Synthetic Biology of Yeasts Tools and Applications (2022)
Synthetic Biology of Yeasts Tools and Applications (2022)
Synthetic Biology of Yeasts Tools and Applications (2022)
Darvishi Harzevili Editor
Synthetic
Biology
of Yeasts
Tools and Applications
Synthetic Biology of Yeasts
Farshad Darvishi Harzevili
Editor
Synthetic Biology
of Yeasts
Tools and Applications
Editor
Farshad Darvishi Harzevili
Department of Microbiology
Alzahra University
Tehran, Iran
© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Switzerland AG 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface
Yeasts are one of the most attractive microbial cell factories for the production
of a wide range of valuable products, including pharmaceuticals, nutraceuticals,
cosmetics, agrochemicals, biofuels, and so on. Tools of synthetic biology have
been developed to improve the metabolic engineering of yeasts in a faster and
more reliable manner. Nowadays, synthetic biology tools have been used to make
synthetic pathways and rewiring metabolism to produce valuable products at high
titer, rate, and yield.
This book covers recent advances and future trends in the synthetic biology
of yeasts. The book consists of two parts. The first part opens with an intro-
duction to rational metabolic pathway prediction and design using computational
tools and their applications for yeast systems and synthetic biology (Chap. “Ratio-
nal Metabolic Pathway Prediction and Design: Computational Tools and Their
Applications for Yeast Systems and Synthetic Biology”). After rational metabolic
pathway design, synthesis, and engineering tools for assembly of standardized bio-
bricks, powerful genome editing and optimization of synthetic pathways are very
important in metabolic engineering of yeasts. Hence, construction and assem-
bly of standardized biobricks for synthetic pathways engineering in yeasts were
summarized in Chap. “Construction and Assembly of Standardized Biobricks
for Synthetic Pathways Engineering in Yeasts”.
Chapter “Cellular Engineering of Yarrowia lipolytica for Biomanufacturing
of High-Value Products from Oils and Fats” presented an example about cellular
engineering of yeasts. A combined cell morphology engineering and metabolic
pathway optimization of non-conventional yeast Yarrowia lipolytica is used to
biomanufacturing of high-value products from oils and fats as hydrophobic sub-
strates. Chapter “Whole Cell Yeast-Based Biosensors” provided an overview about
engineering of yeast cells to act as whole-cell yeast-based biosensors. The highly
modular nature of yeast-based biosensors suggests that they can be a useful tool for
detecting a wide range of biologically relevant analytes, including pharmaceuticals,
toxins, and chemically undefined or complex mixtures ranging from environmental
mutagens to bioavailable phosphorous.
The second part, Chaps. “Yeast Synthetic Biology Approaches for the Produc-
tion of Valuable Polyphenolic Compounds”–“Synthetic Biology in the Candida
v
vi Preface
vii
viii Contents
Pedro A. Saa
Abstract
Continuous progress in metabolic engineering of microbial cell factories like
yeast requires the support of computational tools for finding novel unintuitive
biotransformations routes. In this chapter, a succinct overview is provided of the
most relevant computational tools for pathway prediction by retro-biosynthesis,
and pathway design through stoichiometry-based optimization methods. Illus-
trative case studies are also presented showcasing different strategies for
pathway optimization in yeast, namely redox cofactor balancing, improved pre-
cursor supply, and heterologous expression of carbon fixation pathways. Finally,
challenges and limitations hindering the broad adoption and implementation of
these tools for metabolic engineering will be discussed.
Keywords
1 Introduction
P. A. Saa (B)
Departament of Chemical and Bioprocess Engineering, Pontificia Universidad Católica de Chile,
Santiago, Chile
e-mail: [email protected]
Institute for Mathematical and Computational Engineering, Pontificia Universidad Católica de
Chile, Santiago, Chile
Fig. 1 Workflow for reaction network reconstruction and application of metabolic pathway pre-
diction and design tools. a Assembly of accumulated metabolic reaction data into a comprehensive
reaction network is a requirement for the application of the reviewed tools. Depending on the
application objective, different network representations are employed for either predicting de novo
pathways (i.e., retrosynthesis typically using a graph representation), or (re)designing pathways
for higher metabolic performance (i.e., optimization-based pathway design using an stoichiometric
representation). b Retro-biosynthetic tools explore a substrate graph seeking to connect the target
molecule with some predefined precursors. Starting with the target molecule and moving back-
wards, these tools can generate several possible pathways that are typically ranked using different
criteria (e.g., length, enzyme availability, thermodynamics, among others). c Stoichiometry-based
pathway prediction methods employ a reaction network with known and fixed reactions to enu-
merate mass-balanced pathways that optimize a desired objective such as product yield, pathway
length, thermodynamic favorability, and enzyme cost. In this case, different types of constraints
can be defined to restrict the feasible solution space and narrow the search upfront
Databases for pathway search are an absolute requirement for exploring the fea-
sible reaction space, as they contain the critical information of how metabolites
are connected to others through biochemical reactions. There are numerous pub-
lic data- and knowledge-bases populated with metabolic reaction data. Among
the most popular, KEGG (Kanehisa et al. 2016), MetaCyc (Caspi et al. 2016),
BIGG (King et al. 2016), KBase (Arkin et al. 2018), ModelSEED (Henry et al.
2010a, b), MetRxn (Kumar et al. 2012), and MetaNetX (Ganter et al. 2013; Moretti
et al. 2021) stand out to name a few (for a more details refer to Wang et al.
(2017)). Some of these databases (KEGG, MetaCyc, and Kbase) integrate multi-
ple sources of biological information, e.g., genetic, molecular, physicochemical,
6 P. A. Saa
and experimental, which makes them not only useful for metabolic pathway pre-
diction purposes but also data integration (Lewis et al. 2012). The rest of the
databases are mostly devoted to metabolic network reconstruction, offering either
highly curated reconstructions for specific organisms (e.g., BIGG) or broader albeit
possibly less curated biochemical reaction networks (e.g., MetRxn, ModelSEED,
and MetaNetX). Ultimately, the modeling purpose will dictate the most convenient
source of information considering their specific scope, breadth, and informa-
tion quality. Complementary databases like BRENDA (Jeske et al. 2019) (kinetic
information) and eQuilibrator (Flamholz et al. 2012) (thermodynamic informa-
tion) also constitute valuable resources for parameterizing different optimization
formulations.
The aforementioned databases contain information for known reactions, which
may restrict the pathway search considering the current enzymatic knowledge
gaps. Resources like the ATLAS of Biochemistry (Hadadi et al. 2016) (derived
from the BNICE tool (Hatzimanikatis et al. 2005)) and MINE (Jeffryes et al.
2015) offer larger networks including hypothetical reactions and metabolites that
can expand the reachable chemical space and allow higher complexity. Briefly,
these resources exploit user-defined reaction rules that can act on chemically sim-
ilar compounds, thereby yielding new hypothetical reactions. The latter reactions
have recently been shown to enable filling some of the gaps in current enzyme-
reaction associations (Hadadi et al. 2019). Lastly, another significant and recent
tool for proposing hypothetical reactions that has been employed for pathway
prediction is rePrime used by the novoStoic tool (Kumar et al. 2018) (see sub-
section 2.3 for more details). The former method extracts reaction rules from
molecular signatures found in annotated reactions—defined by the presence of a
set of chemical ‘moieties’—for proposing hypothetical enzymatic transformations
with a high structural encoding fidelity. Unfortunately, this tool currently lacks an
associated open database for its use.
Table 1 (continued)
Tool Database Pathway scoring Applications and references Source
SimPheny BiGG Pathway length, In silico prediction and -
(BioPathway thermodynamics, subsequent experimental
Predictor) product yield, validation of an
number of known unprecedented synthetic
metabolites/enzymes pathway for 1,4-butanediol
production in E. coli (Yim
et al. 2011)
RetroPath MetaCyc, Pathway Prediction of https://
BioCyc thermodynamics, successfully-implemented www.mye
sequence similarity, biosynthetic pathways for xperiment.
pathway length, 146 compounds in metabolic org/workfl
number of putative engineering projects. 81% of ows/4987.
enzyme steps, and the compounds were html
product yield predicted with at least one
pathway (Delépine et al.
2018)
At the end of the algorithm, pathways are ranked by features such as pathway
length, thermodynamics, among others. Methodologically close to BNICE, Path-
Pred (Moriya et al. 2010) uses instead RDM patterns consisting of reaction center
atoms (R), atoms of different regions (D), and atoms of the matched region (M) for
exploring the substrate graph. Pruning of the network is executed using structural
similarity criteria, and pathway ranking is performed using compound similar-
ity and pathway scores. SimPheny (Yim et al. 2011; Schilling et al. 2005) uses
reactions rules from the third Enzyme Commission (EC) number level for gen-
erating reaction rules that enable reaction promiscuity for broader explorations.
In this case, a retro-synthetic search is employed for enumerating feasible routes
that produced intermediates of reasonable size (i.e., below a predefined size), and
later they are ranked based on various criteria. Another retro-biosynthesis tool
with recent important applications is RetroPath (Delépine et al. 2018). This tool
uses a retro-synthetic search, albeit combined with MILP formulations for the
application of various ranking criteria, such as thermodynamics, gene prediction,
pathway length, number of putative steps, and product yield. In contrast to BNICE,
RetroPath maintains a stoichiometric representation of the network (as opposed to
a substrate graph) that enables computation of various scores. Moreover, molecular
signatures are used to generate reaction rules based on a substructure of adja-
cent atoms, enabling the generation of substantially more and flexible reaction
rules (Duigou et al. 2018). A recent implementation of reinforcement learning in
RetroPath (RetroPath RL) has yielded promising results in the retro-biosynthetic
prediction of biologically relevant pathways (Koch et al. 2020).
Rational Metabolic Pathway Prediction and Design … 9
PathTracer between reactants and products putrescine from glutamate, and CO2 fixation pathways in
using the shortest pathway E. coli (Tervo and Reed 2016)
(continued)
11
12
Table 2 (continued)
Pathway design objective Method Goal(s) Applications and references
Thermodynamic favorability OptMDFpathway Maximization of the minimum Computation of all substrate-product combinations in E.
driving force of the production coli that enable net CO2 assimilation via
pathway thermodynamically feasible pathways (Hädicke et al.
2018)
In silico design of multi-strain communities of a
single-species (E. coli) that achieve maximum
thermodynamic driving force for product synthesis under
the assumption of pathway specialization (i.e., parts of the
production pathway are performed by different community
members) (Bekiaris and Klamt 2021)
Enzymatic cost ECM Minimize enzyme cost or In silico computation of putative efficient CO2 fixation
or burden metabolic burden pathways into C2-C4 compounds (Bar-Even et al. 2010)
Optimality analysis of enzyme cost minimization in central
carbon metabolism of E. coli (Noor et al. 2016)
Prediction of rate/yield trade-offs in metabolic pathways
under oxygen-limited conditions in E. coli (Wortel et al.
2018)
P. A. Saa
Rational Metabolic Pathway Prediction and Design … 13
Regeneration of either redox and/or energy cofactors often limits the production of
high-value metabolites. In order to increase the availability of the required cofac-
tor(s), central carbon metabolism must be intervened and engineered in such a
way that it favors bioproduction without extremely affecting microbial growth (Lee
et al. 2013). This challenge is particularly relevant for many NAD(P)H-expensive
valuable compounds that are being produced in yeast (Cataldo et al. 2020; López
et al. 2020, 2019) and other microbes (Ko et al. 2020).
Increased supply of redox cofactors can be achieved by either overexpressing
key enzymes involved in cofactor generation (Lee et al. 2007; San et al. 2002;
Lim et al. 2002) or by increasing the expression of alternative redox partner
systems. A recent application of the latter has proved effective for enhancing
the unprecedented heterologous production of violaxanthin in S. cerevisiae by
approx. two-fold (Cataldo et al. 2020). However, the success of these approaches
is likely limited due to the presence of different intrinsic balancing mechanisms
for maintaining homeostasis in yeast (Hou et al. 2010). An illustrative example of
the latter can be found in the study of Nissen et al. (2001). Here, heterologous
expression of the pyridine nucleotide transhydrogenase system (sth gene, absent
in yeast) that transfers reducing equivalents from NADPH to NADH (and vice
versa), did not improve ethanol formation in anaerobic conditions. On the contrary,
ethanol production was reduced concomitantly with the increase of fermentation
by-products (glycerol and 2-oxoglutarate) required for redox rebalancing. Another,
less intuitive and possibly more effective, strategy for (re)balancing redox cofac-
tors supply and demand involves cofactor swapping (Verho et al. 2003; Martínez
et al. 2008). Computational studies in S. cerevisiae and E. coli support this strategy
as a promising intervention for forcing higher metabolic performance (King and
Feist 2014). Simply put, this approach seeks to replace native (redox-consuming)
enzymes with heterologous counterparts with a different cofactor specificity (e.g.,
NAD(P)H—for a NA(D)H-dependent enzyme).
The first application of the latter strategy involved the optimization of D-
xylose utilization for ethanol production in S. cerevisiae (Verho et al. 2003).
This carbon source is assimilated through the pentose phosphate pathway (PPP)
as D-xylulose-5-phosphate and then incorporated as glyceraldehyde 3-phosphate
in glycolysis. In theory, D-xylose should produce CO2 and ethanol in a 1:1
molar ratio under redox-neutral anaerobic conditions (Kötter and Ciriacy 1993).
However, D-xylose assimilation requires extra NADPH and NAD+ that must be
regenerated by other separate processes, which are very inefficient in yeast, ren-
dering D-xylose fermentation slow. To overcome this bottleneck and force higher
16 P. A. Saa
Rational Metabolic Pathway Prediction and Design … 17
Fig. 2 Illustration of selected reported strategies for achieving improved cofactor balancing,
increased acetyl-CoA supply and engineering CO2 fixation in yeast. The details of each
strategy are discussed in Sect. 3. Relevant metabolite names are represented by uppercase
bold fonts, whereas enzyme names are represented by uppercase italics fonts. Abbreviations:
13DPG, 1,3-diphosphoglycerate; 3PG, 3-phosphoglycerate; 6PGL, 6-phospho-D-glucono-1,5-
lactone; ACCOA, acetyl-CoA; ACE, acetate; ACETAL, acetaldehyde; AcP, acetyl phosphate;
ACS, acetyl-CoA synthetase; AKG, alpha-ketoglutarate; ALD, aldehyde dehydrogenase; DHAP,
dihydroxyacetone phosphate; ETOH, ethanol; F6P, D-fructose 6-phosphate; FDP, D-fructose 1,6-
disphosphate; FOR, formate; G3P, glycerol 3-phospate; G6P, D-glucose 6-phosphate; G6PD,
glucose 6-phosphate dehydrogenase; GAP, glyceraldehyde 3-phosphate; GAPDH, glyceralde-
hyde 3-phosphate dehydrogenase; GDH, glutamate dehydrogenase; GLC, D-glucose; GLU, glu-
tamate; GLY, glycerol; HMGCOA, 3-hydroxy-3-methyl-glutaryl-CoA; HMGCOAR, HMG-CoA
reductase; MEOH, methanol; MEV, mevalonate; NH4, ammonia; PDH, pyruvate dehydrogenase;
PFL, pyruvate formate lyase; PK, phosphoketolase; PRK, phosphoribulokinase; PTA, phospho-
transacetylase; PYR, pyruvate; R5P, D-ribose 5-phosphate; Ru15P, ribulose 1,5-disphosphate;
Ru5P, D-ribulose 5-phosphate; STH, transhydrogenase; Xu5P, D-xylulose 5-phosphate; XYL, D-
xylose
NADPH supply and flux through lower glycolysis, the native NAD-dependent
GAP dehydrogenase (GAPDH) was replaced by an NADP-dependent GAPDH and
the NADPH-dependent glucose-6-phosphate dehydrogenase (G6PD) was knocked
out, which also prevented carbon loss as CO2 (Verho et al. 2003). This strat-
egy almost doubled the ethanol yield on D-xylose (from 18 to 41%) and reduced
the CO2 /ethanol molar ratio close to the theoretical 1:1 (from 2.5 to 1.3). Later,
expression of the heterologous phosphotransacetylase (PTA) and phosphoketolase
(PK) for improving NADH reoxidation in the D-xylose utilization pathway gen-
erated an increase in ethanol yield (25% higher) without affecting the growth rate
(Sonderegger et al. 2004).
