crRNA.

Figure 2. Engineering strategies for microalgae. (a) Overexpression strategies for a single gene or multiple genes. (b) Random DNA insertional mutagenesis. (c) RNAi-mediated gene knockdown. (d) Homologous recombination. (e) Targeted genome editing by CRISPR/Cas9 technology. Abbreviations: R, resistance marker; Gene, interested gene(s); S, sense strand; AS, antisense strand; RuvC, RuvC nuclease domain; PAM, protospacer adjacent motif; HNH, HNH nuclease domain; crRNA, CRISPR RNA; tracrRNA, trans-activating crRNA.

In addition to promoters, incorporating regulatory elements such as introns and featured transcript sequences were also applied to improve the transformation efficiency and increase exogenous gene expression [Citation95]. Synergistic effects were utilized to improve transgene expression by either incorporating different intron portfolios into vectors or including a consensus Kozak sequence in the 5′ UTRs [Citation132] (). To allow proteins to target specific organelles, a leader-targeting sequence [Citation23], comprising transit peptides [Citation133], could be designed. These results highlight the key role of regulatory elements in the design of engineering systems and the necessity of dissecting the mechanisms and functions of different regulatory elements. Unfortunately, the diversity and potential of these regulatory mechanisms (or elements) in microalgae are largely unknown [Citation125] as a result of the low-throughput methods for characterizing the molecular mechanisms.

The ongoing expansion of sequenced genomes of microalgae (60 already completed or in preparation) and several cyanobacteria, and the associated omics information facilitates quantitative dissection of the mechanisms underpinning the functionality of promoters and other regulatory elements in a wide range of microalgae [Citation134]. Together with the development of sophisticated trapping systems [Citation135] and the computational analysis of biological components, it is increasingly viable to rationally and systematically identify, characterize, and standardize promoters, untranslated regions, terminators, enhancers, silencers, codon preferences, and other yet-to-be-discovered elements [Citation136]. This knowledge-driven strategy offers several potential advantages over traditional methods for establishing microalgal transformation systems. For instance, in silico prediction and investigation of regulatory elements considerably increases the likelihood of discovering active regulatory components. Additionally, system-level investigations on gene structure can help identify species-specific regulatory mechanisms and biological components, which facilitates the customized design of synthetic promoters, biological bricks, and circuits [Citation137]. Finally, cross-species genome comparisons are helpful to unveil universal regulatory rules operating in different microalgae and thereby enable the design of universal (at the species or genus levels) transgene vectors.

We see notable progress in the dissection of algal genetic elements in such a knowledge-driven manner. A number of strong promoters have been discerned from different Nannochloropsis species [Citation109] (). To enable multiple-gene expression, bidirectional promoters [Citation111] have also been isolated from this microalgal genus. Promoters, both constitutive and inducible under nitrogen starvation, were employed for customized transgene expression in the diatom P. tricornutum [Citation104]. Quantitative profiling of such components can enable their use in BioBricks – DNA sequences conforming to a restriction enzyme assembly standard and encoding one or more functional units, that can be used as part of a scalable transgenic toolbox. The assembly and engineering of individual components with defined functions from libraries of such standardized interchangeable parts could enable the design and the development of generalizable engineering tools.

Engineering strategies: a trend to editing methods

Gene overexpression

Deliberate overexpression of genes of interest (GOIs) is widely used for functional analysis (). Distinctive to the capacity for expressing multigene in a single operon in prokaryotic or prokaryotic-derived genomes (e.g. microalgal chloroplasts), and marker genes should be located in different expression cassettes in microalgae. Overexpression of multiple genes is a distinct advantage and requirement for trait optimization. Various strategies have been designed for multi-gene expression in microalgae (). For example, co-expressed genes can be cloned into separate vectors with different selective markers and then transformed into one strain sequentially [Citation140]. Alternatively, recently developed strategies for multigene engineering, involve designing expression vectors using bidirectional promoters [Citation111,Citation139], isocaudamers [Citation141], or using the self-cleaving peptide F2A (e.g. in C. reinhardtii [Citation142], Nannochloropsis [Citation111], or Neochloris oleoabundans [Citation99]). Gene stacking overexpression systems have been established in some microalgae (e.g. C. reiharditti, Nannochloropsis sp., Chlorella sp., and P. tricornutum). This system is used to study specific metabolic pathways and to express intracellular [Citation143] or secreted [Citation144] recombinant proteins. These technologies simplify the transformation process, diminish the number of marker genes required for multi-step integration, and improve the dosage of target protein(s) after fusion with proper selectable markers.

Random mutagenesis

Overexpression platforms serve as foundations for the establishment of a comprehensive engineering system incorporating technologies of insertional mutagenesis, RNAi-mediated gene knockdown, and genome editing. Identifying mutants with an interesting phenotype is a classical genetic approach for probing life processes (). One way is to use transfer DNA (T-DNA), which is randomly integrated into genomes. It features tag-like sequences that facilitate the identification of the affected gene(s) [Citation145]. Therefore, gene inactivation by T-DNA mediated transposon insertion has been used in functional studies on diverse plant species that have in turn accelerated both basic research and applied plant biotechnology. However, T-DNA has not been demonstrated in algae thus far. Alternatively, by random insertion of transforming cassettes, whole-scale genome mutagenesis libraries have been generated only in a single microalga C. reinhardtii [Citation146]. The mutant population has proved to be effective to dissect the function of key genes in photosynthesis [Citation147] and TAG biosynthesis [Citation148]. However, current methods for flanking sequence retrieval tend to be species-specific: the application of these methods is always constrained to a single species and can only determine insertion sites for that particular species, and even worse, at a low frequency [Citation149]. A tagged mutant library was recently created for C. reinhardtii in a similar fashion to T-DNA that was successful in A. thaliana [Citation148]. It was utilized to dissect genes involved in biological processes, such as photosynthesis [Citation147] and CO2 concentration [Citation150,Citation151]. The generality of this approach highlights its potential in investigating the function of non-lethal genes in many other microalgae. However, a drawback is that loss-of-function mutations caused by random disruptions are unlikely to probe the function of lethal genes. In addition, the random mutagenesis approach is of limited utility in studies targeting specific genes and in hypothesis-driven tests.

RNA interference (RNAi)-mediated gene knockdown

RNAi-based gene silencing is a superior approach for the oriented knockdown of arbitrarily chosen genes (). Despite the apparent loss of core elements in many microalgal species [Citation152], RNAi-mediated silencing appears to be widespread among algal lineages [Citation128]. Although its biological machinery remains elusive [Citation153], sophisticated systems for small RNAs have been discovered in green, red, brown algae, diatoms, and dinoflagellates [Citation154–158]. RNAi-mediated knockdown has been established for species including C. reinhardtii [Citation159], N. oceanica [Citation128], and P. tricornutum [Citation160]. Nevertheless, the RNAi method has several general shortcomings such as unstable phenotype heritability and unpredictable activities of target genes [Citation161].

Targeted genome editing

To avoid unwanted phenotypes brought by random transforming cassette insertion (when using overexpression) or off-target knock-down (when using RNAi), it is desirable to develop tools for the precise modification of pre-selected sequences. Traditionally, homologous recombination is a routine practice for the genetic manipulation of prokaryotes. It was also developed in recombinogenic lower eukaryotes [Citation162] (). However, this method delivers only limited success in most eukaryotic microalgae, where non-homologous recombination tends to occur in preference to the homologous one [Citation163]. These problems have been circumvented by developing strategies based on engineered nucleases, which substantially facilitates targeted gene disruption and is revolutionizing many areas of science [Citation164]. Although the zinc-finger nuclease (ZFN) and transcription activator-like effectors (TALEN) are excellent tools for targeted gene knockout, they have only been proven to be useful in C. reinhardtii [Citation165] and P. tricornutum [Citation166]. Clustered regularly interspaced palindromic sequences (CRISPR/Cas9) technology is more efficient and much easier in design than ZFN and TALEN technologies [Citation167]. The straightforward design of CRISPR systems allow simultaneous disruption in multiple genes [Citation168] and the creation of CRISPR mutant libraries [Citation169] (). The list of applications of CRISPR has expanded to include knock-in [Citation170] and CRISPR interference (CRISPRi), which involves inducing sequence-specific interference of a target gene’s transcription [Citation171]. By CRISPR/Cas9-derived activator systems, multiple endogenous (non)coding genes can be activated simultaneously [Citation172]. The ability to upregulate any endogenous gene(s) provides unprecedented opportunities to look into and reconstruct cellular behavior. Therefore, an increasing number of microalgal species are engineered by CRISPR/Cas9 systems, such as C. reinhardtii [Citation165], Nannochloropsis sp. [Citation24], and P. tricornutum [Citation173]. However, the CRISPR/Cas9-based technology in microalgae is still in its infancy, as evidenced by low efficiencies and the need for more intensive mutant screening than in plants and animals. To ensure efficiency and accuracy, standardized targeted genome editing protocols should be developed for microalgae on the basis of in-depth fundamental investigations, particularly on the effects of endogenous silence of introduced genetic materials [Citation174].

From nucleus to organelle engineering

Nuclei, chloroplasts, and mitochondria are three membrane-bounded organelles in the microalgal cell. Although most of the preceding examples involve nuclear transformation, the nuclear expression has the drawback of sensitivity to epigenetic effects and random insertion, which can cause variable transgene expression. In contrast, chloroplast engineering shows several key advantages, such as absence of epigenetic gene silencing, targeted localization of transgenes in the genomes, and relatively high expression levels [Citation175]. Microalgal chloroplast genomes (e.g. in Nannochloroposis sp.) possess approximately 120 genes involved in photosynthesis and the gene expression system [Citation176]. In principle, all of these genes could be precisely manipulated. Furthermore, a number of nuclear gene products (such as those involved in photosynthesis, oil metabolism, or plant hormone synthesis) are targeted to plastids [Citation177,Citation178]. For example, in Arabidopsis, more than 2000 nuclear-encoded proteins are predicted to localize in chloroplasts and contribute to cellular properties critical for the growth or the production of essential cellular compounds [Citation179]. This further enlarges the category of biological traits which could be optimized by plastome engineering.

Although chloroplast transformation was initially developed in the green microalga C. reinhardtii [Citation117], plastome genetic engineering was more thoroughly developed in and applied to plant species. The use of chloroplast engineering to produce pharmaceuticals in crop plants presents several fundamental challenges [Citation180]. Problems include slow plant growth [Citation4], especially in cases when a high yield and a recombinant product amount is required. This problem is further accentuated by the seasonal growth of crops, as opposed to the year-round cultivation of microalgae. Moreover, transgene containment [Citation181] is always more difficult to cope with in plants than that in photobioreactor-enclosed photosynthetic microalgae. Microalgal platforms avoid many undesirable issues since these systems can be cultivated under tightly controlled and contained conditions in closed photobioreactors. In addition, the difference in codon usage in microalgae from that of plants helps avoid contamination of the human food chain with heterologous proteins due to cross-pollination [Citation182]. Biocontainment can be improved further by codon reassignment of the transgenes in the chloroplast [Citation119]. Taken together, barriers encountered in the use of plant-based systems to produce recombinant protein or other heterologous products would make them less attractive than photosynthetic microorganisms, for example, microalgae and cyanobacteria [Citation29]. More than 100 foreign or native proteins have thus far been produced in algal chloroplasts, including 40 different therapeutic proteins produced in C. reinhardtii chloroplasts [Citation183]. shows some case studies of chloroplast transformation. Further examples are given in recent reviews [Citation183].

However, despite some notable advances, microalgal plastid transformation is only available for a relatively small number of species [Citation82]. Chloroplast transformation has mainly relied on bombardment with gold microparticles laced with the transforming DNA, which is intractable for most microalgae. The smallest readily available gold particles have a diameter of around 0.6 µm which exceeds the dimensions of most microalgal chloroplasts [Citation123]. An alternative method is via polyethylene glycol (PEG)-mediated protoplast transformation [Citation184]. However, an intrinsic drawback of this method is the requirement of cell wall removal prior to transformation. Protoplast preparation is technically demanding and effectively impossible for many microalgal species with complex cell walls [Citation185]. A simple and straightforward electroporation-based method was recently devised for transforming Nannochloropsis chloroplasts [Citation123]. However, expression levels of transgenes in microalgal chloroplasts remain low compared to plant plastid and cyanobacterial expression systems. To bring plastome engineering into full play, it is critical to establish protocols that enable efficient and reliable expression for individual transgenes or entire pathways. It is thus indispensable to identify several components, such as constitutive promoters for robust expression [Citation118], controllable expression systems for spatially or temporally inducible expression, and optimized genetic elements for multigene expression [Citation84,Citation120]. Synthetic biology principles have been applied to plastome engineering, allowing ex vivo assembly, modification and duplication of the entire Chlamydomonas chloroplast genome [Citation186]. This proof of concept experiment demonstrates that de novo synthesis of algal chloroplast genomes is now possible.

From genetically modified (GM) to “non-GM” engineering

GM crops have been commercially used for decades. Available impact studies of pest-resistant crops show that these technologies are beneficial to farmers and consumers, producing large aggregate welfare gains as well as positive effects on the environment and human health [Citation187,Citation188]. Nonetheless, there has been substantial debate regarding the merits and ethics of GM and non-GM breeding technologies. Widespread public reservations have led to a complex system of regulations and in most countries, the commercial use and cultivation of GM organisms (GMOs) is heavily regulated [Citation189]. For example, the EU arguably has the strictest regulations in the world for the use of GMOs in both food and feed. Specifically, organisms obtained by techniques of genome editing (that cannot be distinguished from a conventionally-bred variety or a naturally-occurring variant; without GM tags) are regarded as GMOs and subject to the same restriction as other transgenic organisms [Citation189]. GM microalgae are created by inserting DNA-containing marker genes and additional GOIs into the host’s genome. A critical concern regarding the GM approach is the risk of promoting antibiotic resistance. To address this concern, techniques for generating transgenic algae should not rely on antibiotic resistance markers. Thus, successful C. reinhardtii chloroplast transformation and strain selection protocol with zero false-positives was achieved upon reconstitution of a functional rbcL gene recovery of Rubisco catalytic activity, coupled with the heterologous expression of the Saccharomyces cerevisiae alcohol dehydrogenase ADH1 gene in the algal chloroplast [Citation190]. This recovery of function, however, requires that photosynthesis-lethal microalgal mutants, which act as the transgene recipient strain, can be propagated heterotrophically. Alternatively, a recombinase has been applied to remove marker genes in engineered cyanobacteria [Citation191]. Extrachromosomal vectors, or episomes, offer another tool for “non-GM” breeding in microalgae [Citation192]. Combined with genome editing techniques, engineered cells of diatoms [Citation193] and N. oceanica [Citation194] maintained the desired traits while episomes (harboring the marker gene) were lost. It highlights the potential to create a “non-GM” microalga by the development of tools derived from episomes with specific and universal algal hosts [Citation195]. This highly efficient genetic system will accelerate knowledge transfer from fundamental research to tangible applications.

History of microalgal engineering approaches: a starting point for future development

Genetic manipulation has been accomplished in a wide spectrum of microalgae. In this section, we will revisit viable ways to generate useful products by engineered microalgae.

Conventional gene engineering – technologies based on single-enzyme encoding genes

Metabolic engineering is a purposeful modification of metabolic pathways for customized or improved production of useful compounds [Citation196] (). Historically, metabolic engineering efforts in microalgae have largely centered on modifying a single gene at a time, typically one encoding a rate-limiting enzyme. Biosynthetic pathways are retrofitted to preferentially flux endogenous carbon toward desirable-end chemicals [Citation23], but not to byproducts [Citation197]. Strategies adopted for this purpose include yield improvement [Citation102,Citation198], substrate range expansion [Citation138], novel chemicals synthesis [Citation26], and optimization of cellular robustness [Citation199,Citation200]. Subsequent developments in molecular and chemical biology led to the design and reconstruction of complete metabolic networks and even whole organisms rather than just individual reactions. It is from these studies that the phrase “cell factories” emerged. High-throughput omics techniques make it possible to rapidly link genotypes and phenotypes and to accelerate the engineering processes [Citation201]. These revolutionary techniques and mindsets have transformed metabolic engineering into a new field comprising genome-scale engineering, which incorporates computational biology, versatile approaches for interrogating and understanding cellular metabolism, and multiplex optimizations to reach diverse phenotypes [Citation202].

Figure 3. Comparison between metabolic engineering and system biology. Adapted from [Citation6].

Figure 3. Comparison between metabolic engineering and system biology. Adapted from [Citation6].

Transcriptional engineering – a control-knob-gene-based approach

Mixed success has been achieved in microalgae using the aforementioned conventional genetic engineering strategies. In some cases, the abundance of the desired metabolite remained unaltered. Transcriptional engineering provides an alternative strategy where a control-knob gene can be modified to simultaneously modulate multiple steps of a metabolic pathway [Citation203]. Initial attempts at trait improvement of microalgae by this strategy involved expressing plant transcription factors (TFs) and resulted in enhanced production of target chemicals [Citation204]. This success prompted a pursuit of genome-wide identification of microalgal TFs [Citation205], which were subsequently utilized to optimize the desired properties [Citation24,Citation202].