Cofactor rebalancing and swapping strategies for the synthesis of NADPH-
expensive isoprenoid-derived compounds have shown to be particularly effective
in yeast. For instance, α-santalene production yields a net production of NADH
and consumption of NADPH, which calls for the rebalancing of the cofactor sup-
ply (Scalcinati et al. 2012). By deleting known reactions involved in glutamate
metabolism (ammonium assimilation) that consume NADPH (GDH1) and acti-
vating NAD-dependent counterparts (GDH2) (Nissen et al. 2000), the production
of α-santalene was substantially improved. Similarly in a different study of pro-
topanaxadiol production—another isoprenoid-derived compound—the availability
of NADPH was enhanced by replacing the native NADH-generating acetaldehyde
dehydrogenase (ALD2) with a functionally equivalent NADPH-generating enzyme
(ALD6), resulting in a 11-fold increase in titer (Kim et al. 2018). Lastly, swap-
ping of the native NADP-dependent 3-hydroxy-3-methyl-glutaryl-CoA reductase
(HMG-CoA reductase)—third committed step of the mevalonate pathway respon-
sible for the production of isoprenoid precursors—has also shown to increase the
overall pathway flux in E. coli (Ma et al. 2011). This result was leveraged by
Meadows et al. (2016) whereby an NADH-consuming HMG-CoA reductase from
Silicibacter pomeroyi was employed for the overproduction of the sesquiterpene
18 P. A. Saa
There is a growing interest in the field for engineering carbon assimilation path-
ways in heterotrophs for improving product yields—e.g., by reducing carbon loss
as CO2 –, and most notably, for implementing one-carbon (C1) compounds (e.g.,
CO2 ) fixation pathways to develop more sustainable fermentation bioprocesses.
Rational Metabolic Pathway Prediction and Design … 19
During the past decade, yeast metabolic engineering has shown great progress and
promise (Smolke and Tyo 2012; Nielsen and Keasling 2016), quickly becoming
one of the preferred microbial factories for realizing the bioproduction of new
chemicals or improving the production of traditional ones. This success has been
largely driven by the continuous advances in the development of genetic and
molecular tools (Smolke and Tyo 2012), as well as novel computational frame-
works for pathway discovery and optimization (Wang et al. 2017; Saa et al.
2019). The latter has brought not only new possibilities for the evaluation of
novel biochemical synthesis routes but also has provided more rational meth-
ods for designing metabolic pathways with superior performance by rewiring
metabolism at a whole-cell scale (Saa et al. 2019). In time, such capabilities will
become increasingly essential for arriving at designs that scale industrially and
meet commercial expectations.
Pathway discovery is supported by the use of retro-synthesis tools that generate
putative routes connecting substrates to products. A comprehensive exploration of
the chemical space typically rests on the availability of reaction rules, which fills
the gaps between the metabolic precursors and target chemical(s). Generation and
application of such rules must be carefully performed, as they may provide infea-
sible pathways that may obscure results interpretation (Wang et al. 2017). Atom
mapping information can be of great aid for validating the application and gen-
eration of certain reaction rules, see e.g., RouteSearch (Latendresse et al. 2014)
and ReTrace (Pitkänen et al. 2009), which can be further completed with enzyme
promiscuity knowledge if available (Mazurenko et al. 2020). Another incipient
alternative for learning novel chemical reaction routes rests on machine learn-
ing techniques (Koch et al. 2020), which can potentially increase exponentially
the size of the reachable chemical space as shown elsewhere (Coley et al. 2019;
Mikulak-Klucznik et al. 2020). Efficient navigation of such vast space would nec-
essarily have to rely on the introduction of pathway scores and rankings to focus
the attention on the most promising and realizable designs. For this task, eval-
uation of the objectives reviewed here along with others—e.g., use of enzymes
with known promiscuous activity or cofactor specificity—constitutes a natural way
for prioritizing and selecting desired pathways. Rational integration of the vari-
ous objectives can be achieved by leveraging mature multi-decision multi-criteria
techniques (Bonissone et al. 2009), which remains largely unexplored in the field.
Notably, the latter techniques are also transferable to optimization-based methods
for pathway prediction, which could enable a more holistic evaluation of pathway
performance and robustness.
While the revised computational methods and tools for pathway prediction have
provided unintuitive and useful insights, their experimental application and vali-
dation remain still limited. Although there have been recent applications in yeast
(Hafner et al. 2021) and other model organisms (Yim et al. 2011) where some
of the tools have proven to be critical for finding effective in vivo metabolic
Rational Metabolic Pathway Prediction and Design … 21
designs, there is still resistance to their broad adoption. Indeed, in vivo implemen-
tation of complex in silico metabolic designs is not trivial, typically demanding
great amounts of experimentation time before arriving at a working pathway
(Antonovsky et al. 2016; Schwander et al. 2016; Savile et al. 2010). Such efforts
could gain from recent computational frameworks for kinetic model construction
(Saa and Nielsen 2017) that could help to predict a priori the expected performance
of the pathway (see, for example, (Theisen et al. 2016)), greatly reducing the time
and resources needed. As the metabolic prediction capabilities of current models
increase (Foster et al. 2021), it is expected that the use of these tools for ratio-
nal pathway engineering in yeast and other microbial factories will progressively
become part of the basic toolbox for metabolic engineering.
Acknowledgements The author acknowledges the financial support of ANID Fondecyt Iniciacion
Grant No 11190871 and FONDAP Grant No 15090007 of the Center for Genome Regulation (CGR).
References
Alcántara R, Axelsen KB, Morgat A, Belda E et al (2012) Rhea-a manually curated resource of
biochemical reactions. Nucleic Acids Res 40:D754–D760
Antonovsky N, Gleizer S, Noor E, Zohar Y et al (2016) Sugar Synthesis from CO2 in Escherichia
coli. Cell 166:115–125
Arkin AP, Cottingham RW, Henry CS, Harris NL et al (2018) KBase: the united states department
of energy systems biology knowledgebase. Nat Biotechnol 36:566–569
Bang J, Lee SY (2018) Assimilation of formic acid and CO2 by engineered Escherichia coli
equipped with reconstructed one-carbon assimilation pathways. Proc Natl Acad Sci 115:E9271
Bar-Even A, Noor E, Lewis NE, Milo R (2010) Design and analysis of synthetic carbon fixation
pathways. Proc Natl Acad Sci 107:8889
Bar-Even A, Noor E, Savir Y, Liebermeister W et al (2011) The moderately efficient enzyme: evo-
lutionary and physicochemical trends shaping enzyme parameters. Biochemistry-Us 50:4402–
4410
Bekiaris PS, Klamt S (2021) Designing microbial communities to maximize the thermodynamic
driving force for the production of chemicals. Plos Comput Biol 17:e1009093
Bogorad IW, Lin TS, Liao JC (2013) Synthetic non-oxidative glycolysis enables complete carbon
conservation. Nature 502:693–697
Bonissone PP, Subbu R, Lizzi J (2009) Multicriteria decision making (mcdm): a framework for
research and applications. IEEE Comput Intell Mag 4:48–61
Campodonico MA, Andrews BA, Asenjo JA, Palsson BO, Feist AM (2014) Generation of an
atlas for commodity chemical production in Escherichia coli and a novel pathway prediction
algorithm GEM-Path. Metab Eng 25C:140–158
Caspi R, Billington R, Ferrer L, Foerster H et al (2016) The MetaCyc database of metabolic path-
ways and enzymes and the BioCyc collection of pathway/genome databases. Nucleic Acids Res
44:D471-480
Cataldo VF, Arenas N, Salgado V, Camilo C et al (2020) Heterologous production of the epoxy-
carotenoid violaxanthin in Saccharomyces cerevisiae. Metab Eng 59:53–63
Chowdhury A, Maranas CD (2015) Designing overall stoichiometric conversions and intervening
metabolic reactions. Sci Rep 5
Coley CW, Thomas DA, Lummiss JAM, Jaworski JN et al (2019). A robotic platform for flow
synthesis of organic compounds informed by AI planning. Science 365:eaax1566
Court SJ, Waclaw B, Allen RJ (2015) Lower glycolysis carries a higher flux than any biochemically
possible alternative. Nat Commun 6:8427
22 P. A. Saa
Delépine B, Duigou T, Carbonell P, Faulon J-L (2018) RetroPath2.0: a retrosynthesis workflow for
metabolic engineers. Metab Eng 45:158–170
Du B, Zielinski DC, Monk JM, Palsson BO (2018) Thermodynamic favorability and pathway yield
as evolutionary tradeoffs in biosynthetic pathway choice. Proc Natl Acad Sci 115:11339
Duigou T, du Lac M, Carbonell P, Faulon J-L (2018) RetroRules: a database of reaction rules for
engineering biology. Nucleic Acids Res 47:D1229–D1235
Edwards JS, Palsson BO (2000) The Escherichia coli MG1655 in silico metabolic genotype: Its
definition, characteristics, and capabilities. P Natl Acad Sci USA 97:5528–5533
Ellis LBM, Roe D, Wackett LP (2006) The University of Minnesota biocatalysis/biodegradation
database: the first decade. Nucleic Acids Res 34:D517–D521
Finley SD, Broadbelt LJ, Hatzimanikatis V (2009) Computational framework for predictive
biodegradation. Biotechnol Bioeng 104:1086–1097
Finley SD, Broadbelt LJ, Hatzimanikatis V (2010) In silico feasibility of novel biodegradation
pathways for 1,2,4-trichlorobenzene. Bmc Syst Biol 4:7
Flamholz A, Noor E, Bar-Even A, Milo R (2012) eQuilibrator-the biochemical thermodynamics
calculator. Nucleic Acids Res 40:770–775
Flamholz A, Noor E, Bar-Even A, Liebermeister W, Milo R (2013) Glycolytic strategy as a tradeoff
between energy yield and protein cost. Proc Natl Acad Sci 110:10039
Foster CJ, Wang L, Dinh HV, Suthers PF, Maranas CD (2021) Building kinetic models for
metabolic engineering. Curr Opin Biotech 67:35–41
Galanie S, Thodey K, Trenchard IJ, Filsinger Interrante M, Smolke CD (2015) Complete biosyn-
thesis of opioids in yeast. Science 349:1095–1100
Ganter M, Bernard T, Moretti S, Stelling J, Pagni M (2013) MetaNetX.org: a website and repos-
itory for accessing, analysing and manipulating metabolic networks. Bioinformatics 29:815–
816
Gassler T, Sauer M, Gasser B, Egermeier M et al (2020) The industrial yeast Pichia pastoris is
converted from a heterotroph into an autotroph capable of growth on CO2 . Nat Biotechnol
38:210–216
Guadalupe-Medina V, Wisselink HW, Luttik MA, de Hulster E et al (2013) Carbon dioxide fixation
by Calvin-cycle enzymes improves ethanol yield in yeast. Biotechnol Biofuels 6:125
Hadadi N, Hafner J, Shajkofci A, Zisaki A, Hatzimanikatis V (2016) ATLAS of Biochemistry: a
repository of all possible biochemical reactions for synthetic biology and metabolic engineering
studies. ACS Synth Biol 5:1155–1166
Hadadi N, MohammadiPeyhani H, Miskovic L, Seijo M, Hatzimanikatis V (2019) Enzyme annota-
tion for orphan and novel reactions using knowledge of substrate reactive sites. Proc Natl Acad
Sci 116:7298
Hädicke O, von Kamp A, Aydogan T, Klamt S (2018) OptMDFpathway: Identification of
metabolic pathways with maximal thermodynamic driving force and its application for analyz-
ing the endogenous CO2 fixation potential of Escherichia coli. Plos Comput Biol 14:e1006492
Hafner J, Payne J, MohammadiPeyhani H, Hatzimanikatis V, Smolke C (2021) A computational
workflow for the expansion of heterologous biosynthetic pathways to natural product deriva-
tives. Nat Commun 12:1760
Hatzimanikatis V, Li C, Ionita JA, Henry CS et al (2005) Exploring the diversity of complex
metabolic networks. Bioinformatics 21:1603–1609
Henry CS, Broadbelt LJ, Hatzimanikatis V (2007) Thermodynamics-based metabolic flux analysis.
Biophys J 92:1792–1805
Henry CS, Broadbelt LJ, Hatzimanikatis V (2010a) Discovery and analysis of novel metabolic
pathways for the biosynthesis of industrial chemicals: 3-hydroxypropanoate. Biotechnol Bio-
eng 106:462–473
Henry CS, DeJongh M, Best AA, Frybarger PM et al (2010) High-throughput generation, opti-
mization and analysis of genome-scale metabolic models. Nat Biotechnol 28:977–982
Hou J, Scalcinati G, Oldiges M, Vemuri GN (2010) Metabolic impact of increased NADH avail-
ability in Saccharomyces cerevisiae. Appl Environ Microb 76:851–859
Rational Metabolic Pathway Prediction and Design … 23
Isaacs FJ, Carr PA, Wang HH, Lajoie MJ et al (2011) Precise manipulation of chromosomes in vivo
enables genome-wide codon replacement. Science 333:348
Jeffryes JG, Colastani RL, Elbadawi-Sidhu M, Kind T et al (2015) MINEs: open access databases
of computationally predicted enzyme promiscuity products for untargeted metabolomics. Jour-
nal of Cheminformatics 7:44
Jeske L, Placzek S, Schomburg I, Chang A, Schomburg D (2018) BRENDA in 2019: a European
ELIXIR core data resource. Nucleic Acids Res gky1048-gky1048
Kamp AV, Klamt S (2020) MEMO: A Method for computing metabolic modules for cell-free
production systems. ACS synthetic biology 9:556–566
Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M (2016) KEGG as a reference resource
for gene and protein annotation. Nucleic Acids Res 44:D457–D462
Khersonsky O, Tawfik DS (2010) Enzyme promiscuity: a mechanistic and evolutionary perspec-
tive. Annu Rev Biochem 79:471–505
Kim J, Reed JL, Maravelias CT (2011) Large-scale bi-level strain design approaches and mixed-
integer programming solution techniques. Plos One 6:e24162
Kim J-E, Jang I-S, Sung BH, Kim SC, Lee JY (2018) Rerouting of NADPH synthetic pathways
for increased protopanaxadiol production in Saccharomyces cerevisiae. Sci Rep 8:15820
King ZA, Feist AM (2014) Optimal cofactor swapping can increase the theoretical yield for chem-
ical production in Escherichia coli and Saccharomyces cerevisiae. Metab Eng 24:117–128
King ZA, Lu J, Dräger A, Miller P et al (2016) BiGG Models: a platform for integrating, standard-
izing and sharing genome-scale models. Nucleic Acids Res 44:D515–D522
Ko Y-S, Kim JW, Lee JA, Han T et al (2020) Tools and strategies of systems metabolic engineer-
ing for the development of microbial cell factories for chemical production. Chem Soc Rev
49:4615–4636
Koch M, Duigou T, Faulon J-L (2020) Reinforcement learning for bioretrosynthesis. ACS Synth
Biol 9:157–168
Kötter P, Ciriacy M (1993) Xylose fermentation by Saccharomyces cerevisiae. Appl Microbiol Biot
38:776–783
Kozak BU, van Rossum HM, Benjamin KR, Wu L et al (2014a) Replacement of the Saccharomyces
cerevisiae acetyl-CoA synthetases by alternative pathways for cytosolic acetyl-CoA synthesis.