Synthetic biology – a systems-level precise engineering

The advance of metabolic engineering along with the increasing blending of DNA technology gave genesis to the field of synthetic biology (). A genetic toggle switch prototype was first devised in bacteria by metabolic engineers [Citation206], elaborated by synthetic biologists [Citation207], and consummated in mice [Citation208]. Unlike metabolic engineering, synthetic biology may use unnatural molecules to create artificial biological systems [Citation6]. While this approach has been successfully implemented in bacteria, yeasts, higher plants, and animals by characterizing and collecting reusable and standard biological parts [Citation209,Citation210], microalgal synthetic biology remains in its infancy. Attempts have been undertaken in only a handful of species. Their successes have been modest because of the underdevelopment of genetic tools and the limited knowledge of gene regulation mechanisms in microalgae, as discussed precedingly [Citation211]. However, advances in several species (e.g. Nannochloropsis sp., C. reinhardtii, and P. tricornutum) are now allowing an expansion to the realm of genome-wide reprogramming beyond conventional individual gene manipulation [Citation212]. Together with the ongoing expansion of microalgal omics datasets, it is possible to use these species as chassis for synthetic biology. A versatile and modular vector toolkit has been generated for C. reinhardtii, which features a standardized collection of reporters, selectable markers, and targeting peptides that can be assembled and repurposed [Citation213]. Broadly speaking, this toolkit generation highlights the inaugural application of synthetic biology in microalgal engineering and enables the development of digital photosynthetic cells.

Future perspectives

Digital photosynthetic cells

Principles and methods of synthetic biology from well-established systems are increasingly introduced into microalgae. However, the development of microalgae as true “programmable” entities is hampered not only by the accessibility of well-characterized and repurposed genetic elements but also by a lack of high-throughput methodology to identify, standardize, and rationalize basic parts and modules, not to mention systems-level circuitry. This is true even for C. reinhardtii, of which, among all microalgal species, the transcriptional regulation and cellular metabolism are best understood. The systematic knowledge of its regulatory elements and potential BioBricks remains elusive. Despite these challenges, the development of standardized and scalable methods for module assembly (e.g. Gibson assembly and Golden Gate cloning), engineering (i.e. transformation), colony handling (e.g. colony picking robots), and high-throughput strain screening (e.g. microfluidics) will give microalgal researchers unprecedented access to “digital photosynthetic cells” in which each gene, pathway, and the biological process can be quantitatively designed and modulated.

From proof of concept to tangible industrial application

Enhancement in the microalgal biofuel R&D occurred in the period between 2006 and 2011 because the global petroleum price skyrocketed [Citation214]. However, with the 2014 collapse in the price of fossil fuels, the renewable biofuels field is at a crossroads and its further development becomes foggy. End-products with an interest of the algal community now mainly focus on primary and secondary metabolites, such as proteins, essential oils, specialty chemicals, and biopharmaceuticals. These comprise sizable markets, albeit not as large as that offered by the transportation fuels field. Moreover, the diversity of natural secondary metabolite in microalgae is far from being fully exploited, suggesting the promising development of future R&D and commercialization. Meanwhile, it may be necessary to fill the gaps between laboratory R&D and tangible industrial production by engineering microalgae with crucial and economically relevant pathways of value-added compounds, such as wax [Citation102], astaxanthin [Citation215], and non-native isoprenoids (e.g. β-phellandrene; lupeol, bisabolene, patchoulol, geranyllinalool, and 13R(+) manoyl oxide) [Citation216,Citation217]. Plant secondary metabolites are a valuable reservoir of drugs and numerous of these compounds, such as ginsenosides and artemisinin, are among the most promising and commercially important biopharmaceuticals that can be generated in microalgae and cyanobacteria. These pathways and the associated key genes could be characterized using microalgal systems which resemble those of plants, but possess unique advantages as mentioned above. Moreover, unlike yeast or bacterial systems, microalgae are promising hosts to produce these metabolites in scalable culturing systems. The development of market-ready drug-producing systems using industrial microalgal strains is just beginning [Citation193,Citation216–218].

To realize the economic viability of commercial microalgal products, there is a need to inform on several major areas, including: (i) For newly identified species with industrial potential, the challenge would be to deliver exogenous DNA into the cells. (ii) Universally applied sophisticated engineering toolkits are essential in efforts to produce versatile biochemicals and biomaterials for agricultural, biomedical, and industrial ends. (iii) The development of genetic circuits, gene switches, systems-level circuitry, synthetic devices, synthetic genomes, genome editing technologies, and practices for microbiome engineering [Citation219] will benefit fundamental studies in microalgal synthetic biology. (iv) Genome-wide activation/knock down/knock out of multiple endogenous genes and simultaneous overexpression of multiple exogenous genes involved in target pathways or networks would greatly facilitate the creation of “super” microalgal strains as customized cell factories. (v) Although the application of glycosylation mechanisms is largely untapped in microalgae [Citation220,Citation221], “humanization” of microalgae glycans via glycoengineering would greatly facilitate the development of active and safe biopharmaceuticals production with proper folding and addition of immunogenic glycans [Citation39].

The sustainability of the target chemical productivity in engineered algal cells has been assessed in bench-scale (10 L) [Citation102] and in pilot-scale photobioreactors (PBRs; 550 L) [Citation216] in controllable laboratory conditions or in mimicked outdoor environments [Citation24]. The scalability and stability of introduced traits in GM microalgae were also probed by culturing cells at different PBR configurations and volumes [Citation222]. Although engineered phenotypes appeared to be stable and robust across these tested scales (from 800 mL to 100 L) and PBR configurations (i.e. vertical column, vertical flat-plate, and open raceway pond), reports on life-cycle assessments of bioenergy or other value-added chemicals using engineered microalgae are not yet available.

A barrier to the commercial-scale algal biofuel production is the relatively high capital and operating costs of PBRs, combined with the need to deliver fuel products at very low cost, typically less than $1 per kg. Coupling biofuel production with environmental applications, or the co-generation of high-value products, could achieve mutual benefits in renewable energy production and mitigation of greenhouse gas emission [Citation223] or removal of pollutants from wastewater [Citation224]. Moreover, extensive R&D is essential for studies on lowering the cost of downstream processing, such as industrial-scale cultivation, contamination control, cell harvesting and disruption, and extraction of target biomolecules. There is no doubt that engineering technology, in particular, simultaneously targeting/activation/overexpression/knockout of multiple genes, would play critical roles in achieving the objective. For example, truncated light-harvesting antennae of the photosystems can be generated to improve sunlight penetration and utilization, resulting in greater photosynthetic productivity of cultures with a high cell density under bright sunlight conditions [Citation225]. Versatile contaminant control strategies have also been developed for open pond cultivation by modifying nutrition mode [Citation226], expressing herbicide resistance genes [227], or genome-based chemical biology [228]. Target biomolecules could be extracted using cost-effective cell disruption methods by expressing algicides [229] or upon bypassing the disruption step employing an inducible green recovery strategy [230]. Overall, sound opportunities exist for the development of market-ready multiple-product systems by employing microalgae as digital photosynthetic cells.

Conclusions

We have reviewed the trend of microalgal engineering in the era of synthetic biology. In this article, the bioengineering of cyanobacteria, genomes, and evolution, and applications of microalgae were not at the center of our attention, but the bottleneck and the development of the engineering techniques of eukaryotic microalgae were considered. As mentioned, microalgal metabolic engineering, genome editing, and synthetic biology are increasingly intervening in microalgal biotechnology due to their numerous advantages over mutagenesis strategies using chemical and physical mutagens. Microalgal engineering technologies rapidly transfer from empirical methods to quantitative design, as that target genome editing and synthetic biology play more and more important roles in breeding new algae and the digital design of algal cellular metabolism. The current trend explicitly demonstrates that alongside technological progress in transforming techniques, algal engineering develops mostly toward customized transformation systems, systems-level precision engineering, and cells created without GM tags. All these aspects are going to be integrated with big data analysis of the increasing omics dataset. Despite positive horizons, the costs, the development of scale-up configurations, as well as the public concerns about GMOs, are dominant challenges.

Acknowledgments

We thank Dr. Jian Xu (Qingdao Institute of BioEnergy and Bioprocess Technology, Chinese Academy of Sciences) for very helpful discussion. We are grateful to the anonymous reviewers for their valuable improvement to this manuscript.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

This manuscript was supported by the Basic and Applied Basic Research Programs for the Talents of Hainan Province [2019RC033], the National Natural Science Foundation of China [32060061], the Project of Innovation & Development of Marine Economy [HHCL201803], the Chinese Government Scholarship [CSC201807565004], the Foundation of Hainan University [KYQD1561], and the Project of Collaborative Innovation of Tianjin University and Hainan University [HDTDU201903], the Maryland Industrial Partnerships Program [07432143, 07431629].

References

  • Guiry MD. How many species of algea are there? J Phycol. 2012;48(5):1057–1063.
  • Melillo JM, McGuire AD, Kicklighter DW, et al. Global climate change and terrestrial net primary production. Nature. 1993;363(6426):234–240.
  • Lu Y, Jiang J, Zhao H, et al. Clade-specific sterol metabolites in dinoflagellate endosymbionts are associated with coral bleaching in response to environmental cues. mSystems. 2020;5(5):e00765.
  • Melis A. Solar energy conversion efficiencies in photosynthesis: minimizing the chlorophyll antennae to maximize efficiency. Plant Sci. 2009;177(4):272–280.
  • Scaife MA, Nguyen GTDT, Rico J, et al. Establishing Chlamydomonas reinhardtii as an industrial biotechnology host. Plant J. 2015;82(3):532–546.
  • Nielsen J, Fussenegger M, Keasling J, et al. Engineering synergy in biotechnology. Nat Chem Biol. 2014;10(5):319–322.
  • Sengupta A, Pakrasi HB, Wangikar PP. Recent advances in synthetic biology of cyanobacteria. Appl Microbiol Biotechnol. 2018;102(13):5457–5471.
  • Brodie J, Chan CX, De Clerck O, et al. The algal revolution. Trends Plant Sci. 2017;22(8):726–738.
  • Wijffels RH, Kruse O, Hellingwerf KJ. Potential of industrial biotechnology with cyanobacteria and eukaryotic microalgae. Curr Opin Biotech. 2013;24(3):405–413.
  • Renuka N, Guldhe A, Prasanna R et al. Microalgae as multi-functional options in modern agriculture: current trends, prospects and challenges. Biotechnol Adv. 2018;36(4):1255–1273.
  • Jagadevan S, Banerjee A, Banerjee C, et al. Recent developments in synthetic biology and metabolic engineering in microalgae towards biofuel production. Biotechnol Biofuels. 2018;11:185.
  • Steinrücken P, Erga SR, Mjøs SA, et al. Bioprospecting North Atlantic microalgae with fast growth and high polyunsaturated fatty acid (PUFA) content for microalgae-based technologies. Algal Res. 2017;26:392–401.
  • Poliner E, Pulman JA, Zienkiewicz K, et al. A toolkit for Nannochloropsis oceanica CCMP1779 enables gene stacking and genetic engineering of the eicosapentaenoic acid pathway for enhanced long-chain polyunsaturated fatty acid production. Plant Biotechnol J. 2018;16(1):298–309.
  • Song X, Wang J, Wang Y, et al. Artificial creation of Chlorella pyrenoidosa mutants for economic sustainable food production. Biores Technol. 2018;268:340–345.
  • Ryckebosch E, Bruneel C, Termote-Verhalle R, et al. Nutritional evaluation of microalgae oils rich in omega-3 long chain polyunsaturated fatty acids as an alternative for fish oil. Food Chem. 2014;160:393–400.
  • Kyriakopoulou K, Papadaki S, Krokida M. Life cycle analysis of β-carotene extraction techniques. J Food Eng. 2015;167:51–58.
  • Suh IS, Joo HN, Lee C-G. A novel double-layered photobioreactor for simultaneous Haematococcus pluvialis cell growth and astaxanthin accumulation. J Biotechnol. 2006;125(4):540–546.
  • Wang Y, Liu Z, Qin S. Effects of iron on fatty acid and astaxanthin accumulation in mixotrophic Chromochloris zofingiensis. Biotechnol Lett. 2013;35(3):351–357.
  • Guo B, Liu B, Yang B, et al. Screening of diatom strains and characterization of Cyclotella cryptica as a potential fucoxanthin producer. Marine Drugs. 2016;14(7):125.
  • Ho SH, Chan MC, Liu CC, et al. Enhancing lutein productivity of an indigenous microalga Scenedesmus obliquus FSP-3 using light-related strategies. Biores Technol. 2014;152:275–282.
  • Lin J, Lee D, Chang J. Lutein production from biomass: marigold flowers versus microalgae. Biores Technol. 2015;184:421–428.
  • Randhir A, Laird DW, Maker G, et al. Microalgae: a potential sustainable commercial source of sterols. Algal Res. 2020;46:101772.
  • Xue J, Niu YF, Huang T, et al. Genetic improvement of the microalga Phaeodactylum tricornutum for boosting neutral lipid accumulation. Metab Eng. 2015;27:1–9.
  • Ajjawi I, Verruto J, Aqui M, et al. Lipid production in Nannochloropsis gaditana is doubled by decreasing expression of a single transcriptional regulator. Nat Biotechnol. 2017;35(7):647–652.
  • Li Y, Han D, Hu G, et al. Chlamydomonas starchless mutant defective in ADP-glucose pyrophosphorylase hyper-accumulates triacylglycerol. Metab Eng. 2010;12(4):387–391.
  • Tran M, Van C, Barrera DJ, et al. Production of unique immunotoxin cancer therapeutics in algal chloroplasts. Proc Natl Acad Sci USA. 2013;110(1):E15–E14.
  • Jones CS, Luong T, Hannon M, et al. Heterologous expression of the C-terminal antigenic domain of the malaria vaccine candidate Pfs48/45 in the green algae Chlamydomonas reinhardtii. Appl Microbiol Biotechnol. 2013;97(5):1987–1995.
  • Gregory JA, Topol AB, Doerner DZ, et al. Alga-produced cholera toxin-Pfs25 fusion proteins as oral vaccines. Appl Environ Microbiol. 2013;79(13):3917–3925.
  • Betterle N, Martinez DH, Melis A. Cyanobacterial production of biopharmaceutical and biotherapeutic proteins. Front Plant Sci. 2020;11:237.
  • Werpy T. Petersen G. Top value added chemicals from biomass: volume I–results of screening for potential candidates from sugars and synthesis gas. Golden (CO): National Renewable Energy Lab; 2004.
  • Macek B, Forchhammer K, Hardouin J, et al. Protein post-translational modifications in bacteria. Nat Rev Microbiol. 2019;17(11):651–664.
  • Griesbeck C, Kobl I, Heitzer M. Chlamydomonas reinhardtii: a protein expression system for pharmaceutical and biotechnological proteins. Mol Biotechnol. 2006;34(2):213–224.
  • Wang D, Lu Y, Huang H. et al. Establishing oleaginous microalgae research models for consolidated bioprocessing of solar energy. Adv Biochem Eng Biol. 2012. doi: https://doi.org/10.10077/10_2011_122.
  • Wildt S, Gerngross TU. The humanization of N-glycosylation pathways in yeast. Nat Rev Microbiol. 2005;3(2):119–128.
  • Dorai H, Ganguly S. Mammalian cell-produced therapeutic proteins: heterogeneity derived from protein degradation. Curr Opin Biotechnol. 2014;30:198–204.
  • Webster DE, Thomas MC. Post-translational modification of plant-made foreign proteins; glycosylation and beyond. Biotechnol Adv. 2012;30(2):410–418.
  • Doran PM. Foreign protein production in plant tissue cultures. Curr Opin Biotechnol. 2000;11(2):199–204.
  • Huang T, McDonald K. Bioreactor systems for in vitro production of foreign proteins using plant cell cultures. Biotechnol Adv. 2012;30(2):398–409.
  • Barolo L, Abbriano RM, Commault AS, et al. Perspectives for glyco-engineering of recombinant biopharmaceuticals from microalgae. Cells. 2020;9(3):633.
  • Mathieu-Rivet E, Kiefer-Meyer M-C, Vanier G, et al. Protein N-glycosylation in eukaryotic microalgae and its impact on the production of nuclear expressed biopharmaceuticals. Front Plant Sci. 2014;5:359.
  • Lindberg P, Park S, Melis A. Engineering a platform for photosynthetic isoprene production in cyanobacteria, using Synechocystis as the model organism. Metab Eng. 2010;12(1):70–79.
  • Ungerer J, Tao L, Davis M, et al. Sustained photosynthetic conversion of CO2 to ethylene in recombinant cyanobacterium Synechocystis 6803. Energy Environ Sci. 2012;5(10):8998–9006.
  • Formighieri C, Melis A. Heterologous synthesis of geranyllinalool, a diterpenol plant product, in the cyanobacterium Synechocystis. Appl Microbiol Biotechnol. 2017;101(7):2791–2800.
  • Formighieri C, Melis A. A phycocyanin·phellandrene synthase fusion enhances recombinant protein expression and β-phellandrene (monoterpene) hydrocarbons production in Synechocystis (cyanobacteria). Metab Eng. 2015;32:116–124.
  • Betterle N, Melis A. Heterologous leader sequences in fusion constructs enhance expression of geranyl diphosphate synthase and yield of β-phellandrene production in cyanobacteria (Synechocystis). ACS Synth Biol. 2018;7(3):912–921.
  • Chaves JE, Rueda-Romero P, Kirst H, et al. Engineering isoprene synthase expression and activity in cyanobacteria. ACS Synth Biol. 2017;6(12):2281–2292.