Metab Eng 21:46–59
Kozak BU, van Rossum HM, Luttik MAH, Akeroyd M et al (2014) Engineering acetyl coenzyme
a supply: Functional expression of a bacterial pyruvate dehydrogenase complex in the cytosol
of Saccharomyces cerevisiae. mBio 5:e01696–01614
Kumar A, Suthers PF, Maranas CD (2012) MetRxn: a knowledgebase of metabolites and reactions
spanning metabolic models and databases. BMC Bioinformatics 13:6
Kumar A, Wang L, Ng CY, Maranas CD (2018) Pathway design using de novo steps through
uncharted biochemical spaces. Nat Commun 9:184
Kummel A, Panke S, Heinemann M (2006) Putative regulatory sites unraveled by network-
embedded thermodynamic analysis of metabolome data. Mol Syst Biol 2006(2):0034
Latendresse M, Krummenacker M, Karp PD (2014) Optimal metabolic route search based on atom
mappings. Bioinformatics 30:2043–2050
Lee SY, Kim HU (2015) Systems strategies for developing industrial microbial strains. Nat
Biotechnol 33:1061–1072
Lee WH, Park JB, Park K, Kim MD, Seo JH (2007) Enhanced production of epsilon-caprolactone
by overexpression of NADPH-regenerating glucose 6-phosphate dehydrogenase in recombi-
nant Escherichia coli harboring cyclohexanone monooxygenase gene. Appl Microbiol Biotech-
nol 76:329–338
Lee WH, Kim MD, Jin YS, Seo JH (2013) Engineering of NADPH regenerators in Escherichia
coli for enhanced biotransformation. Appl Microbiol Biotechnol 97:2761–2772
Lerman JA, Hyduke DR, Latif H, Portnoy VA et al (2012) In silico method for modelling
metabolism and gene product expression at genome scale. Nat Commun 3
Lewis NE, Nagarajan H, Palsson BO (2012) Constraining the metabolic genotype-phenotype rela-
tionship using a phylogeny of in silico methods. Nat Rev Microbiol 10:291–305
24 P. A. Saa
Lim SJ, Jung YM, Shin HD, Lee YH (2002) Amplification of the NADPH-related genes zwf and
gnd for the oddball biosynthesis of PHB in an E. coli transformant harboring a cloned phbCAB
operon. J Biosci Bioeng 93:543–549
Lin G-M, Warden-Rothman R, Voigt CA (2019) Retrosynthetic design of metabolic pathways to
chemicals not found in nature. Curr Opin Syst Biol 14:82–107
López J, Bustos D, Camilo C, Arenas N, Saa PA (2020) Engineering Saccharomyces cerevisiae for
the overproduction of β -ionone and its precursor β -carotene. Front Bioeng Biotechnol 8:1–13
López J, Cataldo VF, Peña M, Saa PA et al (2019) Build your bioprocess on a solid strain—β-
carotene production in recombinant Saccharomyces cerevisiae. Front Bioeng Biotechnol 7
Ma SM, Garcia DE, Redding-Johanson AM, Friedland GD et al (2011) Optimization of a het-
erologous mevalonate pathway through the use of variant HMG-CoA reductases. Metab Eng
13:588–597
Martínez I, Zhu J, Lin H, Bennett GN, San K-Y (2008) Replacing Escherichia coli NAD-dependent
glyceraldehyde 3-phosphate dehydrogenase (GAPDH) with a NADP-dependent enzyme from
Clostridium acetobutylicum facilitates NADPH dependent pathways. Metab Eng 10:352–359
Mazurenko S, Prokop Z, Damborsky J (2020) Machine learning in enzyme engineering. ACS Catal
10:1210–1223
Meadows AL, Hawkins KM, Tsegaye Y, Antipov E et al (2016) Rewriting yeast central carbon
metabolism for industrial isoprenoid production. Nature 537:694–697
Mikulak-Klucznik B, Goł˛ebiowska P, Bayly AA, Popik O et al (2020) Computational planning of
the synthesis of complex natural products. Nature 588:83–88
Moretti S, Tran, Van Du T, Mehl F, Ibberson M, Pagni M (2021) MetaNetX/MNXref: unified
namespace for metabolites and biochemical reactions in the context of metabolic models.
Nucleic Acids Res 49:D570–D574
Moriya Y, Shigemizu D, Hattori M, Tokimatsu T et al (2010) PathPred: an enzyme-catalyzed
metabolic pathway prediction server. Nucleic Acids Res 38:W138–W143
Moura M, Finkle J, Stainbrook S, Greene J et al (2016) Evaluating enzymatic synthesis of small
molecule drugs. Metab Eng 33:138–147
Ng CY, Wang L, Chowdhury A, Maranas CD (2019) Pareto optimality explanation of the glycolytic
alternatives in nature. Sci Rep 9:2633
Nielsen J, Keasling JD (2016) Engineering cellular metabolism. Cell 164:1185–1197
Nissen TL, Kielland-Brandt MC, Nielsen J, Villadsen J (2000) Optimization of ethanol production
in Saccharomyces cerevisiae by metabolic engineering of the ammonium assimilation. Metab
Eng 2:69–77
Nissen TL, Anderlund M, Nielsen J, Villadsen J, Kielland-Brandt MC (2001) Expression of a cyto-
plasmic transhydrogenase in Saccharomyces cerevisiae results in formation of 2-oxoglutarate
due to depletion of the NADPH pool. Yeast 18:19–32
Noor E, Bar-Even A, Flamholz A, Reznik E et al (2014) Pathway thermodynamics highlights
kinetic obstacles in central metabolism. Plos Comput Biol 10:e1003483
Noor E, Flamholz A, Bar-Even A, Davidi D et al (2016) The protein cost of metabolic fluxes:
prediction from enzymatic rate laws and cost minimization. Plos Comput Biol 12: e1005167
Pharkya P, Burgard AP, Maranas CD (2004) OptStrain: a computational framework for redesign of
microbial production systems. Genome Res 14:2367–2376
Pitkänen E, Jouhten P, Rousu J (2009) Inferring branching pathways in genome-scale metabolic
networks. Bmc Syst Biol 3:103
Saa PA, Nielsen LK (2017) Formulation, construction and analysis of kinetic models of
metabolism: a review of modelling frameworks. Biotechnol Adv 35:981–1003
Saa PA, Cortés MP, López J, Bustos D et al (2019) Expanding metabolic capabilities using novel
pathway designs: computational tools and case studies. Biotechnol J 14:1800734
San KY, Bennett GN, Berrios-Rivera SJ, Vadali RV et al (2002) Metabolic engineering through
cofactor manipulation and its effects on metabolic flux redistribution in Escherichia coli. Metab
Eng 4:182–192
Rational Metabolic Pathway Prediction and Design … 25
Sánchez BJ, Zhang C, Nilsson A, Lahtvee P-J et al (2017) Improving the phenotype predictions of
a yeast genome-scale metabolic model by incorporating enzymatic constraints. Mol Syst Biol
13:935
Savile CK, Janey JM, Mundorff EC, Moore JC et al (2010) Biocatalytic asymmetric synthesis of
chiral amines from ketones applied to sitagliptin manufacture. Science 329:305
Scalcinati G, Partow S, Siewers V, Schalk M et al (2012) Combined metabolic engineering of pre-
cursor and co-factor supply to increase α-santalene production by Saccharomyces cerevisiae.
Microb Cell Fact 11:117
Schilling C, Thakar R, Travnik E, Dien S, Wiback S (2005) SimPheny: a Computational Infras-
tructure for Systems Biology. In: US Department of Energy, Genomic Science Program publi-
cations
Schramm M, Racker E (1957) Formation of erythrose-4-phosphate and acetyl phosphate by a
phosphorolytic cleavage of fructose-6-phosphate. Nature 179:1349
Schwander T, von Schada Borzyskowski L, Burgener S, Cortina NS, Erb TJ (2016) A synthetic
pathway for the fixation of carbon dioxide in vitro. Science 354: 900–904
Smolke CD, Tyo KEJ (2012) Synthetic biology: Emerging methodologies to catalyze the metabolic
engineering design cycle. Metab Eng 14:187–188
Sonderegger M, Schumperli M, Sauer U (2004) Metabolic engineering of a phosphoketolase path-
way for pentose catabolism in Saccharomyces cerevisiae. Appl Environ Microbiol 70:2892–
2897
Tervo CJ, Reed JL (2016) MapMaker and PathTracer for tracking carbon in genome-scale
metabolic models. Biotechnol J 11:648–661
Theisen MK, Lafontaine Rivera JG, Liao JC (2016) Stability of ensemble models predicts produc-
tivity of enzymatic systems. Plos Comput Biol 12:e1004800
Tokic M, Hadadi N, Ataman M, Neves D et al (2018) Discovery and evaluation of biosynthetic
pathways for the production of five methyl ethyl ketone precursors. ACS Synth Biol 7:1858–
1873
van Rossum HM, Kozak BU, Pronk JT, van Maris AJA (2016) Engineering cytosolic acetyl-
coenzyme A supply in Saccharomyces cerevisiae: pathway stoichiometry, free-energy conser-
vation and redox-cofactor balancing. Metab Eng 36:99–115
Verho R, Londesborough J, Penttilä M, Richard P (2003) engineering redox cofactor regenera-
tion for improved pentose fermentation in Saccharomyces cerevisiae. Appl Environ Microb
69:5892–5897
Voigt CA (2020) Synthetic biology 2020–2030: six commercially-available products that are
changing our world. Nat Commun 11:6379
von Schada Borzyskowski L, Carrillo M, Leupold S, Glatter T et al (2018) An engineered Calvin-
Benson-Bassham cycle for carbon dioxide fixation in Methylobacterium extorquens AM1.
Metab Eng 47:423–433
Wang L, Dash S, Ng CY, Maranas CD (2017) A review of computational tools for design and
reconstruction of metabolic pathways. Synth Syst Biotechnol 2:243–252
Wortel MT, Noor E, Ferris M, Bruggeman FJ, Liebermeister W (2018) Metabolic enzyme cost
explains variable trade-offs between microbial growth rate and yield. Plos Comput Biol
14:e1006010
Yang X, Mao Z, Zhao X, Wang R et al (2020) Integrating thermodynamic and enzymatic con-
straints into genome-scale metabolic models. bioRxiv 2020.2011.2030.403519
Yim H, Haselbeck R, Niu W, Pujol-Baxley C et al (2011) Metabolic engineering of Escherichia
coli for direct production of 1,4-butanediol. Nat Chem Biol 7:445–452
Construction and Assembly
of Standardized Biobricks
for Synthetic Pathways Engineering
in Yeasts
Abstract
Over the past years, a tremendous variety of modular cloning and DNA assem-
bly strategies has been developed. By definition, they enable one-step fusion of
multiple DNA modules, arranged in a defined position and orientation. They
differ in terms of flexibility, robustness and precision, but they all overcome
limitations of time-consuming traditional cloning. Irrespectively of technical
details, modular cloning and standardized bioparts assembly strategies con-
tribute to great progress within the field of genetic engineering of yeast. Their
exploitation allows the researchers to focus on addressing the scientific ques-
tions, rather than on execution of tedious laboratory procedures. Modularity
and standardization facilitate wide-spread collaboration within the yeast com-
munity and speed-up accomplishing of set goals. This chapter provides an
overview of synthetic biology tools for assembly of large DNA constructs (in
many cases—from predesigned modules) to construct and fine-tune complex
DNA assemblies for yeasts. The principles of the techniques are presented, and
then followed by examples of yeast species-specific implementations for Sac-
charomyces cerevisiae, Komagataella phaffii and Yarrowia lipolytica. A general
division into restriction digestion/recombination-based techniques and overlap
extension assembly strategies were adopted.
Paulina Korpys-Woźniak and Monika Kubiak—These Authors contributed equally to the work and
should be considered as the two first Authors.
Gene cloning has gone a long way since its origin in ’70 of the XX century. For
many years, the DNA fragments to be cloned were processed with specific and
unique Restriction Enzymes (RE or ER—Endonuclease Restriction) and individu-
ally ligated with a vector that enabled maintenance of the fragment in the host cell,
and in some cases—its expression. The procedure had to be individually tailored
to a given sequence’s specificity, i.e., its nucleotide sequence and the presence of
ER Recognition Sites (RS) at a specific location. Along with time and complex-
ity of generated DNA constructions, the need for streamlined cloning procedures
has been growing. Traditional Restriction Digestion (RD) with specific RE, and
cloning of the fragments one-by-one did not meet growing demands. Until the con-
cept of modularity has been transplanted to molecular biology from engineering
and programming fields.
Modularity dissects a system (= complex DNA construction) into individual
modules (= biobricks/bioparts), each having a specific position and task to exe-
cute, in order to develop the desired functionality of the whole system. A given
module (e.g., promoter/ORF) may have different variants (e.g., strong, weak,
inducible/fluorescent, enzymatic) that can be interchangeably shuffled. Impor-
tantly, in a specific design of the system, each module has a strictly defined location
and function that cannot be exchanged with another module (only its variants can
be exchanged at the same position). A given module’s location (and function)
is predefined by a scaffold, which is a system of linking and docking adjacent
modules with each other. By organizing the modules within the scaffold, they
are correctly ordered and oriented, and cannot be translocated/shifted to another
location (only to another variant of the same module). Modularity gives the advan-
tage of an open upgrading option, provided that the new module/variant will fit
into the scaffold. This quality is not available for cloning through the traditional,
sequence-specific RE-based strategies.
To put the modularity concept into a molecular biology perspective, we need to
translate a term “module” into “a function at a specific location” within a multi-
part DNA construction (Fig. 1). This can be, for example, a promoter element at
the second T ranscription Unit (TU; TU is a system of promoter-ORF-terminator)
within a two TU-bearing DNA construction (P2 in Fig. 1).
This (P2) element is defined by specific linking/docking elements, flanking the
module. These may be specific ER RS, Recombinase (Rec) RS, OverLapping
protruding elements (OL; overlaps) generated by PCR or in vitro DNA synthe-
sis—anything that facilitates correct assembly (hybridization and ligation) of the
subsequent DNA fragments. They are specific for a given element (as P2 in our
example), and unique within the whole DNA construction. These elements guar-
antee correct positioning and orientation of the (P2) element within the multipart
DNA assembly. Depending on the linking/docking system adapted for a particular
technique, the final assembly may bear scars or be scarless.
Construction and Assembly of Standardized Biobricks …
Box 1
A scar is a short nucleotide sequence, present in the final assembly, that is
not a functional part of any of the subsequent bioparts, but was necessary at
the assembly stage for correct hybridization of the modules. Typical scars, in
traditional and in BioBrick clonings, are ER RS, in GoldenGate—4 nt over-
hangs, in the OL extension methods—depends on the technique design. The
ER-based techniques typically lead to scar-bearing assemblies. On the other
hand, OLs may be designed to leave a scarless assembly, when each OL is
composed solely of the 3’ and 5’ sequence of the two adjacent bioparts. But,
OL-based methods may leave the scars—when the OLs bear standardized
but unique short sequences, altogether generating a scaffold.
The (P2) module may have different variants, as stated above—e.g., strong,
weak, constitutive, inducible, etc., provided that they are fitted into the scaffold
by flanking with specific linking/docking elements. Such systematic exchange of
different variants of a given module, to test their functionality, is termed combina-
torial cloning. The process of transforming a given DNA fragment into a module
requires endowing it with the flanks, compatible with the system’s scaffold, and
ensuring that the nucleotide sequence of the fragment will remain inert upon the
enzymatic reaction taking place during the assembly reaction (restriction diges-
tion, recombination, ligation, amplification, etc., depending on the method). Thus,
the module must not contain any nucleotide sequence recognized and modified by
the enzymes used in the following enzymatic reaction, that physically assemble the
modules into a single DNA construction. The collection of such elements (DNA
fragments transformed into modules) is called a library; and, together with the
predefined scaffold, they can be also termed a standard. Sometimes, the term
“standard” refers only to the scaffold itself. In the majority of modular cloning
strategies, it is possible to generate extensive, multi-element DNA construction
in a single-tube and one-step reaction. It is facilitated by careful design of the
scaffold and the modules that are precisely positioned in the pre-designed order
and orientation; even if >10 elements are assembled (put into the same tube and
processed by the enzymes) at the same time.
Box 2
Once assigned to a given module functionality can be later modified. Consid-
ering our example, if a gene (ORF; G2) cloned under control of P2 must be
transcriptionally fused to an element (signal peptide or fusion partner) at its
5’ terminus, and this element will be shuffled in the following cloning (dif-
ferent variants will be tested), two different strategies could be followed: (i)
expanding the scaffold by modifying the linking/docking elements of P2 at 3’
terminus and G2 at 5’ terminus (adding additional fusion site) so that another
module can be inserted; (ii) functionality of all the elements upstream of the
Construction and Assembly of Standardized Biobricks … 31
Box 3
Modular cloning may have different formats in DNA construction assembly.
The number of shuffled modules in a single reaction may differ signifi-
cantly—from a single biopart (when all the remaining elements necessary
for the gene expression in yeast cell are already present in the destina-
tion/acceptor vector), up to all the “yeast elements” (when only the “bacterial
part” in a shuttle vector* remains unchanged). Technically, the approach
depends on the scaffold’s structure. An extensive and highly specific scaf-
fold (with a large number of unique and specific linking/docking elements)
accepts more elements to be shuffled in a single reaction. Limited scaffold,
allow shuffling one or two bioparts at a time. In the latter case, all the remain-
ing elements, necessary for full functionality of the expression cassette must
be already present in the acceptor vector. (* typically the DNA bioparts
assembly is conducted in a shuttle vector—bearing a “bacterial part”, to
be assembled and amplified in E. coli, and a “yeast part”—composed of
yeast-specific regulatory elements and the target genes, to be transformed
and expressed in a yeast host cell).
32 P. Korpys-Woźniak et al.
Over the past years, a tremendous variety of modular cloning and DNA assem-
bly strategies has been developed. They differ in terms of flexibility, robustness
and precision of cloning, but they all overcome limitations of time-consuming
traditional cloning using RD and DNA ligation, typically joining only two, “single-
use” DNA parts simultaneously. DNA assembly (incl. modular cloning) typically
enables one-step fusion of multiple DNA modules, arranged in a defined position
and orientation. This chapter provides an overview of “state-of-the-art” synthetic
biology tools for assembly of large DNA constructs (in many cases—from pre-
designed modules) to construct and fine-tune complex DNA assemblies for yeasts.
For the sake of clarity and comprehensiveness, the principles of the techniques are
presented and then followed by yeast species-specific examples. Saccharomyces
cerevisiae, Komagataella phaffii and Yarrowia lipolytica were chosen as the case
studies, due to their fundamental role or high interest in the yeast community.
Selection of methods was an arbitrary choice of the authors, based on their pop-
ularity in research by the yeast community. General division into RE-dependent
and OL assembly methods was adopted.
2 Historical Outline
The evolution of modular cloning and DNA bioparts assembly strategies spans
the last 30 years. General trend of the techniques’ evolution leads from founding
the basic principles of a method, through demonstration of its applicability in a
given yeast species, up to development of the so-called, toolkit (or toolbox), built
of species-specific standardized modules and/or DNA constructions and recipi-
ent strains. In order to put the techniques into a research community perspective
and illustrate dynamics of the methods’ evolution, in this section we summarize
timeline of development of the most popular techniques, according to that key—
establishing the technical concept, through “the first mention” for a given yeast
species, up to a toolkit development (where applicable; Fig. 2a, b). A more detailed
description of the techniques with deeper insight into current state of the art is
given in the following sections of this chapter.
Technical implementation of the DNA bioparts assembling concept was at first
addressed by Overlap Extension PCR (OE-PCR). The method was initially adopted
for assembly of relatively short elements (300–400 bp), incl. gene’s exons, or
parts located upstream and downstream of the mutation site (two parts) (Ho et al.
1989; Horton et al. 1989). The technique was later adopted for creating chimeric
genes or gene alleles in S. cerevisiae (Anthony and Liebman 1995), K. phaffii
(Eldin et al. 1997) and Y. lipolytica (Zhang et al. 2010). In many cases, due to
simplicity of the method, lack of spectacular technical background behind it and
low efficiency when compared to the other assembly method, the authors tend
to use it as a routine laboratory technique for simple assemblies, without special
mention. Nevertheless, later elaborated variants of OE-PCR, combined with “in
yeasto” assembly, demonstrated the great potential of that simple method, which
ultimately enabled assembly of extensive DNA constructions (Shao et al. 2009;
Gao et al. 2014).
Construction and Assembly of Standardized Biobricks … 33
Fig. 2 Historical outline of DNA bioparts assembly and modular cloning strategies invention and
implementation. a Inventions and implementations are given chronologically. b Inventions and
implementations are sorted by the principal technique
Box 4
The term “in yeasto” refers to phenomena or reactions taking place within
the living yeast cell. It may be considered as an abbreviation of stating
“in vivo in the yeast cell”. It can be sometimes found in the research papers
issued by/dedicated to the yeast community.
3.1 BioBricks
Box 5
Isocaudomers are pairs of RE that recognize slightly different RS but upon
cleavage, they generate compatible overhangs. After ligation of such pro-
truding elements, an asymmetrical sequence is generated and thus cannot be
recognized by any of the ERs—it is lost, but in this way, the sequence is sta-
ble during the next cycle of assembly. Examples of IsoCMs: 1: BamHI/BglII,
2: NdeI/MseI, 3: XbaI/AvrII/NheI/SpeI, 4: XhoI/SalI.
Fig. 3 General strategy for BioBrick parts assembly (e.g.: promoter, ORF and terminator). Each of
the modules is first equipped with prefix and suffix bearing RS for a single URE (unique restriction
endonuclease recognition site) and a single IsoCM (isocaudomer restriction endonuclease recog-
nition site). Full description is given in the text. “Scar” site is shown as a purple circle in the final
assembly. Circular objects indicate the bacterial ori of replication (black) and ampicillin resistance
gene AmpR (orange) contained in the “bacterial” part of the assembly, which is typically discarded
before yeast cell transformation
modules are prepared by adding specific prefixes and suffixes in the PCR to a
biopart, each bearing RS for URE at 5’ and IsoCM at 3’. The two modules to
be assembled are then cleaved in a specific pattern each with URE and IsoCM to
generate compatible overhangs at the linking/docking site. The overhangs at the
fusion site are generated by different IsoCM for each of the modules. RS of URE
flank the assembly, and compatible RS are present in the destination vector. As
a result of the following ligation reaction, the sequence between the fused DNA
fragments bears a 6 bp scar, but the URE-bearing prefix and suffix sequences
remain unchanged, hence the other BB parts can be assembled in the next step. In
this technique, the destination vector bearing two assembled bioparts becomes an
entry vector in the next cycle of assembly.
The reaction conditions are standard for ER and ligation (typically 1 h at 37 °C
and 1 h 16–22 °C). Most of compatible BBs for yeasts are listed on the website
Construction and Assembly of Standardized Biobricks … 37
executed, and positional effects and different genes’ configurations could be stud-
ied. The system of reusable modules compatible with the BB standard is available
from Addgene.
3.2 GoldenGate
Box 6
Type IIS RE belongs to a specific group of RE that recognize asymmetric
DNA sequences (BsaI: GGTCTC) and cleave at a defined distance outside
of their RS, (1–20 bp) but the combination of nucleotides at the cleavage
site is not defined. Since the sequence at the cutting site is indifferent to
the RE, it can be modified/designed. It is possible to generate a system of
unique sequences, that will form overlaps after the cleavage with type IIS
RE, which is the essence of designing a scaffold for GG cloning.
In this way, the user may design overhangs that will be generated upon cleavage.
A set of such unique overhangs, serving as a linking/docking system, consti-
tutes a scaffold of the GG standard defining position and orientation of assembled
parts (Fig. 4). To assure compatibility of the modules, the scaffold must remain
unchanged. Adjacent bioparts share the same overhang at the junction points. Mod-
ules are prepared by PCR amplification with primers adding ER IIS RS at 5’,
followed by the 4 bp overhang in the middle of the primer, and a sequence com-
patible with the biopart at its 3’ end. Such construction allows to “loose” the type
IIS RE RS after cleavage, leaving only the 4 bp overhang after digestion and a scar
after fusion. To generate a library of GG modules, the PCR-amplified bioparts are
cloned in the entry vectors. The entry vectors with the modules are pooled together
with a destination vector in the assembly reaction. The destination vector typically
bears some chromophore gene that serves as a negative control after assembly and
cloning in E. coli. The chromophore region in the destination vector is flanked with
type IIS RE RS cloned inwardly, so the cleavage leaves 4 bp overhangs at the vec-
tor’s termini, and the chromophore cassette with the type IIS RE RS is discarded.
The reaction is conducted in a cycling profile of subsequent incubations at 37 °C
for ER action and 16 °C for the action of T4 ligase. With each ER/ligation cycle,
new fusions are generated, devoid of type IIS RE RS—disallowing any further
modifications. The number of elements that can be fused in a single cycling reac-
tion in a single tube is limited by the scaffold design. More than ten modules are
Construction and Assembly of Standardized Biobricks … 39
Fig. 4 General outline of GoldenGate parts assembly (e.g., promoter, ORF and terminator). Each
of the modules is first equipped with RS for type IIS ER and 4 bp overhang sequences (indicated
as red, purple, orange and green XXXX), according to a predesigned scaffold. Full description is
given in the text. 4-bp scar sites are shown as purple and orange rectangles in the final assembly.
Circular objects indicate the bacterial ori of replication (black) and bacterial resistance gene: KanR
(blue) and AmpR (orange) contained in the “bacterial” part of the assembly, which is typically
discarded before yeast cell transformation
regularly fused (examples are given hereafter) with a cloning efficiency of about
90% (Engler et al. 2008).
The GG concept-based variation, called MoClo (Weber et al. 2011), allows gen-
erating large multigene construction (up to 33 kb) in three cloning steps, from basic
modules, through pre-assembled TUs, finally merged into an expression construct.
40 P. Korpys-Woźniak et al.
Box 7
VEGAS is a molecular biology technique exploiting the innate capacity of a
microbial cell to combine DNA fragments with terminal complementarity by
homologous recombination using orthogonal adapter sequences (VAs). The
adapters are introduced to the DNA fragments as terminal modules flanking
the TUs in the core. Their location and orientation direct desired order of
the core pathway genes.
First, five individual DNA assemblies (violacein synthesis pathway), each com-
posed of five modules in a 6-overhang scaffold, were constructed in vitro in a
typical GG reaction. In each assembly, central TU (promoter, gene, terminator)
was flanked with terminal adapters, facilitating in vivo homologous recombina-
tion. The five DNA assemblies were transformed into E. coli and correctly fused
Construction and Assembly of Standardized Biobricks … 41
the module variants allowed to rapid construction and test of 18 DNA assemblies.
A very similar approach was undertaken in a very recent study by (Püllmann et al.
2020) aiming at overexpression of eight enzyme-encoding genes. In that study, the
authors prepared destination vectors with a pre-cloned dominant selection marker,
promoter and terminator for the gene of interest transcription, and ARS for episo-
mal maintenance. Thus, only a 4-fusion sites GG scaffold was used for the modules
shuffling. The three shuffled modules were: (i) signal peptide (one of 17 variants),
(ii) target gene (out of 8), (iii) protein tag for purification (7 variants available).
The authors deposited all the modules in Addgene for the yeast community.
A remarkable variety of GG modular cloning implementations were reported for
Y. lipolytica, with the first mention published in 2017 (Celińska et al. 2017). In that
report, a scaffold composed of 13 fusion sites was developed for modular cloning
of 12 bioparts, equipped with specific 4 bp overhangs. The authors proposed 48
new modules, to be fitted into the scaffold, covering different promoters, termina-
tors, selection markers and elements directing targeted genomic integration. Soon
after that first mention, a report on biotechnologically relevant exploitation of that
scaffold and the modules was published (Larroude et al. 2018). The authors shuf-
fled promoters of different strengths to optimize expression of a 3-gene carotenoids
synthesis pathway in Y. lipolytica. The shuffling process was conducted in a single-
tube, single-step reaction to assembly 12-module pathway of >12 kb. The initial
scaffold was later on modified by the addition of a new overhang into the scaffold
(Celińska et al. 2018). The additional fusion site was used to enable combinatorial
cloning of different signal peptides, to be tested with different heterologous secre-
tory proteins. The initial scaffold was designed to enable the rapid fusion of 12
bioparts—3 TUs + selection marker +2 integration targeting sites. Introduction of
this additional fusion site enabled the reuse of the previously developed modules
(the problem is addressed in Box 1) and insertion of an additional one for a sig-
nal peptide (between promoter and target gene). The site was carefully designed
to maintain the reading frame unchanged and to minimize amino acid changes
within the encoded secretory proteins. The authors shuffled ten novel signal pep-
tides with two reporter proteins to indicate optimal fusions. In the following study
(Celińska et al. 2020), the best fusions (signal peptide + gene) were amplified as
single modules and could be cloned in the former scaffold, as multiple TUs bear-
ing DNA assembly. In that study, the authors investigated positional effect on the
genes’ expression, by combinatorial cloning of optimized fusions (signal peptide
and the genes) in either the first or the second TU. Thanks to the modularity of
the GG method, it was possible to rapidly generate the necessary set of DNA con-
structions. An improved scaffold with an enriched library of modules (altogether
64 modules, validated with three different fluorescent reporters) was finally made
available to the yeast community through Addgene (Larroude et al. 2019). The
same scaffold, either in its full length or limited to a single/double TUs was later
exploited in different applications, like rapid single genes cloning under strong
promoter (Korpys-Woźniak et al. 2020) or testing the strength of newly described
promoters (Park et al. 2019).
Construction and Assembly of Standardized Biobricks … 43
3.3 Gateway
Box 8
There are four different variants of att sites (B, P, L, R). All of them con-
tain a core 25-bp “recognition region” (black box) whereas L, R/P contain
“arms” on either/both sides of the core, acting as interaction sites for the
recombination enzymes (left arm—white box; right arm—lined box). The
core region bears a central 7-bp “asymmetric overlap” (yellow box) that is
recognized by clonases and thus determines where DNA is cut and rejoined.
44 P. Korpys-Woźniak et al.
Box 9
Definition of att sites’ subtype is executed by making nucleotide changes
within the 25-bp region. Setting different subtypes of att sites facilitates
cloning of each DNA fragment in only one position and orientation.
In the final cloning stage, the LR clonase executes the reverse reaction to BP
clonase, generating attB sites at the fusion sites. The destination vector bears attR
sequences flanking ccdB toxic cassette, which, when maintained, disable clones
growth. The reaction is conducted in vitro, over short incubation (1 h) at 25 °C.
The multipart format allows for one-step cloning of up to five modules. As the
other RD-/Rec-based assembly techniques, GW leaves (relatively long—25-bp)
scars at the fusion sites. One of the biggest advantages of GW techniques is a wide
array of destination vectors available, which can be either purchased or obtained
from Addgene/iGEM collections.