  • Ungerer J, Wendt KE, Hendry JI, et al. Comparative genomics reveals the molecular determinants of rapid growth of the cyanobacterium Synechococcus elongatus UTEX 2973. Proc Natl Acad Sci USA. 2018;115(50):E11761–E11770.
  • Bernstein HC, McClure RS, Hill EA, et al. Unlocking the constraints of cyanobacterial productivity: acclimations enabling ultrafast growth. mBio. 2016;7(4):e00949.
  • Kindle KL. High-frequency nuclear transformation of Chlamydomonas reinhardtii. Proc Natl Acad Sci USA. 1990;87(3):1228–1232.
  • Feng S, Xue L, Liu H, Lu P. Improvement of efficiency of genetic transformation for Dunaliella salina by glass beads method. Mol Biol Rep. 2009;36(6):1433–1439.
  • Cui Y, Wang J, Jiang P, et al. Transformation of Platymonas (Tetraselmis) subcordiformis (Prasinophyceae, Chlorophyta) by agitation with glass beads. World J Microbiol Biotechnol. 2010;26(9):1653–1657.
  • Kumar SV, Misquitta RW, Reddy VS, et al. Genetic transformation of the green alga: Chlamydomonas reinhardtii by Agrobacterium tumefaciens. Plant Sci. 2004;166(3):731–738.
  • Kathiresan S, Sarada R. Towards genetic improvement of commercially important microalga Haematococcus pluvialis for biotech applications. J Appl Phycol. 2009;21(5):553–558.
  • Ortiz-Matamoros MF, Islas-Flores T, Voigt B, et al. Heterologous DNA uptake in cultured Symbiodinium spp. aided by Agrobacterium tumefaciens. PLoS One. 2015;10(7):e0132693.
  • Prasad B, Vadakedath N, Jeong H-J, et al. Agrobacterium tumefaciens-mediated genetic transformation of haptophytes (Isochrysis species). Appl Microbiol Biotechnol. 2014;98(20):8629–8639.
  • Cheng R, Ma R, Li K, et al. Agrobacterium tumefaciens mediated transformation of marine microalgae Schizochytrium. Microbiol Res. 2012;167(3):179–186.
  • Cha T, Chen C, Yee W, et al. Cinnamic acid, coumarin and vanillin: alternative phenolic compounds for efficient Agrobacterium-mediated transformation of the unicellular green alga, Nannochloropsis sp. J Microbiol Methods. 2011;84(3):430–434.
  • Anila N, Chandrashekar A, Ravishankar GA, et al. Establishment of Agrobacterium tumefaciens-mediated genetic transformation in Dunaliella bardawil. Euro J Phycol. 2011;46(1):36–44.
  • Srinivasan R, Gothandam KM. Synergistic action of D-glucose and acetosyringone on Agrobacterium strains for efficient Dunaliella transformation. PLoS One. 2016;11(6):e0158322.
  • Steinbrenner J, Sandmann G. Transformation of the green alga Haematococcus pluvialis with a phytoene desaturase for accelerated astaxanthin biosynthesis. Appl Environ Microbiol. 2006;72(12):7477–7484.
  • Apt KE, Kroth-Pancic PG, Grossman AR. Stable nuclear transformation of the diatom Phaeodactylum tricornutum. Mol Gen Genet. 1996;252(5):572–579.
  • Poulsen N, Chesley PM, Kroger N. Molecular genetic manipulation of the diatom Thalassiosira pseudonana (Bacillariophyceae). J Phycol. 2006;42(5):1059–1065.
  • Dunahay TG, Jarvis EE, Roessler PG. Genetic transformation of the diatoms Cyclotella cryptica and Navicula saprophila. J Phycol. 1995;31(6):1004–1012.
  • Miyagawa-Yamaguchi A, Okami T, Kira N, et al. Stable nuclear transformation of the diatom Chaetoceros sp. Phycol Res. 2011;59(2):113–119.
  • Fischer H, Robl I, Sumper M, et al. Targeting and covalent modification of cell wall and membrane proteins heterologously expressed in the diatom Cylindrotheca fusiformis (Bacillariophyceae). J Phycol. 1999;35(1):113–120.
  • Brown LE, Sprecher SL, Keller LR. Introduction of exogenous DNA into Chlamydomonas reinhardtii by electroporation. Mol Cell Biol. 1991;11(4):2328–2332.
  • Guo SL, Zhao XQ, Tang Y, et al. Establishment of an efficient genetic transformation system in Scenedesmus obliquus. J Biotechnol. 2013;163(1):61–68.
  • Jaeger D, Hübner W, Huser T, et al. Nuclear transformation and functional gene expression in the oleaginous microalga Monoraphidium neglectum. J Biotechnology. 2017;249:10–15.
  • Kilian O, Benemann CSE, Niyogi KK, et al. High-efficiency homologous recombination in the oil-producing alga Nannochloropsis sp. Proc Natl Acad Sci USA. 2011;108(52):21265–21269.
  • Niu YF, Yang ZK, Zhang MH, et al. Transformation of diatom Phaeodactylum tricornutum by electroporation and establishment of inducible selection marker. Biotech. 2012;52:1–3.
  • Run C, Fang L, Fan J, et al. Stable nuclear transformation of the industrial alga Chlorella pyrenoidosa. Algal Res. 2016;17:196–201.
  • Sun Y, Yang Z, Gao X, et al. Expression of foreign genes in Dunaliella by electroporation. Mol Biotechnol. 2005;30(3):185–192.
  • Demirer GS, Zhang H, Matos JL, et al. High aspect ratio nanomaterials enable delivery of functional genetic material without DNA integration in mature plants. Nat Nanotechnol. 2019;14(5):456–464.
  • Kwak SY, Lew TTS, Sweeney CJ, et al. Chloroplast-selective gene delivery and expression in planta using chitosan-complexed single-walled carbon nanotube carriers. Nat Nanotechnol. 2019;14(5):447–455.
  • Qin S, Lin H, Jiang P. Advances in genetic engineering of marine algae. Biotechnol Adv. 2012;30(6):1602–1613.
  • Matsunaga T, Takeyama H. Genetic engineering in marine cyanobacteria. J Appl Phycol. 1995;7(1):77–84.
  • Matsunaga T, Sakaguchi T, Tadakoro F. Magnetite formation by a magnetic bacterium capable of growing aerobically. Appl Microbiol Biotechnol. 1991;35(5):651–655.
  • Qu B, Eu Y, Jeong W, et al. Droplet electroporation in microfluidics for efficient cell transformation with or without cell wall removal. Lab Chip. 2012;12(21):4483–4488.
  • Im DJ, Jeong SN, Yoo BS, et al. Digital microfluidic approach for efficient electroporation with high productivity: transgene expression of microalgae without cell wall removal. Anal Chem. 2015;87(13):6592–6599.
  • Bae S, Park S, Kim J, et al. Exogenous gene integration for microalgal cell transformation using a nanowire-incorporated microdevice. ACS Appl Mater Interfaces. 2015;7(49):27554–27561.
  • Kim HS, Devarenne TP, Han A. Microfluidic systems for microalgal biotechnology: a review. Algal Res. 2018;30:149–161.
  • Bock R. Engineering plastid genomes: methods, tools, and applications in basic research and biotechnology. Annu Rev Plant Biol. 2015;66(1):211–241.
  • Yadav VG, De Mey M, Giaw Lim C, et al. The future of metabolic engineering and synthetic biology: towards a systematic practice. Metab Eng. 2012;14(3):233–241.
  • Doron L, Segal N, Shapira M. Transgene expression in microalgae—from tools to applications. Front Plant Sci. 2016;7:505.
  • Jiang W, Brueggeman AJ, Horken KM, et al. Successful transient expression of Cas9 and single guide RNA genes in Chlamydomonas reinhardtii. Eukaryot Cell. 2014;13(11):1465–1469.
  • Schroda M, Blocker D, Beck CF. The HSP70A promoter as a tool for the improved expression of transgenes in Chlamydomonas. Plant J. 2000;21(2):121–131.
  • Davies JP, Weeks DP, Grossman AR. Expression of the arylsulfatase gene from the beta 2-tubulin promoter in Chlamydomonas reinhardtii. Nucleic Acids Res. 1992;20(12):2959–2965.
  • Barahimipour R, Strenkert D, Neupert J, et al. Dissecting the contributions of GC content and codon usage to gene expression in the model alga Chlamydomonas reinhardtii. Plant J. 2015;84(4):704–717.
  • Mini P, Demurtas OC, Valentini S, et al. Agrobacterium-mediated and electroporation-mediated transformation of Chlamydomonas reinhardtii: a comparative study. BMC Biotechnol. 2018;18(1):11.
  • Shimogawara K, Fujiwara S, Grossman AR, et al. High-efficiency transformation of Chlamydomonas reinhardtii by electroporation. Genetics. 1998;148(4):1821–1828.
  • Dunahay TG. Transformation of Chlamydomonas reinhardtii with silicon carbide whiskers. BioTechniques. 1993;15(3):452–460.
  • Lumbreras V, Stevens DR, Purton S. Efficient foreign gene expression in Chlamydomonas reinhardtii mediated by an endogenous intron. Plant J. 1998;14(4):441–447.