Apart from the GW assembly system, some other recombination-based cloning
systems have been issued, like Univector (Liu et al. 1998) and Creator (Siegel
et al. 2004) relying on the use of Cre recombinase and loxP sites, or In-Fusion
Construction and Assembly of Standardized Biobricks … 45
Fig. 5 General outline of Gateway cloning strategy (e.g., promoter, ORF and terminator). Each
of the modules is first equipped with RecRS by the introduction of attB sites (indicated as red,
purple, orange and green arrows), according to a predesigned scaffold. The reaction is two-stage as
described in detail in the text. 25-bp scar sites are shown as red, green, purple and orange rectangles
in the final assembly. Circular objects indicate the bacterial ori of replication (black) and bacterial
resistance gene: KanR (blue) and AmpR (orange) contained in the “bacterial” part of the assembly,
which is typically discarded before yeast cell transformation
46 P. Korpys-Woźniak et al.
(Berrow et al. 2007), MAGIC (Li and Elledge 2005) and SEFC (Zhu et al. 2010)
systems using homologous recombination either in vitro or within E. coli cells.
Soon after setting the principles of the GW method, first reports on its imple-
mentation into yeasts systems were published. Funk proposed and validated a
system of 32 acceptor vectors for att-based cloning in S. cerevisiae with four
different promoters (two constitutive and two inducible), four different epitopes
for protein tagging and four auxotrophy selection markers (Funk et al. 2002).
The destination vectors are compatible with donor vectors bearing elements to
be cloned flanked with attB sites. Operability of the system was demonstrated
using a fluorescent reporter protein. Similarly, at approximately the same time, a
set of 20 destination vectors for recombinant protein production in S. cerevisiae
was designed and constructed in accordance with GW cloning (Van Mullem et al.
2003). In that toolkit, the vectors differ in the selection marker modules (2 variants
available) and the epitope modules for protein tagging (four variants available).
Operability of the system was validated with two entry vectors bearing differ-
ent genes. In 2007, an extensive library of 285 S. cerevisiae- and GW-compatible
destination vectors was developed, and made available to the yeast community
by Addgene (Alberti et al. 2007). The vectors were constructed based on com-
monly used pRS shuttle vector series. Variants covered by that collection bear
either constitutive or inducible promoters and one of several options for epitope
tagging (with a short tag or a fluorescent fusion partner). Specific implementations
of the GW toolkits in S. cerevisiae metabolic engineering cover, e.g., enhanced
synthesis of phenylalanine for improved production of styrene (McKenna et al.
2014), or engineering acetyl-CoA synthesis by introducing alternative acetyl-CoA
synthetases (out of 4 modules prepared) (Kozak et al. 2014).
All the mentioned above GW vectors were intended for a single biopart (gene)
modular cloning, but the multisite GW system was also used in S. cerevisiae.
The first GW system for multipart assembly was designed for 3-modules fusion
(Nagels Durand et al. 2012). The authors presented a set of 3 destination vectors
for gene expression in S. cerevisiae, enabling fast assembly of any promoter, open
reading frame, and epitope tag in combination with any of 4 auxotrophic markers
and 3 distinct replication mechanisms. The system was proved operable based on
an example of a 3-domain protein complex. A set of GW shuttle vectors for 2-
biopart assemblies was presented by (Giuraniuc et al. 2013). The authors used the
system to develop 2-parts fusions to make a new, synthetic regulatory element, and
to generate combinatorial fusions between five different promoters and 12 ORFs.
The first implementation of GW standard to K. phaffii system was reported by
(Esposito et al. 2005). The authors generated a GW destination vector by mod-
ification of a popular pPICZ vector for secretory overproduction of proteins in
a methanol-inducible manner. Interestingly, the authors used modified attB sites,
which were reported to give 4-times more clones after transformation, than the
original attBs. Of key importance was the fact, the authors evidenced a lack
of negative impact of the scar (att site) left between N-terminal fusion and the
mature polypeptide-encoding sequence. With the same aim and highly correspond-
ing approach, two GW-/K. phaffii-compatible destination vectors were developed
Construction and Assembly of Standardized Biobricks … 47
4.1 Gibson
Fig. 6 General outline of Gibson cloning (e.g., promoter, ORF and terminator). Each of the mod-
ules is first amplified using binary primers covering sequences of adjacent elements (indicated as
red, purple, orange and green OL XXX). The OLs are typically 40 nucleotides in length and define
the final arrangement of the construct. The reaction is two-stage as described in detail in the text.
No scars are left at the fusion sites. Circular objects indicate the bacterial ori of replication (black)
and ampicillin resistance gene AmpR (orange) contained in the “bacterial” part of the assembly,
which is typically discarded before yeast cell transformation
Construction and Assembly of Standardized Biobricks … 49
OLs must be also present in the destination vector (OL compatible with 5’ of the
first module and 3’ of the last module within the assembly). Overlapping bioparts
are mixed in a single tube with the destination vector and subjected to the action of
three different enzymes. The 5’-exonuclease generates single-stranded DNA OLs
by acting from the 5’-end so that the complementary regions are exposed and can
be annealed. High-fidelity DNA polymerase fills the gaps and Taq DNA ligase
covalently joins fragments. The reaction is a single step and isothermal (50 °C,
over 15–60 min). Typically, with the use of a commercial mix of enzymes, up
to 5 modules can be assembled simultaneously in a single-tube reaction using a
one-step master mix of enzymes, whereas a two-step reaction allows combining as
many as 15 bioparts. The size limit for this in vitro DNA assembly is unknown.
Typically, efficient cloning of the cassettes reaching 20-kbp is reported, but the
largest demonstrated construct has reached 900-kbp (Gibson 2009; Gibson et al.
2009). The GB method was a basis for the development of other techniques, like
MODAL (Casini et al. 2013) and UNS (Unique Nucleotide Sequences)-guided
isothermal assembly (Torella et al. 2014).
The GB method’s principles were first described in a paper on the enzymatic
assembly of DNA bioparts in vitro (Gibson et al. 2009), where an isothermal,
single-step reaction for assembling multiple overlapping DNA molecules by the
concerted action of the three enzymes was described. As in the other OL-based
techniques, no sensu stricte scaffold is needed (as it is for the GG method).
Recently, (Cataldo et al. 2020) constructed an expression cassette for S. cerevisiae
via the so-called “full in vitro Gibson assembly.” The authors observed a toxic
effect of the cloned genes on E. coli (the subcloning host). The in vitro-assembled
GB cassette was composed of 8 bioparts (2 genes, bidirectional promoter, 2 ter-
minators, up and down elements for integration, and a selection marker) bearing
40 bp OLs. To bypass the bottleneck of subcloning in E. coli, the full-length assem-
bly was used as a template in a PCR. The amplified cassette was then transformed
into the final host—S. cerevisiae.
Several interesting implementations of the GB method to complex DNA assem-
blies construction for K. phaffii were described. Combination of OE-PCR and
GB assembly methods was adopted to construct an extensive, 12-bioparts bear-
ing DNA assembly in 12 different variants (Vogl et al. 2016). A 4-gene pathway
for carotenoids synthesis was used as a model. At first, the authors generated a
library of 12 methanol-dependent promoters of different strengths and a differ-
ent mechanism of induction; and an additional collection of 4 terminators. They
assembled a complete expression cassette in a step-wise manner. First, 3 indi-
vidual elements building individual TUs (promoter, gene, terminator) were fused
using OE-PCR. Subsequently, 4 such TUs were assembled via the GB method to
generate a complete DNA construction. A combinatorial approach was followed to
construct different variants of the complete DNA assembly, each bearing 4 TUs.
This way, in the initial OE-PCR, 3 bioparts were assembled, and in the follow-
ing GB reaction, 4 modules (each bearing individual TU) were shuffled to build a
complete carotenoid pathway, but controlled by different regulatory elements. The
strategy enabled fine-tuning of the model pathway.
50 P. Korpys-Woźniak et al.
4.2 USER®
Fig. 7 General outline of USER® cloning (e.g., promoter, ORF and terminator). Each of the mod-
ules is first amplified using binary primers covering sequences of adjacent elements (indicated as
red, purple, orange and green XXX). The primers include a single deoxyuracil residue (dU) flank-
ing the 3’ ends of the homology region. The OLs are typically 7–12 nucleotides in length and define
the final arrangement of the construct. The reaction details are given in the text. No scars are left
at the fusion sites. Circular objects indicate the bacterial ori of replication (black) and ampicillin
resistance gene AmpR (orange) contained in the “bacterial” part of the assembly, which is typically
discarded before yeast cells transformation
report, marker recycling was facilitated by flanking the gene by direct repeats.
Soon after that report, the same Research Group evolved a previously developed
USER-based system and developed EasyClone system for iterative integration of
DNA constructions assisted with a selection marker recycling system based on a
loxP-Cre mechanism (Jensen et al. 2014). The USER-constructed system covered
both episomal and integrative destination vectors. Their construction was iden-
tical as previously (Mikkelsen et al. 2012) with the difference in the selection
marker gene flanks—in the updated version—loxP elements. USER-reaction was
each time conducted with a constitutive promoter and a fluorescent reporter protein
(out of 3 available). The authors sequentially transformed the host strain, recycled
the selection marker, reaching up to 3 reporters coexpression.
USER cloning was combined with the CRISPR-Cas9 genome-editing system
to create the EasyClone-MarkerFree vector toolkit for S. cerevisiae (Jessop-Fabre
et al., 2016). The new system was developed on a backbone of EasyClone 2.0
vectors, by amplifying a fragment with bacterial elements, specific homology-
integration sites, and the USER-cloning site flanked with terminators. Most
importantly – the yeast selection markers were not included in the new backbone
vector. The expression cassettes (two genes flanking a bidirectional promoter) to be
cloned in such a new backbone vector were first assembled (via USER) and then
fused with the marker-less vector. To precisely direct genomic integration of these
cassettes, a gRNA-expressing vector was prepared. In the proof-of-concept exam-
ple, the authors prepared three USER-cloned assemblies, each bearing two genes
under bidirectional promoter and specific homology-integration regions, and a vec-
tor bearing 3 gRNAs, compatible with the homology regions. The three expression
cassettes were then cotransformed with the triple gRNA vector into a S. cerevisiae
strain already synthetizing Cas9 from an episomal vector. The cassettes are then
integrated into loci specified by gRNA and the 500-bp homology regions. The
vector expressing gRNAs is then lost by subculturing the strains in a non-selective
medium. Similarly, Cas9-bearing helper vector could be lost, if no further engi-
neering was planned. The developed system enables simultaneous introduction of
up to three integration cassettes into the genome of S. cerevisiae without the use
of any yeast selection markers. Using standardized USER-bioparts, the integration
cassettes can be constructed for overexpression of one or two genes per integra-
tion site. The authors suggested that for better stability, the selected integration
sites should not be in direct proximity to each other. The system was validated not
only in laboratory strain (CEN.PK) but also in an industrial diploid (EthanolRed),
which is of great interest. The EasyClone-MarkerFree vector toolkit is available to
the yeast community from Addgene.
The USER® assembly system was used in multiple following studies by that
Research Team. One of the interesting examples is a paper presenting a method
enabling simultaneous disruption of multiple genetic targets, where a combina-
tion of CRISPR/Cas9, in vivo recombination, USER assembly, and RNAi was
developed (Kildegaard et al. 2019). USER cloning was used there for assembly
of multipart expression cassettes. Interestingly, the authors improved the modular-
ity of USER by introducing standardized spacers/overhangs of 60-bp at the fusion
Construction and Assembly of Standardized Biobricks … 53
sites, flanking individual TUs. These were used in combination with standard over-
hangs adopted for USER cloning, for fusions inside the TUs. Such specific design
allowed reusing a standard set of primers for amplification of different TUs, so
the genes and regulatory elements could be cloned in a combinatorial manner. By
adopting that sophisticated system of cloning, the authors constructed an exten-
sive assembly, covering 3 TUs, with a selection marker element, and 2 genomic
integration regions flanking the whole assembly.
The USER-based EasyClone system was also adopted for Y. lipolytica in a
form of a comprehensive toolbox (Holkenbrink et al. 2018). The system com-
prises a broad set of integrative expression vectors, compatible with USER cloning,
which allows cloning up to 2 genes. As in the case of the EasyClone toolkit for
S. cerevisiae, the “Yali” system was extensively studied in terms of the selection
of intergenic and highly transcribed integration loci. As for the former system,
stringent requirements were set to propose candidate integration sequences, and
the presumptions were carefully verified using a reporter protein system. Within
the EasyClone Yali toolbox, the Authors proposed 5 integrative, marker-less vec-
tors, precisely directed for integration by CRISPR/Cas9 system, and 11 vectors
to be integrated and maintained using auxotrophic/resistance selection mark-
ers. Additionally, the vectors and protocols for marker-less deletion of up to 2
genes simultaneously using the CRISPR/Cas9 system were provided. Using 90-bp
double-stranded oligos as the DNA repair templates, the authors obtained efficien-
cies of up to 90% and 66% for individual and double knockouts, respectively.
The expression/integration cassette construction in EasyClone Yali is facilitated
by USER cloning and a rich collection of standardized and reusable bioparts. The
EasyCloneYali vectors can be obtained via Addgene. Operability and robustness of
the system were evidenced in another paper by that research Team on cloning and
optimization of astaxanthin synthesis pathway in Y. lipolytica (Kildegaard et al.
2017). In that study, a great variety of diverse combinations of 6 genes of differ-
ent origins was shuffled and cloned in both laboratory auxotrophs and native Y.
lipolytica strains. Thanks to intensive engineering of several side-branches of the
terpenes synthesis pathway, the flux through the target astaxanthin was enhanced.
Among the other, efficient cloning techniques were the enabling factor for that
study.
4.3 OE-PCR
Fig. 8 General outline of OE_PCR assembly strategy. Each of the modules is first amplified using
binary primers (represented as letters A-H) covering sequences of adjacent elements: promoter
(solid lines), open reading frame (ORF, dotted lines), terminator (dashes and dots) and destina-
tion vector. Complementary regions to the adjacent parts are marked in colors (red, purple, orange
and green squares). Product AB is generated using “A” and “B” primers on the promoter template.
Parts “CD” “EF” and “GH” are amplified in the same way. Restriction enzyme sites are intro-
duced in primers (black boxes). The reaction details are given in the text. No scars are left at the
fusion sites. Circular objects indicate the bacterial ori of replication (black) and ampicillin resis-
tance gene AmpR (orange) contained in the “bacterial” part of the assembly, which is typically
discarded before yeast cells transformation
Construction and Assembly of Standardized Biobricks … 55
between adjacent fragments, the fusion product may not be correctly assembled
(Cha-aim et al. 2009). However, several specific improvements and modifications
were incorporated to the techniques, to address these issues (Xiao and Pei 2011;
Guo et al. 2019). For example in 2007, (Dong et al. 2007) improved the OE-PCR
technique by introducing into the protocol several elements, like primers design
by DNA Works software instead of manual design, provided the Melting Temper-
atures™ of the OLs to ensure their specificity, reduced the length of OLs down
to < 5 bp, and promoted the use of high fidelity DNA polymerase, which was
specifically designed for OE-PCR conditions. The technique in the new format
was validated with several genes of ~1–1.5 kbp.
While in its basic format OE-PCR technique may seem inferior to the other
modular cloning and DNA assembly techniques, several interesting modifications
advanced the technique and scope of its exploitation. To more easily apply OE-
PCR to the generation of DNA assemblies and improve its specificity, a system
of short, GC-rich OLs at the fusion sites was proposed (Cha-aim et al. 2009).
The authors demonstrated that introduction of short G/C stretches can significantly
improve specific annealing of compatible DNA fragments in a fusion PCR. Pre-
cisely, C(15) OL (but not GC repeat) turned out to be highly efficient in directing
specific hybridization between the fusion elements under all examined tempera-
tures. Such stretch cannot be used for fusion of functionally related parts (signal
peptide and mature polypeptide or promoter and ORF) but could be used for fusion
of independent bioparts—integration sites, selection markers, TUs. The authors
demonstrated operability of their method on 3 bioparts fusion in a single reaction.
In 2009 (Shao et al. 2009) reported on “DNA Assembler” strategy for “in
yeasto” assembly of complex, multi-element DNA constructions. In fact, the
method highly corresponds to the initial, model experiment described by (Gibson
2009), when a 1170-bp DNA fragment was dissected into 38 short oligomers (60-
bp long) and flanked with 20-bp complementarity regions to adjacent fragments,
that were assembled “in yeasto.” The key difference between the two reports is
the size of the assembled bioparts. Gibson assembled ~1 kb fragment from 38
synthetic oligos, and (Shao et al. 2009) executed complex pathway assembly of
>10-kpb. Indeed, both techniques should be classified in the same sort of cloning
strategies, as they rely on identical principles. Basically, the method relies on
designing short OLs flanking adjacent fragments to be assembled. These can be
either short oligos (Gibson 2009), or modules bearing regulatory elements and
heterologous genes, as in the case of (Shao et al. 2009). For the latter, all proof-
of-concept cassettes were assembled step-wise. The cassettes were as follows: (i)
xylose utilization pathway (∼9 kb DNA consisting of 3 genes), (ii) zeaxanthin
biosynthesis pathway (∼11 kb DNA consisting of 5 genes) and (iii) combined
xylose utilization and zeaxanthin biosynthesis pathway (∼19 kb consisting of 8
genes). Basically, all TUs were assembled in vitro by standard OE-PCR. Adja-
cent TUs were already flanked with specific OLs, enabling correct assembly in
desired orientation upon transformation into the yeast cell. The length of OLs
was individually adjusted to the complexity of the DNA construction—less com-
plex assembly (3–5 genes, ~10 kb) could be correctly fused with only 50-bp OLs
56 P. Korpys-Woźniak et al.
reaching very high efficiency of 80–100%, while for more complex DNA construc-
tions (8 genes, ~19 kb) the efficiency of 40–70% could be achieved with longer
OLs (~125–430 bp). It was also demonstrated that by increasing the amount (in
ng) of the modules per the same amount of the backbone vector, the efficiency
of assembly can be increased. The authors stressed simplicity and rapid character
of the DNA assembly method, which actually requires only appropriate design of
PCR and exploitation of native recombination system of the yeast cell.
Many other, aim-specific implementations of OE-PCR to S. cerevisiae genetic
engineering have been described. Just to give several different examples, OE-PCR
was exploited for high-throughput production of linear DNA templates, that were
later processed by cell-free transcription-translation system from S. cerevisiae (Gan
and Jewett 2014), or for generation of chimeric genes involved in phenylpropanoid
pathway (Jiang et al. 2005) or assembly of resveratrol synthesis pathway (Wang
et al. 2011).
Likewise, multiple literature reports demonstrate wide applicability of OE-PCR
for the assembly of synthetic DNA constructions that are later used for K. phaffii
modifications. OE-PCR was used to adjust standard pGAPZα vector for removing
unnecessary elements, and high-throughput mutagenesis of Candida rugosa lipase
to be overproduced in K. phaffii (Chang et al. 2005). The scientific aim of that
study was to exchange a rare codon for a frequent codon, which improved enzyme
production. OE-PCR was also used in K. phaffii-related applications to conduct
in vitro splicing (Jeya et al. 2009), or generate chimeric DNA templates encoding
a gene with a fusion partner and cell-surface anchor (Yang et al. 2017).
The “DNA Assembler” technique reported by (Shao et al. 2009) was later
“transplanted” to Y. lipolytica system to rapidly assemble multipart DNA cassettes
(Gao et al. 2014). While the authors termed their cloning strategy as one-step
assembly, in fact the approach was executed in a step-wise manner. First, the TUs
and the flanking regions homologous to rDNA sequences in Y. lipolytica genome
were assembled by OE-PCR in groups of 3–4 bioparts. Terminal elements of indi-
vidual OE-PCR-assembled TUs were overlapping adjacent bioparts according to
a design. The OLs between the adjacent parts were 45–60-bp, depending on the
constructed variant. Such pre-assembled TUs were then mixed and transformed
into the recipient strains which assembled them. The efficiency of the assembly
reaction of a 4-gene pathway ranged between 11 and 21%, but the most impor-
tant advancement relates to a decrease in the time necessary for the procedure
completion. According to the authors calculations, the total procedure could be
completed in less than one week, as compared to a previously reported sequential
gene integration method that required n weeks for n genes (Celińska and Grajek
2013).
In 2020, (Guo et al. 2020) used OE-PCR to assemble 9 elements, altogether
forming the first completely synthetic yeast artificial chromosome for Y. lipolytica
(Yl-YAC), with centrally located ARS and flanked with telomeres. As the authors
mainly aimed at optimization of Yl-YAC construction, they adopted combinatorial
approach to testing different combinations of chromosomal elements (ARS and
Construction and Assembly of Standardized Biobricks … 57
TEL) and reporter gene-encoding modules, which were transformed into Y. lipoly-
tica cells as free DNA elements. Each module contained 50-bp flanking regions
compatible with adjacent sequence, according to a design. Importantly, as in the
case of previous works with S. cerevisiae, Y. lipolytica native recombination mech-
anisms correctly assembled pre-designed DNA construction. After examining the
replicative character and transferability of the Yl-YAC, the authors set at trans-
ferring both xylose and cellobiose utilization pathways to Y. lipolytica within the
newly constructed Yl-YAC. A complete construction comprised three key genes for
xylose (XYL1, XYL2 and XKS1) and three genes for cellobiose utilization (CBP1,
CDT1 and scPGM2), of total length 23 kb.
5 Summary
A great variety of other modular cloning and DNA parts assembly methods have
been developed, that were not covered by the main part of this chapter. For specific
reasons, those methods did not gain popularity in the yeast research community.
One of such methods is ligase cycling reaction (LCR) that employs single-stranded
bridging oligomers (partly complementary to adjacent bioparts) that guide correct
hybridization of DNA fragments. The bioparts to be fused must be phosphorylated
at 5’ends (either by kinase or by amplification with 5’phosphorylated primers).
The method is amenable to modularity by possible use of a 40-bp scaffold of
oligonucleotide connectors (complementary to adjacent elements) which can be
reused in alternative assembly design (Kok et al. 2014). The LCR is executed by
hot-start thermostable ligase and polymerase under cycling profile of denaturation,
annealing of the fragments and ligation at 66 °C. As reported, up to 20 bioparts of
total length 20 kbp can be assembled in a single reaction without leaving scars at
the fusion sites. The other such technique is circular polymerase extension cloning
(CPEC) conducted with DNA polymerase solely (Quan and Tian 2009). The frag-
ments must bear 15–25 bp OLs to anneal correctly. CPEC was reported to be
highly efficient, but only up to 4 bioparts.
Irrespectively of technical details, modular cloning and standardized bioparts
assembly strategies contribute to great progress within the field of genetic engi-
neering of yeast. Their exploitation allows the researchers to focus on addressing
the scientific questions, and to expand the scope of experiments to accurately test
hypotheses, rather than execution of tedious laboratory procedures. Modularity and
standardization facilitate wide-spread collaboration within the yeast community
and speed-up accomplishing of new goals.
Acknowledgements PKW was financially supported by the Ministry of Sciences and Higher Edu-
cation in Poland project: DI 2017 000947. MK was financially supported by the Ministry of Sciences
and Higher Education in Poland project: DI2017 001047.
58 P. Korpys-Woźniak et al.
References
Agmon N et al (2015) Yeast Golden Gate (yGG) for the efficient assembly of S. cerevisiae tran-
scription units. ACS Synthetic Biology. American Chemical Society 4(7):853–859. https://doi.
org/10.1021/sb500372z
Alberti S, Gitler AD, Lindquist S (2007) A suite of Gateway® cloning vectors for high-throughput
genetic analysis in Saccharomyces cerevisiae. Yeast 24(10):913–919. https://doi.org/10.1002/
yea.1502
Anderson JC et al (2010) BglBricks: a flexible standard for biological part assembly. J Biol Eng
4:1–12. https://doi.org/10.1186/1754-1611-4-1
Anthony RA, Liebman SW (1995) Alterations in ribosomal protein RPS28 can diversely affect
translational accuracy in Saccharomyces cerevisiae. Genetics 140(4):1247–1258
Berrow NS et al (2007) A versatile ligation-independent cloning method suitable for high-
throughput expression screening applications. Nucleic Acids Res 35(6):e45. https://doi.org/10.
1093/nar/gkm047
Bhutada G et al (2017) Sugar versus fat: elimination of glycogen storage improves lipid accu-
mulation in Yarrowia lipolytica. FEMS Yeast Res 17(3):1–10. https://doi.org/10.1093/femsyr/
fox020
Casini A et al (2013) One-pot DNA construction for synthetic biology: the modular overlap-
directed assembly with linkers (MODAL) strategy. Nucleic Acids Res 42(1):1–13. https://doi.
org/10.1093/nar/gkt915
Cataldo VF et al (2020) Genomic integration of unclonable gene expression cassettes in Saccha-
romyces cerevisiae using rapid cloning-free workflows. MicrobiologyOpen 9(3):1–10. https://
doi.org/10.1002/mbo3.978
Celińska E et al (2017) Golden gate assembly system dedicated to complex pathway manipulation
in Yarrowia lipolytica. Microbial Biotechnology. Wiley 10(2):450–455. https://doi.org/10.1111/
1751-7915.12605
Celińska E et al (2018) Robust signal peptides for protein secretion in Yarrowia lipolytica: identifi-
cation and characterization of novel secretory tags. Applied Microbiology and Biotechnology.
Springer Verlag 102(12):5221–5233. https://doi.org/10.1007/s00253-018-8966-9
Celińska E et al (2020) Optimization of Yarrowia lipolytica-based consolidated biocatalyst through
synthetic biology approach: transcription units and signal peptides shuffling. Appl Microbiol
Biotechnol 104(13):5845–5859. https://doi.org/10.1007/s00253-020-10644-6
Celińska E, Grajek W (2013) A novel multigene expression construct for modification of glycerol
metabolism in Yarrowia lipolytica. Microb Cell Fact 12(1):1–16. https://doi.org/10.1186/1475-
2859-12-102
Cha-aim K et al (2009) Reliable fusion PCR mediated by GC-rich overlap sequences. Gene.
Elsevier B.V. 434(1–2):43–49. https://doi.org/10.1016/j.gene.2008.12.014
Chang SW et al (2005) Multiple mutagenesis of the Candida rugosa LIP1 gene and optimum
production of recombinant LIP1 expressed in Pichia pastoris. Appl Microbiol Biotechnol
67(2):215–224. https://doi.org/10.1007/s00253-004-1815-z
Chuang J, Boeke JD, Mitchell LA (2018) Coupling yeast golden gate and VEGAS for efficient
assembly of the violacein pathway in Saccharomyces cerevisiae. Methods Mol Biol 1671:211–
225. https://doi.org/10.1007/978-1-4939-7295-1_14
Dong B et al (2007) An improved method of gene synthesis based on DNA works software and
overlap extension PCR. Mol Biotechnol 37(3):195–200. https://doi.org/10.1007/s12033-007-
0039-8
Dueñas-Santero E et al (2019) A new toolkit for gene tagging in Candida albicans containing
recyclable markers. PLoS ONE 14(7):1–17. https://doi.org/10.1371/journal.pone.0219715
Dulermo R et al (2017) Using a vector pool containing variable-strength promoters to optimize
protein production in Yarrowia lipolytica. Microb Cell Fact. BioMed Central Ltd. 16(1). https://
doi.org/10.1186/s12934-017-0647-3
Construction and Assembly of Standardized Biobricks … 59
Eldin P et al (1997) High-level secretion of two antibody single chain Fv fragments by Pichia
pastoris. J Immunol Methods 201(1):67–75. https://doi.org/10.1016/S0022-1759(96)00213-X
Engler C, Kandzia R, Marillonnet S (2008) A one pot, one step, precision cloning method with
high throughput capability. PLoS ONE 3(11). https://doi.org/10.1371/journal.pone.0003647
Esposito D et al (2005) Gateway cloning is compatible with protein secretion from Pichia pastoris.
Protein Expr Purif 40(2):424–428. https://doi.org/10.1016/j.pep.2004.12.006
Funk M et al (2002) Vector systems for heterologous expression of proteins in Saccha-
romyces cerevisiae. Methods Enzymol 350(1994):248–257. https://doi.org/10.1016/S0076-687
9(02)50967-8
Gan R, Jewett MC (2014) A combined cell-free transcription-translation system from Saccha-
romyces cerevisiae for rapid and robust protein synthesis. Biotechnol J 9(5):641–651. https://
doi.org/10.1002/biot.201300545
Gao S et al (2014) One-step integration of multiple genes into the oleaginous yeast Yarrowia lipoly-
tica. Biotechnol Lett. Kluwer Academic Publishers 36(12):2523–2528. https://doi.org/10.1007/
s10529-014-1634-y
Gassler T et al (2019) CRISPR/Cas9-mediated homology-directed genome editing in Pichia pas-
toris. Methods Mol Biol (Clifton, N.J.). United States, 1923 211–225. https://doi.org/10.1007/
978-1-4939-9024-5_9
Gibson DG (2009) Synthesis of DNA fragments in yeast by one-step assembly of overlapping
oligonucleotides. Nucleic Acids Res 37(20):6984–6990. https://doi.org/10.1093/nar/gkp687
Gibson DG et al (2009) Enzymatic assembly of DNA molecules up to several hundred kilobases.
Nat Methods 6(5):343–345. https://doi.org/10.1038/nmeth.1318
Giuraniuc CV, MacPherson M, Saka Y (2013) Gateway vectors for efficient artificial gene assem-
bly in vitro and expression in yeast Saccharomyces cerevisiae. PLoS ONE 8(5). https://doi.org/
10.1371/journal.pone.0064419
Guo W et al (2019) An improved overlap extension PCR for simultaneous multiple sites large frag-
ments insertion, deletion and substitution. Sci Rep. Springer US 9(1):1–6. https://doi.org/10.
1038/s41598-019-52122-8
Guo Y et al (2015) YeastFab: The design and construction of standard biological parts for
metabolic engineering in Saccharomyces cerevisiae. Nucleic Acids Res 43(13):e88. https://doi.
org/10.1093/nar/gkv464
Guo Z-P et al (2020) An artificial chromosome ylAC enables efficient assembly of multiple genes
in Yarrowia lipolytica for biomanufacturing. Commun Biol. Springer US 3(1). https://doi.org/
10.1038/s42003-020-0936-y
Hartley JL, Temple GF, Brasch MA (2000) DNA cloning using in vitro site-specific recombination.
Genome Res 10(11):1788–1795. https://doi.org/10.1101/gr.143000
Ho SN et al (1989) Site-directed mutagenesis by overlap extension using the polymerase chain
reaction. Gene 77(1):51–59. https://doi.org/10.1016/0378-1119(89)90358-2
Holkenbrink C et al (2018) EasyCloneYALI: CRISPR/Cas9-based synthetic toolbox for engineer-
ing of the yeast Yarrowia lipolytica. Biotechnol J. Germany 1–23. https://doi.org/10.1002/biot.