  • Rasala BA, Chao S, Pier M, et al. Enhanced genetic tools for engineering multigene traits into green algae. PLoS One. 2014;9(4):e94028.
  • Rasala BA, Lee PA, Shen Z, et al. Robust expression and secretion of xylanase1 in Chlamydomonas reinhardtii by fusion to a selection gene and processing with the FMDV 2A peptide. PLoS One. 2012;7(8):e43349.
  • Iddo W, Atar S, Schweitzer S, et al. Enhancing heterologous expression in Chlamydomonas reinhardtii by transcript sequence optimization. Plant J. 2018;94:22–31.
  • Eichler-Stahlberg A, Weisheit W, Ruecker O, et al. Strategies to facilitate transgene expression in Chlamydomonas reinhardtii. Planta. 2009;229(4):873–883.
  • Fan J, Ning K, Zeng X, et al. Genomic foundation of starch-to-lipid switch in oleaginous Chlorella spp. Plant Physiol. 2015;169:2444–2461.
  • Chungjatupornchai W, Kitraksa P, Fa-Aroonsawat S. Stable nuclear transformation of the oleaginous microalga Neochloris oleoabundans by electroporation. J Appl Phycol. 2016;28(1):191–199.
  • Muñoz CF, Weusthuis RA, D’Adamo S, et al. Effect of single and combined expression of lysophosphatidic acid acyltransferase, glycerol-3-phosphate acyltransferase, and diacylglycerol acyltransferase on lipid accumulation and composition in Neochloris oleoabundans. Front Plant Sci. 2019;10:1573.
  • Zaslavskaia LA, Lippmeier JC, Kroth PG, et al. Transformation of the diatom Phaeodactylum tricornutum (Bacillariophyceae) with a variety of selectable marker and reporter genes. J Phycology. 2001;36(2):379–386.
  • Zhang C, Hu H. High-efficiency nuclear transformation of the diatom Phaeodactylum tricornutum by electroporation. Mar Genomics. 2014;16:63–66.
  • Cui Y, Zhao J, Wang Y, et al. Characterization and engineering of a dual-function diacylglycerol acyltransferase in the oleaginous marine diatom Phaeodactylum tricornutum. Biotechnol Biofuels. 2018;11:32.
  • Rosenwasser S, Graff van Creveld S, Schatz D, et al. Mapping the diatom redox-sensitive proteome provides insight into response to nitrogen stress in the marine environment. Proc Natl Acad Sci USA. 2014;111(7):2740–2745.
  • Adler-Agnon Z, Leu S, Zarka A, et al. Novel promoters for constitutive and inducible expression of transgenes in the diatom Phaeodactylum tricornutum under varied nitrate availability. J Appl Phycol. 2018;30(5):2763–2710.
  • Li F, Gao D, Hu H. High-efficiency nuclear transformation of the oleaginous marine Nannochloropsis species using PCR product. Biosci Biotechnol Biochem. 2014;78(5):812–817.
  • Radakovits R, Jinkerson RE, Fuerstenberg SI, et al. Draft genome sequence and genetic transformation of the oleaginous alga Nannochloropis gaditana. Nat Commun. 2012;3:686.
  • Dolch L-J, Rak C, Perin G, et al. A palmitic acid elongase affects eicosapentaenoic acid and plastidial monogalactosyldiacylglycerol levels in Nannochloropsis. Plant Physiol. 2017;173(1):742–759.
  • Vieler A, Wu G, Tsai CH, et al. Genome, functional gene annotation, and nuclear transformation of the heterokont oleaginous alga Nannochloropsis oceanica CCMP1779. PLoS Genet. 2012;8:e1003064.
  • Wang Q, Lu Y, Xin Y, et al. Genome editing of model oleaginous microalgae Nannochloropsis spp. by CRISPR/Cas9. Plant J. 2016;88(6):1071–1081.
  • Xin Y, Lu Y, Lee YY, et al. Producing designer oils in industrial microalgae by rational modulation of co-evolving type-2 diacylglycerol Acyltransferases. Mol Plant. 2017;10(12):1523–1539.
  • Zienkiewicz K, Zienkiewicz A, Poliner E, et al. Nannochloropsis, a rich source of diacylglycerol acyltransferases for engineering of triacylglycerol content in different hosts. Biotechnol Biofuels. 2017;10:8.
  • Kaye Y, Grundman O, Leu S, et al. Metabolic engineering toward enhanced LC-PUFA biosynthesis in Nannochloropsis oceanica: overexpression of endogenous Δ12 desaturase driven by stress-inducible promoter leads to enhanced deposition of polyunsaturated fatty acids in TAG. Algal Res. 2015;11:387–398.
  • Nobusawa T, Hori K, Mori H, et al. Differently localized lysophosphatidic acid acyltransferases crucial for triacylglycerol biosynthesis in the oleaginous alga Nannochloropsis. Plant J. 2017;90(3):547–559.
  • Kang NK, Jeon S, Kwon S, et al. Effects of overexpression of a bHLH transcription factor on biomass and lipid production in Nannochloropsis salina. Biotechnol Biofuels. 2015;8:200.
  • Ma X, Yao L, Yang B, et al. RNAi-mediated silencing of a pyruvate dehydrogenase kinase enhances triacylglycerol biosynthesis in the oleaginous marine alga Nannochloropsis salina. Sci Rep. 2017;7(1):11485.
  • Boynton JE, Gillham NW, Harris EH, et al. Chloroplast transformation in Chlamydomonas with high velocity microprojectiles. Science. 1988;240(4858):1534–1538.
  • Rasala BA, Muto M, Sullivan J, et al. Improved heterologous protein expression in the chloroplast of Chlamydomonas reinhardtii through promoter and 5' untranslated region optimization. Plant Biotechnol J. 2011;9(6):674–683.
  • Young REB, Purton S. Codon reassignment to facilitate genetic engineering and biocontainment in the chloroplast of Chlamydomonas reinhardtii. Plant Biotechnol J. 2016;14(5):1251–1260.
  • Specht EA, Mayfield SP. Synthetic oligonucleotide libraries reveal novel regulatory elements in Chlamydomonas chloroplast mRNAs. ACS Synth Biol. 2013;2(1):34–46.
  • O’Neill BM, Mikkelson KL, Gutierrez NM, et al. An exogenous chloroplast genome for complex sequence manipulation in algae. Nucleic Acids Res. 2012;40(6):2782–2792.
  • Lapidot M, Raveh D, Sivan A, et al. Stable chloroplast transformation of the unicellular red alga Porphyridium species. Plant Physiol. 2002;129(1):7–12.
  • Gan Q, Jiang J, Han X, et al. Engineering the chloroplast genome of oleaginous marine microalga Nannochloropsis oceanica. Front Plant Sci. 2018;9:439.
  • Xie W-H, Zhu C-C, Zhang N-S, et al. Construction of novel chloroplast expression vector and development of an efficient transformation system for the diatom Phaeodactylum tricornutum. Mar Biotechnol. 2014;16(5):538–546.
  • Gimpel JA, Vitalia H, Mayfield SP. In metabolic engineering of eukaryotic microalgae: potential and challenges come with great diversity. Front Microbiol. 2015;6:1376.
  • Davies JP, Weeks DP, Grossman AR. Expression of the arylsulfatase gene from the β2-tubulin promoter in Chlamydomonas reinhardtii. Nucl Acids Res. 1992;20(12):2959–2965.
  • Wei Li, Xin Yi, Wang Q, et al. RNAi-based targeted gene knockdown in the model oleaginous microalgae Nannochloropsis oceanica. Plant J. 2017;89(6):1236–1250.
  • Belliveau NM, Barnes SL, Ireland WT, et al. Systematic approach for dissecting the molecular mechanisms of transcriptional regulation in bacteria. Proc Natl Acad Sci USA. 2018;115(21):E4796–E4805.
  • Zhang C, Hu H. High-efficiency nuclear transformation of the diatom Phaeodactylum tricornutum by electroporation. Marine Gen. 2014;16:63–66.
  • Mehrotra R, Renganaath K, Kanodia H, et al. Towards combinatorial transcriptional engineering. Biotechnol Adv. 2017;35(3):390–405.
  • DeHoff P, Soriaga L. Nannochloropsis kozak consensus sequence. US Patent. 2016;9:523.
  • Rosenwasser S, Graff van Creveld S, Schatz D, et al. Mapping the diatom redox-sensitive proteome provides insight into response to nitrogen stress in the marine environment. Proc Nat Acad Sci. 2014;111(7):2740–2745.
  • Salama ES, Govindwar SP, Khandare RV, et al. Can omics approaches improve microalgal biofuels under abiotic stress? Trends Plant Sci. 2019;24(7):611–624.
  • Vila M, Díaz-Santos E, de la Vega M, et al. Promoter trapping in microalgae using the antibiotic paromomycin as selective agent. Marine Drugs. 2012;10(12):2749–2765.
  • Anderson M, Muff TJ, Georgianna DR, et al. Towards a synthetic nuclear transcription system in green algae: characterization of Chlamydomonas reinhardtii nuclear transcription factors and identification of targeted promoters. Algal Res. 2017;22:47–55.