201700543
Horton RM et al (1989) Engineering hybrid genes without the use of restriction enzymes: gene
splicing by overlap extension. Gene 77(1):61–68. https://doi.org/10.1016/0378-1119(89)903
59-4
Jensen NB et al (2014) EasyClone: Method for iterative chromosomal integration of multiple genes
in Saccharomyces cerevisiae. FEMS Yeast Res. Blackwell Publishing Ltd, 14(2):238–248.
https://doi.org/10.1111/1567-1364.12118
Jessop-Fabre MM et al (2016) EasyClone-MarkerFree: a vector toolkit for marker-less integration
of genes into Saccharomyces cerevisiae via CRISPR-Cas9. Biotechnol J. Wiley-VCH Verlag
11(8):1110–1117. https://doi.org/10.1002/biot.201600147
Jeya M et al (2009) Cloning and expression of a GH11 xylanase gene from Aspergillus fumigatus
MKU1 in Pichia pastoris. J Biosci Bioeng. The Society for Biotechnology, Japan 108(1):24–
29. https://doi.org/10.1016/j.jbiosc.2009.02.003
60 P. Korpys-Woźniak et al.
Jiang H, Wood KV, Morgan JA (2005) Metabolic engineering of the phenylpropanoid pathway in
Saccharomyces cerevisiae. Appl Environ Microbiol 71(6):2962–2969. https://doi.org/10.1128/
AEM.71.6.2962-2969.2005
Kadkhodaei S et al (2016) Multiple overlap extension PCR (MOE-PCR): an effective technical
shortcut to high throughput synthetic biology. RSC Adv 6(71):66682–66694. https://doi.org/
10.1039/c6ra13172g
Kildegaard KR et al (2017) Engineering of Yarrowia lipolytica for production of astaxanthin.
Synth Syst Biotechnol. KeAi Communications Co. 2(4):287–294. https://doi.org/10.1016/j.syn
bio.2017.10.002
Kildegaard KR et al (2019) CRISPR/Cas9-RNA interference system for combinatorial metabolic
engineering of Saccharomyces cerevisiae. Yeast 36(5):237–247. https://doi.org/10.1002/yea.
3390
Knight T (2003) Idempotent vector design for standard assembly of Biobricks, MIT Libraries.
http://hdl.handle.net/1721.1/21168
Kok SD et al (2014) Rapid and reliable DNA assembly via ligase cycling reaction. ACS Synth Biol
3(2):97–106. https://doi.org/10.1021/sb4001992
Korpys-Woźniak P et al (2020) Impact of overproduced heterologous protein characteristics on
physiological response in Yarrowia lipolytica steady-state-maintained continuous cultures.
Appl Microbiol Biotechnol. https://doi.org/10.1007/s00253-020-10937-w
Kozak BU et al (2014) Replacement of the Saccharomyces cerevisiae acetyl-CoA synthetases by
alternative pathways for cytosolic acetyl-CoA synthesis. Metab Eng. Elsevier 21:46–59. https://
doi.org/10.1016/j.ymben.2013.11.005
Larroude M et al (2018) A synthetic biology approach to transform Yarrowia lipolytica into a com-
petitive biotechnological producer of β-carotene. Biotechnol Bioeng 115(2):464–472. https://
doi.org/10.1002/bit.26473
Larroude M et al (2019) A modular Golden Gate toolkit for Yarrowia lipolytica synthetic biology.
Microb Biotechnol 12(720824):1249–1259. https://doi.org/10.1111/1751-7915.13427
Larroude M et al (2020) A set of Yarrowia lipolytica CRISPR/Cas9 vectors for exploiting wild-
type strain diversity. Biotechnol Lett. Springer Netherlands 42(5):773–785. https://doi.org/10.
1007/s10529-020-02805-4
Lebel K, MacPherson S, Turcotte B (2006) New tools for phenotypic analysis in Candida albi-
cans: the WAR1 gene confers resistance to sorbate. Yeast (Chichester, England) 23(4):249–259.
https://doi.org/10.1002/yea.1346
Leguia M et al (2013) 2ab assembly: A methodology for automatable, high-throughput assembly
of standard biological parts. J Biol Eng 7(1):1. https://doi.org/10.1186/1754-1611-7-2
Leplat C, Nicaud J-MM, Rossignol T (2018) Overexpression screen reveals transcription factors
involved in lipid accumulation in Yarrowia lipolytica. FEMS Yeast Res. England 18(5):1–9.
https://doi.org/10.1093/femsyr/foy037
Leplat C, Nicaud JM, Rossignol T (2015) High-throughput transformation method for Yarrowia
lipolytica mutant library screening. FEMS Yeast Res. Oxford University Press 15(6):1–9.
https://doi.org/10.1093/femsyr/fov052
Li MZ, Elledge SJ (2005) ‘MAGIC, an in vivo genetic method for the rapid construction of recom-
binant DNA molecules. Nature genetics. United States, 37(3):311–319. https://doi.org/10.1038/
ng1505g
Liu D et al (2019) Constructing yeast chimeric pathways to boost lipophilic terpene synthesis. ACS
Synth Biol. United States 8(4):724–733. https://doi.org/10.1021/acssynbio.8b00360
Liu Q et al (1998) The univector plasmid-fusion system, a method for rapid construction of recom-
binant DNA without restriction enzymes. Curr Biol. England 8(24):1300–1309. https://doi.org/
10.1016/s0960-9822(07)00560-x
Lund AM et al (2014) A versatile system for USER cloning-based assembly of expression vec-
tors for mammalian cell engineering. PLoS ONE 9(5). https://doi.org/10.1371/journal.pone.
0096693
Construction and Assembly of Standardized Biobricks … 61
Abstract
Yarrowia lipolytica is a safe and robust yeast to efficiently use lipid as the sole
carbon source, which provides us opportunities for biomanufacturing of a series
of high-value products from cost-effective agriculture feedstocks such as plant
oils and animal fats. Y. lipolytica has a unique propensity for high flux through
tricarboxylic acid (TCA) cycle intermediates and biological precursors such as
acetyl-CoA and malonyl-CoA that can be diverted into a variety of heterolo-
gous value-added bioproducts. With recent advances in metabolic engineering
and synthetic biology tools, the potential of using Y. lipolytica for biomanufac-
turing of high-value products has been expanded. Examples include industrial
enzymes, extracellular proteins, fatty alcohols, wax esters, long-chain diacids,
omega-3 fatty acids, and carotenoids. For large-scale biomanufacturing using
oils/fats as substrate, the poor mixing and mass transfer caused by the insolu-
bility of substrates in an aqueous medium is one of the major challenges that
have to be addressed in addition to the pathway engineering and optimization
for both fatty acid biosynthesis and conversion. The multi-phase computational
fluid dynamics (CFD) simulation can be used as a powerful tool for analy-
sis of mixing and mass transfer behaviors in bioreactors and further guide the
bioreactor design and optimization of operating conditions. Cell morphology
has a profound effect on cell growth, oil substrate uptake, and product forma-
tion. Both PKA and cAMP-dependent signaling pathways are involved in the
dimorphic transition in Y. lipolytica. Maintaining the dimorphic yeast shape via
Na Liu and Ya-Hue Valerie Soong are contribute equally to the paper.
morphology engineering strategies has been explored. This chapter also intro-
duced several examples of how we combined cell morphology engineering,
metabolic pathway optimization, and bioreaction engineering to significantly
improve the production of intracellular lipids, citric acid, and wax esters from
plant oils. Potential strategies for further improving the biosynthesis efficiency
via transporter engineering in yeast were also introduced.
Keywords
1 Introduction
yeast an important biotechnological yeast, which has been proposed for the treat-
ment of petroleum oil-polluted soil or water (Ledesma-Amaro and Nicaud 2016b).
In addition, Y. lipolytica has been efficiently cultivated on oils or fats for pro-
ducing many intra- or extracellular metabolites of industrial significance (Fickers
et al. 2005a, b; Papanikolaou et al. 2007). Vegetable oils and animal fats were
reported as promising substrates for biosurfactant production, as the oily or fatty
carbon sources are consumed by microorganisms could work as a building block
for biosurfactant synthesis (Goncalves et al. 2014). Various studies have indicated
that the production and secretion of lipases in Y. lipolytica strains are stimulated
by the presence of long-chain fatty acids (LCFAs) (Fickers et al. 2003), plant oils
(Braga et al. 2012; Deive et al. 2010; Kebabci and Cihangir 2012; Najjar, et al.
2011), and animal fats (Kamzolova et al. 2005) in the culture medium. Conversely,
glucose in the medium might repress the production of lipases (Liu et al. 2021).
The oleaginous yeast Y. lipolytica has been reported capable of accumulating a
significant amount of intracellular lipid, stored in lipid bodies, during growth on
vegetable oil (Najjar et al. 2011) and animal fats or their industrial derivatives
(Papanikolaou et al. 2002, 2007). With recent improvements in the synthetic biol-
ogy tools, the industrial potential of Y. lipolytica has been expanded to include
organic acids such as citric acid, α-ketoglutaric acid, and itaconic acid, enzymes
such as lipases, RNase, and esterase, lipids and lipid-derived compounds such as
biodiesel, dicarboxylic acids, and biosurfactants (Liu et al. 2021).
The great potential for industrial applications of Y. lipolytica has driven the
development of metabolic engineering tools for it as well as the basic research to
understand its physiological features. In recent years, different metabolic engineer-
ing tools and strategies have been created and applied in Y. lipolytica, which has
further expanded the application of this yeast.
This chapter summarizes the degradation of oil and fat substrates by Y. lipoly-
tica, fermentation and cellular engineering efforts of the Y. lipolytica metabolic
engineering program to improve the utilization efficiency and bioconversion of
oils and fats. The distribution of oil–water mixing in stirred bioreactor was studied
by computational fluid dynamics (CFD) simulation. The CFD results were further
used to guide the optimization of the bioreactor design and operation conditions
during bioreactor fermentation experiments. The effect of cell morphology on the
fermentation with lipid substrates was examined by creating the hyphal strain and
the yeast-like strain through overexpression and disruption of the gene YlMHY1
in the wild-type Y. lipolytica ATCC20362, respectively. The strategies of biore-
actor design and operation conditions and cell morphology engineering can also
be used for the production of other high-value products (e.g., fatty alcohol, wax
esters). The metabolic engineering for the biosynthesis of wax esters by Y. lipolyt-
ica from oils and fats is introduced as a case study. This chapter also summarizes
the engineering of transporters in yeast that are used to enhance bioproduction.
66 N. Liu et al.
Y. lipolytica is often found in environments with the presence of plant oils or ani-
mal fats, being able to assimilate FAs, oils, and fats efficiently. The route of oils
or fats substrates into the cells induces several modifications of the substrates to
improve their accessibility. The initial challenge is a contact between the poorly
water-miscible substrate, and the cell surface. Y. lipolytica can produce surfactants,
amphiphilic compounds consist of hydrophilic and hydrophobic moieties, which
can reduce surface and interfacial tensions in aqueous media and hydrophobic sub-
strates and also reduce the size of the oils/fats droplets, thus increasing the contact
between substrates and cell surface (Fickers et al. 2005a, b). In bioreactor fermen-
tation, this step could be additionally facilitated by powerful agitation. Moreover,
surfactants can also facilitate cell adhesion to oils/fats droplets (Beopoulos et al.
2009).
The main component of oils and fats is triglycerides (TAGs) which cannot be
directly uptake by Y. lipolytica, as no TAGs transporters have been identified (Liu
et al. 2021). Y. lipolytica secretes extracellular lipases (LIP2, LIP7 and LIP8),
which help hydrolyze TAGs to form glycerol and the respective free fatty acids
(FFAs). Grown in oils/fats leads to structural changes on the cells surface of Y.
lipolytica and results in the formation of protrusions that enable cells to take up
FFAs from the medium (Mlickova et al. 2004). The action of biosurfactant and
lipase occurs progressively, the formation of numerous small-sized droplets facil-
itating the surface-mediated substrate transport (Beopoulos et al. 2009). Once in
the cytoplasm, FFA becomes activated by conversion into a fatty acyl-CoA by
the enzyme acyl-CoA synthetase FAA1. Then the fatty acyl-CoA can enter the
Kennedy pathway to be stored into the lipid body as TAGs (formed by DGA1 and
DGA2) (Abghari and Chen 2014), as precursors for the synthesis of fatty-acid-
derived products, or be transported into the peroxisome to carry out β-oxidation
(POX1-6, MFE1, and POT1). FFAs can be released from the TAGs through intra-
cellular lipases (TGL3 and TGL4), and thereafter, they can be activated to form
fatty acyl-CoA then transported into the peroxisome (Ledesma-Amaro et al. 2016).
The β-oxidation pathway, the main pathway for the breakdown of these fatty acyl-
CoA esters, is a four-reaction cycle. After each cycle, the CoA ester of FA gets
two carbons shorter and one molecule of acetyl-CoA is released. In Y. lipolytica,
the first and most important step of β-oxidation is carried out by six acyl-CoA
oxidases (POX1-6). The strain with POX1-6 genes knock-out is unable to degrade
FFAs resulting in a lipid accumulation (Wang et al. 1999). The second and third
steps in β-oxidation are catalyzed by the multi-functional enzyme (MFE). In con-
trast to acyl-CoA oxidases, which are encoded by six genes, MFE is encoded by
a single gene MFE1. The deletion of the MFE1 gene has been extensively studied
in Y. lipolytica for lipid production due to its technical simplicity (Blazeck et al.
2014). The fourth and last step is carried out by a peroxisomal thiolase POT1. As
β-oxidation takes place in the peroxisome, this pathway has been blocked by dis-
rupting the genes, PEX3, PEX10, and PEX11, involved in peroxisome biogenesis
Cellular Engineering of Yarrowia lipolytica … 67
Fig. 1 a Culture medium (blue) and oil (yellow) in the flask, oil stay on the top of the aque-
ous medium due to hydrophobic nature and light density. b Flask culture of Y. lipolytica with oil
substrate. c The assimilation of oils by Y. lipolytica: (i) Oil–water mixing was facilitated by surfac-
tant from Y. lipolytica; (ii) TAGs (oils/fats) are cleaved by extracellular lipases to give FFAs; (iii)
oils/FAAs droplets bind onto the cell surface; (iv) FAAs enter into cell interior via transport/export
mechanisms; (v) FAAs in the cytoplasm can be activated by cytoplasmic fatty acyl-CoA synthase
(FAA1); (vi) the activated fatty acid (fatty acyl-CoA) directly transported into the peroxisome for
β-oxidation degradation, (vii) converted into fatty acids derived products, or (viii) store into lipid
bodies as TAGs; (ix) lipid bodies could work as storage pools for fatty acids derived products,
TAG in lipid bodies could be hydrolyzed by lipases to release FFAs. Dash lines: Putative route,
not confirmed
(Ledesma-Amaro and Nicaud 2016b). The cycle is repeated several times, the-
oretically until the fatty acyl-CoA been completely breakdown. The acetyl-CoA
formed in β-oxidation is a key intracellular metabolite, which plays a major role
in various metabolic pathways that link catabolism and anabolism such as TCA
cycle, biosynthesis of acetyl-CoA-derived products (e.g., polyphenol, carotenoids)
(Fig. 1) (Chen et al. 2012).
the aqueous medium to the individual cells surface is still the critical and limiting
step. Hence, optimization of bioreactor design (e.g., impeller types, baffled) and
operation conditions (e.g., agitation speed, gas flow rate) is regarded as the promis-
ing solution to the mixing and mass transfer issues in oils/fats substrate involved
bioprocesses.