  • Scranton MA, Ostrand JT, Georgianna DR, et al. Synthetic promoters capable of driving robust nuclear gene expression in the green alga Chlamydomonas reinhardtii. Algal Res. 2016;15:135–142.
  • Neupert J, Karcher D, Bock R. Generation of Chlamydomonas strains that efficiently express nuclear transgenes. Plant J. 2009;57(6):1140–1150.
  • Jia B, Zheng Y, Xiao K, et al. A vector for multiple gene co-expression in Chlamydomonas reinhardtii. Algal Res. 2016;20:53–56.
  • Plucinak TM, Horken KM, Jiang W, et al. Improved and versatile viral 2A platforms for dependable and inducible high-level expression of dicistronic nuclear genes in Chlamydomonas reinhardtii. Plant J. 2015;82(4):717–729.
  • Moog D, Schmitt J, Senger J, et al. Using a marine microalga as a chassis for polyethylene terephthalate (PET) degradation. Microb Cell Fact. 2019;18(1):171.
  • Ramos-Martinez EM, Fimognari L, Sakuragi Y. High-yield secretion of recombinant proteins from the microalga Chlamydomonas reinhardtii. Plant Biotechnol J. 2017;15(9):1214–1224.
  • Parinov S, Sundaresan V. Functional genomics in Arabidopsis: large-scale insertional mutagenesis complements the genome sequencing project. Curr OpBiotechnol. 2000;11(2):157–161.
  • Dent RM, Sharifi MN, Malnoë A, et al. Large‐scale insertional mutagenesis of Chlamydomonas supports phylogenomic functional prediction of photosynthetic genes and analysis of classical acetate‐requiring mutants. Plant J. 2015;82(2):337–351.
  • Li X, Patena W, Fauser F, et al. A genome-wide algal mutant library and functional screen identifies genes required for eukaryotic photosynthesis. Nat Genet. 2019 Apr;51(4):627–635.
  • Li Xiaobo, Zhang Ru, Patena W, et al. An indexed, mapped mutant library enables reverse genetics studies of biological processes in Chlamydomonas reinhardtii. Plant Cell. 2016;28(2):367–387.
  • Bonente G, Formighieri C, Mantelli M, et al. Mutagenesis and phenotypic selection as a strategy toward domestication of Chlamydomonas reinhardtii strains for improved performance in photobioreactors. Photosynth Res. 2011;108(2–3):107–120.
  • Mackinder LCM, Chen C, Leib RD, et al. A spatial interactome reveals the protein organization of the algal CO2-concentrating mechanism. Cell. 2017;171(1):133–147.
  • Freeman Rosenzweig ES, Xu B, Kuhn Cuellar L, et al. The eukaryotic CO2-concentrating organelle is liquid-like and exhibits dynamic reorganization. Cell. 2017;171(1):148–162.
  • Cerutti H, Ma X, Msanne J, et al. RNA-mediated silencing in algae: biological roles and tools for analysis of gene function. Eukaryot Cell. 2011;10(9):1164–1172.
  • Taylor RS, Tarver JE, Hiscock SJ, et al. Evolutionary history of plant microRNAs. Trends Plant Sci. 2014;19(3):175–182.
  • Zhao T, Li G, Mi S, et al. A complex system of small RNAs in the unicellular green alga Chlamydomonas reinhardtii. Genes Develop. 2007;21(10):1190–1203.
  • Tarver JE, Cormier A, Pinzón N, et al. MicroRNAs and the evolution of complex multicellularity: identification of a large, diverse complement of microRNAs in the brown alga Ectocarpus. Nucleic Acids Res. 2015;43(13):6384–6398.
  • Huang A, He L, Wang G. Identification and characterization of microRNAs from Phaeodactylum tricornutum by high-throughput sequencing and bioinformatics analysis. BMC Genomics. 2011;12:337.
  • Gao F, Nan F, Feng J, et al. Identification of conserved and novel microRNAs in Porphyridium purpureum via deep sequencing and bioinformatics. BMC Genomics. 2016;17(1):612.
  • Geng H, Sui Z, Zhang S, et al. Identification of microRNAs in the toxigenic dinoflagellate Alexandrium catenella by high-throughput illumina sequencing and bioinformatic analysis. Plos One. 2015;10(9):e0138709.
  • Molnar A, Bassett A, Thuenemann E, et al. Highly specific gene silencing by artificial microRNAs in the unicellular alga Chlamydomonas reinhardtii. Plant J. 2009;58(1):165–174.
  • De Riso V, Raniello R, Maumus F, et al. Gene silencing in the marine diatom Phaeodactylum tricornutum. Nucleic Acids Res. 2009;37( 14):e96.
  • Hannon GJ. RNA interference. Nature. 2002;418(6894):244–251.
  • Fujiwara T, Ohnuma M, Yoshida M, et al. Gene targeting in the red alga Cyanidioschyzon merolae: single- and multi-copy insertion using authentic and chimeric selection markers. PLoS One. 2013;8(9):e73608.
  • Sodeinde OA, Kindle KL. Homologous recombination in the nuclear genome of Chlamydomonas reinhardtii. Proc Natl Acad Sci USA. 1993;90(19):9199–9203.
  • Jeon S, Lim J-M, Lee H-G, et al. Current status and perspectives of genome editing technology for microalgae. Biotechnol Biofuels. 2017;10:267.
  • Greiner A, Kelterborn S, Evers H, et al. Targeting of photoreceptor genes in Chlamydomonas reinhardtii via zinc-finger nucleases and CRISPR/Cas9. Plant Cell. 2017;29(10):2498–2518.
  • Daboussi F, Leduc S, Maréchal A, et al. Genome engineering empowers the diatom Phaeodactylum tricornutum for biotechnology. Nat Commun. 2014;5(1):3831.
  • Gupta SK, Shukla P. Gene editing for cell engineering: trends and applications. Crit Rev Biotechnol. 2017;37(5):672–684.
  • Serif M, Dubois G, Finoux A-L, et al. One-step generation of multiple gene knock-outs in the diatom Phaeodactylum tricornutum by DNA-free genome editing. Nat Commun. 2018;9(1):3924.
  • Meng X, Yu H, Zhang Y, et al. Construction of a genome-wide mutant library in rice using CRISPR/Cas9. Mol Plant. 2017;10(9):1238–1241.
  • Shin S-E, Lim J-M, Koh HG, et al. CRISPR/Cas9-induced knockout and knock-in mutations in Chlamydomonas reinhardtii. Sci Rep. 2016;6:27810.
  • Kao P, Ng I. CRISPRi mediated phosphoenolpyruvate carboxylase regulation to enhance the production of lipid in Chlamydomonas reinhardtii. Biores Technol. 2017;245:1527–1537.
  • Chavez A, Scheiman J, Vora S, et al. Highly efficient Cas9-mediated transcriptional programming. Nat Methods. 2015;12(4):326–328.
  • Nymark M, Sharma A, Sparstad T, et al. A CRISPR/Cas9 system adapted for gene editing in marine algae. Sci Rep. 2016;6:24951.
  • Jeong B-r, Wu-Scharf D, Zhang C, et al. Suppressors of transcriptional transgenic silencing in Chlamydomonas are sensitive to DNA-damaging agents and reactivate transposable elements. Proc Natl Acad Sci USA. 2002;99(2):1076–1081.
  • Daniell H, Lin C, Yu M, et al. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biol. 2016;17(1):134.
  • Wang D, Ning K, Li J, et al. Nannochloropsis genomes reveal evolution of microalgal oleaginous traits. PLoS Genet. 2014;10(1):e1004094.
  • Lu Y, Tarkowská D, Turečková V, et al. Antagonistic roles of abscisic acid and cytokinin during response to nitrogen depletion in oleaginous microalga Nannochloropsis oceanica expand the evolutionary breadth of phytohormone function. Plant J. 2014;80(1):52–68.
  • Lu Y, Xu J. Phytohormones in microalgae: a new opportunity for microalgal biotechnology? Trends Plant Sci. 2015;20(5):273–282.
  • Leister D. Chloroplast research in the genomic age. Trends Genetics. 2003;19(1):47–56.
  • Zhang B, Shanmugaraj B, Daniell H. Expression and functional evaluation of biopharmaceuticals made in plant chloroplasts. Curr Op Chem Biol. 2017;38:17–23.
  • Scotti N, Rigano MM, Cardi T. Production of foreign proteins using plastid transformation. Biotechnol Adv. 2012;30(2):387–397.
  • Wilson SA, Roberts SC. Recent advances towards development and commercialization of plant cell culture processes for the synthesis of biomolecules. Plant Biotechnology J. 2012;10(3):249–268.
  • Dyo YM, Purton S. The algal chloroplast as a synthetic biology platform for production of therapeutic proteins. Microbiology. 2018;164(2):113–121.
  • Golds TJ, Maliga P, Koop H. Stable plastid transformation in PEG-treated protoplasts of Nicotiana tabacum. Nat Biotechnol. 1993;11(1):95–97.
  • Maliga P. Plastid transformation in higher plants. Annu Rev Plant Biol. 2004;55:289–313.
  • O’Neill BM, Mikkelson KL, Gutierrez NM, et al. An exogenous chloroplast genome for complex sequence manipulation in algae. Nucleic Acids Res. 2012;40(6):2782–2792.
  • Qaim M. The economics of genetically modified crops. Annu Rev Resour Econ. 2009;1(1):665–694.
  • Li Y, Peng Y, Hallerman EM, et al. Biosafety management and commercial use of genetically modified crops in China. Plant Cell Rep. 2014;33(4):565–573.
  • Wasmer M. Roads forward for European GMO policy-uncertainties in wake of ECJ judgment have to be mitigated by regulatory reform. Front Bioeng Biotechnol. 2019;7:132.
  • Hsu-Ching C, Anastasios M. Marker-free genetic engineering of the chloroplast in the green microalga Chlamydomonas reinhardtii. Plant Biotechnology J. 2013;11:818–828.
  • Cheah YE, Albers SC, Peebles CAM. A novel counter-selection method for markerless genetic modification in Synechocystis sp. PCC 6803. Biotechnol Prog. 2013;29(1):23–30.
  • Karas BJ, Diner RE, Lefebvre SC, et al. Designer diatom episomes delivered by bacterial conjugation. Nat Commun. 2015;6:6925.
  • Slattery SS, Diamond A, Wang H, et al. An expanded plasmid-based genetic toolbox enables Cas9 genome editing and stable maintenance of synthetic pathways in Phaeodactylum tricornutum. ACS Synth Biol. 2018;7(2):328–338.
  • Poliner E, Takeuchi T, Du ZY, et al. Nontransgenic marker-free gene disruption by an episomal CRISPR system in the oleaginous microalga, Nannochloropsis oceanica CCMP1779. ACS Synth Biol. 2018;7(4):962–968.
  • Nora LC, Westmann CA, Guazzaroni ME, et al. Recent advances in plasmid-based tools for establishing novel microbial chassis. Biotechnology Advances. 2019;37(8):107433.
  • Bailey JE. Toward a science of metabolic engineering. Science. 1991;252(5013):1668–1675.
  • Trentacoste EM, Shrestha RP, Smith SR, et al. Metabolic engineering of lipid catabolism increases microalgal lipid accumulation without compromising growth. Proc Natl Acad Sci USA. 2013;110(49):19748–19753.
  • Han X, Song X, Li F, et al. Improving lipid productivity by engineering a control-knob gene in the oleaginous microalga Nannochloropsis oceanica. Metabolic Eng Commun. 2020;11:e00142.
  • Loera-Quezada MM, Leyva-González MA, Velázquez-Juárez G, et al. A novel genetic engineering platform for the effective management of biological contaminants for the production of microalgae. Plant Biotechnol J. 2016;14:2066–2076.
  • Lu Y, Gan Q, Iwai M, et al. Role of an ancient light-harvesting protein of PSI in light absorption and photoprotection. Nat Commun. 2021;12(1):679.
  • Darby AC, Hall N. Fast forward genetics. Nat Biotechnol. 2008;26(11):1248–1249.
  • Ngan CY, Wong CH, Choi C, et al. Lineage-specific chromatin signatures reveal a regulator of lipid metabolism in microalgae. Nat Plants. 2015;1:15107.
  • Chen X, Hu G, Liu L. Hacking an algal transcription factor for lipid biosynthesis. Trends Plant Sci. 2018;23(3):181–184.
  • Zhang J, Hao Q, Bai L, et al. Overexpression of the soybean transcription factor GmDof4 significantly enhances the lipid content of Chlorella ellipsoidea. Biotechnol Biofuels. 2014;7(1):128.
  • Thiriet-Rupert S, Carrier G, Chénais B, et al. Transcription factors in microalgae: genome-wide prediction and comparative analysis. BMC Genomics. 2016;17:282.
  • Chen W, Kallio PT, Bailey JE. Construction and characterization of a novel cross-regulation system for regulating cloned gene expression in Escherichia coli. Gene. 1993;130(1):15–22.
  • Gardner TS, Cantor CR, Collins JJ. Construction of a genetic toggle switch in Escherichia coli. Nature. 2000;403(6767):339–342.
  • Kemmer C, Gitzinger M, Daoud-El Baba M, et al. Self-sufficient control of urate homeostasis in mice by a synthetic circuit. Nat Biotechnol. 2010;28(4):355–360.
  • Ignea C, Athanasakoglou A, Ioannou E, et al. Carnosic acid biosynthesis elucidated by a synthetic biology platform. Proc Natl Acad Sci USA. 2016;113(13):3681–3686.
  • Chen B, Lee HL, Heng YC, et al. Synthetic biology toolkits and applications in Saccharomyces cerevisiae. Biotechnol Adv. 2018;36(7):1870–1881.
  • Hlavova M, Turoczy Z, Bisova K. Improving microalgae for biotechnology—from genetics to synthetic biology. Biotechnol Adv. 2015;33(6):1194–1203.
  • Butler TO, Kapoore RV, Vaidyanathan S. Phaeodactylum tricornutum: a diatom cell factory. Trends Biotechnol. 2020;38(6):606–622.
  • Lauersen KJ, Kruse O, Mussgnug JH. Targeted expression of nuclear transgenes in Chlamydomonas reinhardtii with a versatile, modular vector toolkit. Appl Microbiol Biotechnol. 2015;99(8):3491–3503.
  • Chen H, Qiu T, Rong J, et al. Microalgal biofuel revisited: an informatics-based analysis of developments to date and future prospects. Appl Energy. 2015;155:585–598.
  • Roth MS, Cokus SJ, Gallaher SD, et al. Chromosome-level genome assembly and transcriptome of the green alga Chromochloris zofingiensis illuminates astaxanthin production. Proc Natl Acad Sci USA. 2017;114(21):E4296–E4305.
  • D’Adamo S, di Visconte GS, Lowe G, et al. Engineering the unicellular alga Phaeodactylum tricornutum for high-value plant triterpenoid production. Plant Biotechnol J. 2019;17(1):75–87.
  • Wichmann J, Baier T, Wentnagel E, et al. Tailored carbon partitioning for phototrophic production of (E)-α-bisabolene from the green microalga Chlamydomonas reinhardtii. Metab Eng. 2018;45:211–222.
  • Lauersen KJ, Baier T, Wichmann J, et al. Efficient phototrophic production of a high-value sesquiterpenoid from the eukaryotic microalga Chlamydomonas reinhardtii. Metab Eng. 2016;38:331–343.
  • Lawson CE, Harcombe WR, Hatzenpichler R, et al. Common principles and best practices for engineering microbiomes. Nat Rev Microbiol. 2019;17(12):725–741.
  • Schulze S, Urzica E, Reijnders MJMF, et al. Identification of methylated GnTI-dependent N-glycans in Botryococcus brauni. New Phytol. 2017;215(4):1361–1369.
  • Baiet B, Burel C, Saint-Jean B, et al. N-glycans of Phaeodactylum tricornutum diatom and functional characterization of its N-acetylglucosaminyltransferase I enzyme. J Biol Chem. 2011;286:6152–6164.
  • Wei L, Shen C, El Hajjami M, et al. Knockdown of carbonate anhydrase elevates Nannochloropsis productivity at high CO2 level. Metab Eng. 2019;54:96–108.
  • Wu W, Wang PH, Lee DJ, et al. Global optimization of microalgae-to-biodiesel chains with integrated cogasification combined cycle systems based on greenhouse gas emissions reductions. Appl Energy. 2017;197:63–82.
  • Cheah WY, Ling TC, Show PL, et al. Cultivation in wastewaters for energy: a microalgae platform. Appl Energy. 2016;179:609–625.
  • Kirst H, Garcia-Cerdan JG, Zurbriggen A, et al. Truncated photosystem chlorophyll antenna size in the green microalga Chlamydomonas reinhardtii upon deletion of the TLA3-CpSRP43 gene. Plant Physiol. 2012;160(4):2251–2260.
  • Loera-Quezada MM, Leyva-González MA, Velázquez-Juárez G, et al. A novel genetic engineering platform for the effective management of biological contaminants for the production of microalgae. Plant Biotechnol J. 2016;14(10):2066–2076.
  • Bruggeman AJ, Kuehler D, Weeks DP. Evaluation of three herbicide resistance genes for use in genetic transformations and for potential crop protection in algae production. Plant Biotechn J. 2014;12(7):894–902.
  • Gan Q, Zhou W, Wang S, et al. A customized contamination controlling approach for culturing oleaginous Nannochloropsis oceanica. Algal Res. 2017;27:376–382.
  • Demuez M, González-Fernández C, Ballesteros M. Algicidal microorganisms and secreted algicides: new tools to induce microalgal cell disruption. Biotechnol Adv. 2015;33:1615–1625.
  • Liu X, Fallon S, Sheng J, et al. CO2-limitation-inducible Green Recovery of fatty acids from cyanobacterial biomass. Proc Natl Acad Sci USA. 2011;108(17):6905–6908.