Fig. 2 Two typical impellers used for fermentation. a The blades on Rushton impellers are flat and
set vertically along an agitation shaft, which produces a unidirectional radial flow. b The blades on
pitched-blade impellers (top and medium) are flat and set at ∼45° angles, which produces a simul-
taneous axial and radial flow. The combination of pitched-blades and Rushton impellers provides
better overall mixing and creates a higher mass transfer rate
In our recent bioreactor fermentation experiments, the cell growth and citric acid
production of wild-type Y. lipolytica ATCC20362 on soybean oil substrate were
studied as a representation of the overall mixing and mass transfer, since better
mixing and mass transfer usually led to faster cell growth and/or higher prod-
ucts formation. To compare the effect of impellers on mixing and mass transfer,
two sets of impellers were tested in 1-L bioreactors, with one set of impellers con-
sisted of three evenly-spaced Rushton impellers (Fig. 2a), while the other consisted
of two pitched-blade impellers on top and one Rushton impeller at the bottom
(Fig. 2b). The agitation speed through fermentation was controlled between 500
and 1400 rpm. The bioreactor is equipped with three impellers, which has one
more impeller as compared to the bioreactor geometry in CFD simulation. The
third impeller was added to the top region of the bioreactor to break up and dis-
perse the oil droplets accumulated in the upper region. Furthermore, two maximal
agitation speeds, 1000 rpm and 1400 rpm, were tested in the bioreactor equipped
with two pitched-blade impellers on top and one Rushton impeller at the bot-
tom to examine if more powerful input provides better mixing and improves the
70 N. Liu et al.
40 80
30 60
20 40
10 20
0 0
1.2 0.3
0.2
0.6
0.1
0.0 0.0
Fig. 3 Fed-batch fermentation of Y. lipolytica ATCC20362 with soybean oil substrate in biore-
actors equipped with different impellers and under different maximal agitation speed. a Dry cells
weight (DCW), b citric acid titer, c specific citric acid, d citric acid yield at 144 h
Two major signal transduction pathways are involved in the regulation of the yeast-
to-hyphal transition in dimorphic yeast. One is the mitogen-activated protein kinase
(MAPK) pathway (Fig. 4). MAPK signaling cascades are multi-functional path-
ways that are evolutionarily conserved in all eukaryotic cells (Ligterink and Hirt
2001; Xu et al. 2017). The MAPK pathway is initiated by specific extracellular
cues, amplifies, and integrates the signals through the activation of a particular
MAPK following the consecutive of a MAPK kinase kinase (MAPKKK) and a
MAPK kinase (MAPKK), and consequently elicits an appropriate physiological
response. Typically, the MAPKKK is activated by interactions with a small GTPase
and/or phosphorylation by protein kinases downstream from cell surface receptors
transducing signal downstream (Cuevas et al. 2007; Zhang and Liu 2002). Sev-
eral types of protein kinases (i.e., YlCLA4p, YlSTE11p, YlSTE7p) contributed
to the MAPK pathway in Y. lipolytica have been characterized (Cervantes-Chávez
and Ruiz-Herrera 2006; Martínez-Soto and Ruiz-Herrera 2017; Martinez-Vazquez
et al. 2013; Szabo 2001). YlCLA4p is highly homologous to CLA4 protein kinase
of Candida albicans and Saccharomyces cerevisiae, which are members of the
p21-activated kinase (PAK) family containing conserved internal Cdc42p-binding
regions (Bartholomew and Hardy 2009; Lai et al. 2012). Deletion of YlCLA4 in
Y. lipolytica is not lethal but completely loses the ability to form filaments or
invade agar (Szabo 2001). Disruption of the YlSTE11, which exhibits high homol-
ogy to fungal MAPKKK, loses the capacity to mate and grows constitutively in
the yeast-like form (Cervantes-Chávez and Ruiz-Herrera 2006). The expression
level of YlSTE7 encoding MAPKK is increased during yeast-to-hyphal transition
(Chaleff and Tatchell 1985; Leberer et al. 1996; Martinez-Vazquez et al. 2013).
Transcription factor is a downstream effector of upstream signaling pathway.
YlZNC1p contains a Zn(II)2 C6 fungal-type zinc finger DNA-binding domain and
a leucine zipper domain that acts as a transcription factor repressing hyphal forma-
tion. YlZNC1p is involved in the MAPK pathway via the regulation of YlCLA4,
Cellular Engineering of Yarrowia lipolytica … 73
Fig. 4 Simplified scheme describing the dimorphic transition in Y. lipolytica. Both MAPK and
cAMP-dependent signaling pathways are involved in the yeast-to-hyphal transition but have oppos-
ing actions. The activation of the MAPK pathway induces filamentation. The activation of the
cAMP-dependent pathway represses the filamentous growth. Abbreviations: PAK, p21-activated
kinase; MAPKKK, MAP kinase kinase kinase; MAPKK, MAP kinase kinase; MAPK, MAP
kinase; TCS, TEC1 consensus binding sequence; FRE, filamentation response element; GPCR,
G-protein coupled receptor; PKA, protein kinase A
The other is the cAMP-dependent protein kinase A (PKA) pathway (Fig. 4). In
yeast and filamentous fungi, this pathway takes part in the pathogenesis, cellular
morphogenesis, nutrient sensing and acquisition, sexual reproduction, and stress
responses (D’Souza and Heitman 2001; Hogan and Sundstrom 2009; Kronstad
et al. 2011). Ras subfamily proteins, the small monomeric GTP-binding proteins,
activate adenylyl cyclase CYR1 by interaction with its Ras associating domain to
synthesize cAMP, which binds to the regulatory subunits of PKA and releases the
catalytic subunits. The unleashed active kinases then phosphorylate a diverse set of
substrates resulting in corresponding biological responses (Cao et al. 2017; Weeks
and Spiegelman 2003).
Unlike other yeast species, which usually has one or two Ras proteins, Y.
lipolytica possesses three Ras proteins, particularly YlRAS1p and YlRAS2p, are
implicated in the control of dimorphism. In comparison with YlRAS1p, YlRAS2p
plays a major role in the yeast-to-hyphal transition. The expression of YlRAS2
is increased dramatically at the transcription level during mycelial development.
Additionally, mutants in the YlRAS2 exhibit a severe defeat in pseudohyphal or
hyphal growth (Li, et al. 2014). The second messenger cAMP is an important
mediator to regulate the PKA activation. The increase of cytosolic cAMP concen-
tration, either by adenylyl cyclase activation or by entry of exogenous nucleotide
into the cell, inhibits the PKA pathway and the dimorphic transition in Y. lipolyt-
ica (Cervantes-Chávez and Ruiz-Herrera 2007; Johnson et al. 2001; Taylor et al.
2004). The core component of this pathway is PKA, which is a heterotetramer
consisting of two catalytic (cPKA) and two regulatory subunits (rPKA). The cat-
alytic subunits are encoded by the YlTPK1 gene and the regulatory subunits are
encoded by the YlRKA1 (homologous to S. cerevisiae BCY1) (Cervantes-Chávez,
et al. 2009; Toda et al. 1987). It has been previously reported that the YlRKA1
gene was up-regulated at the transcriptional level under conditions that induce
mycelial morphology, indicating that dimorphic transition is regulated by the PKA
pathway via YlPKA1p in Y. lipolytica (Cervantes-Chávez and Ruiz-Herrera 2007).
Expression of YlPKA1p in PKA cascade was also mediated by YlZNC1p. Hence,
YlZNC1p acting through both MAPK and PKA pathways represses the yeast-to-
hyphal transition. Mutant strains deleted in the YlTPK1 grew constitutively in the
mycelial form, whereas YlSTE11 and YlTPK1 double mutants grew constitutively
in the yeast form, suggesting that the default growth pattern of the Y. lipolytica is a
yeast-like form (Cervantes-Chávez et al. 2009). Surprisingly, all these data provide
evidence that the opposite actions of the MAPK and PKA pathways in regulating
Y. lipolytica dimorphism, in contrast to S. cerevisiae and C. albicans where both
pathways act together to regulate the dimorphic switching (Biswas et al. 2007).
Multiple genes, including YlHOY1 and YlMHY1, are also responsible for the
filamentous growth of Y. lipolytica. The YlHOY1 encoding a putative nuclear pro-
tein with a homeodomain function as transcriptional regulatory protein. Disruption
of YlHOY1 results in a defect in filamentation (Torres-Guzman and Domínguez
Cellular Engineering of Yarrowia lipolytica … 75
1997). Like the YlHOY1 gene, YlMHY1p is localized to the nucleus during fila-
mentous growth and acts as a transcription factor. YlMHY1 encoding a C2 H2 -type
zinc finger protein, YlMHY1p, exhibits strong homology to the S. cerevisiae stress
response factors MSN2p and MSN4p, which is involved in the regulation of mor-
phogenesis (Kobayashi and McENTEE 1993; Marchler et al. 1993). Like these
factors, YlMHY1p specifically recognizes and binds to putative cis-acting DNA
stress response elements (STREs) located on the upstream of a number of genes
conferring tolerance to a variety of stress conditions, such as heat shock, car-
bon source starvation, osmotic stress, and oxidative shock (Treger et al. 1998).
Transcription of YlMHY1 is dramatically increased during the dimorphic transi-
tion in Y. lipolytica. Deletion of YlMHY1 is unable to undergo mycelial growth,
indicating that YlMHY1p promotes hyphal development. Interestingly, overexpres-
sion of YlMHY1 in YlRAS2 mutant cells form abundant pseudohyphae or hyphae,
while overexpression of YlRAS2 in YlMHY1 mutant cells fail to induce filamentous
growth (Li et al. 2014). YlMHY1 is a gene whose overexpression could restore
hyphal growth and is required for Y. lipolytica YlRAS2p function, like S. cere-
visiae Ras protein does to MSN2p and MSN4p. Therefore, the transcription factor
YlMHY1p may function as a signal transducer downstream of YlRAS2p in the
control of the dimorphic transition.
Besides, the transcriptome analysis reveals the downstream target regulated
by YlMHY1p. These downstream genes encode proteins that are similar to the
flocculin FLO11, cell wall mannoprotein TIR3 and agglutinin AGA1 in S. cere-
visiae and HYR1 in C. albicans (Bailey et al. 1996; Rupp et al. 1999). Therefore,
YlMHY1p mediates the expression of a large group of cell wall proteins and
enzymes involved in cell wall maintenance. In addition to cell wall-related genes,
YlMHY1p also regulates genes involved in nutrient uptake, protein processing,
and lipid metabolism (Wu et al. 2020). YlMHY1p has multiple cellular functions.
2020). Unlike the Y. lipolytica ATCC20362 wild-type strain, cells of the YlMHY1
deletion are unable to form filaments (Fig. 5a). Cells with YlMHY1 overexpression
form filaments longer than those of control cells (Fig. 5a), higher dry cell weight
(DCW) (Fig. 5b), decreased citric acid titer (Fig. 5c), and low lipid accumulation
(Fig. 5d) in liquid medium using glucose, soybean oil or waste cooking oil as
main carbon source. Citric acid is defined as a key precursor in lipid production.
Deletion of YlMHY1 in Y. lipolytica exhibits larger lipid bodies (LB), higher citric
acid titer (Fig. 5c), and increased lipid/DCW (Fig. 5d) in 1-L fed-batch fermenta-
tion, suggesting that the morphology arresting in yeast-like form has a beneficial
effect on the lipid-derived product formation and accumulation (Liu et al. 2021).
In comparison with glucose, lipid-derived feedstock (soybean oil and waste cook-
ing oil) benefits both Y. lipolytica cells’ growth and lipid production. Mutations in
YlMHY1 are shown to be more efficient in lipid feedstock utilization. Consistent
with this finding, Wang et al. reported that YlMHY1p regulates lipid biosynthesis
since the Y. lipolytica with YlMHY1 deletion increased carbon flux through lipid
biosynthesis and accumulated more intracellular oil than the wild-type strain dose
(Wang et al. 2018). Therefore, in combination with morphology engineering and
metabolic engineering strategies together with lipid-derived feedstock utilization
is expected to become an efficient and economical approach for constructing Y.
lipolytica cell factories.
(a) (b)
80
Glucose Soybean Oil Waste Cooking Oil
Wild-type YlMHY1- YlMHY1+ 70
Dry Cell Weight (g/L)
60
LB
50
LB 40
LB
30
20
10
0
Wild-type YlMHY1- YlMHY1+
(c) (d)
140 100%
Glucose Soybean Oil Waste Cooking Oil Glucose Soybean Oil Waste Cooking Oil
120
Citric Acid Titer (g/L)
80%
100
Lipid/DCW
80 60%
60 40%
40
20%
20
0 0%
Wild-type YlMHY1- YlMHY1+ Wild-type YlMHY1- YlMHY1+
Fig. 5 Cellular responses to different carbon sources for Y. lipolytica strains with YlMHY1 deletion
and overexpression. a Cell morphology under microscope: Cells of strains ATCC20362 (wild-
type), YlMHY1 knock-out (YlMHY1−), and YlMHY1 overexpression (YlMHY1+ ) were grown
at 30 °C for 120 h in liquid media containing glucose as the main carbon source. b dry cell
weight (DCW), c citric acid titer, and d lipid/DCW of Y. lipolytica wild-type, YlMHY1- and
YlMHY1+ strains after 144 h cultivation in 1-L fed-batch fermentation using glucose, soybean oil,
or waste cooking oil as the main carbon sources
Cellular Engineering of Yarrowia lipolytica … 77
Wax esters are widely distributed nature compounds that are found in high evo-
lution plants, algae, microorganisms, and even insects and mammals. Natural
occurring waxes, consisting of fatty acids esterified to long-chain alcohols, are a
group of highly hydrophobic neutral lipids but structurally diverse. Physical prop-
erties and applications of wax esters are varied due to different chain lengths of
the fatty acid and the fatty alcohol components as well as the degree of unsat-
uration that affect melting temperature, oxidation stability, and pressure stability.
Wax esters have a variety of biological functions that provide the protective coat-
ing on the surface so that they are resistant to dehydration, ultraviolet rays, and
pathogens (Jetter and Kunst 2008; Wältermann et al. 2005). Wax esters are used
commercially to serve as a wide range of applications, such as cosmetics, printing
inks, lubricants, coatings, pharmaceuticals, and the food industries (Doan et al.
2017; Fiume et al. 2015; Petersson et al. 2005). Although the wax esters are found
in nature universally, the abundance of wax ester source is still low because only
a few organisms such as Jojoba plant (Simmondsia chinensis) and sperm whale
(Physeter macrocephalus) are capable of accumulating a considerable amount of
intracellular wax esters. Currently, wax esters are in a short supply due to the hunt-
ing ban for sperm whale, the high extraction cost, and the harsh requirement of
agriculture system for Jojoba (Miwa 1971). Therefore, establishment of an efficient
expression platform will be expected to advance the economic feasibility of wax
esters from low-cost substrates as carbon sources, especially from waste cooking
oil.
The biosynthesis and accumulation of TAG and wax esters (WEs) have been
reported in some soil and marine bacteria including Acinetobacter (Ishige et al.
2002; Santala et al. 2014), Marinobacter (Willis et al. 2011), Mycobacterium
(Sirakova et al. 2012), Streptomyces (Röttig et al. 2016), Euglena (Tomiyama
et al. 2017), and Rhodococcus (Round et al. 2019) genera under nitrogen-limited
conditions. In prokaryotes, mechanism of WE synthesis proceeds via sequen-
tial reactions (Fig. 6a): First, the fatty acyl-CoA or fatty acyl-ACP substrate is
reduced to a respective long-chain fatty aldehyde by fatty acyl-CoA reductase
(FAR) in an NADPH-dependent manner; the aldehyde is further reduced to corre-
sponding fatty alcohol by uncharacterized fatty aldehyde reductase (Alvarez 2016;
Mcdaniel et al. 2011). Second, the fatty alcohol is esterified with acyl-CoA through
the existence of a CoA-dependent acyltransferase enzyme known as wax ester
synthase/diacylglycerol acyltransferase (WS/DGAT).
In Y. lipolytica, biosynthesis of fatty alcohols by FAR has been intensively
studied (Madzak 2018; Zeng et al. 2018). Heterologous expression of FAR from
Marinobacter hydrocarbonoclasticus strain VT8 produced 5.75 g/L fatty alcohols
in Y. lipolytica when grown on modified YPD medium containing 91 g/L glucose
78 N. Liu et al.
Fig. 6 Introducing wax ester biosynthesis pathway in Y. lipolytica. a Biosynthesis pathway of wax
esters in Y. lipolytica during the growth on oils/fats as carbon source. b Fatty alcohol and wax ester
titers in the engineered Y. lipolytica Po1f strain, which contained the fatty acyl CoA reductase
(ScFAR, MmFAR or MhFAR) and wax ester synthase (AbWS) expression via plasmid transfor-
mation. c Phenotype and d wax ester titer of engineered Y. lipolytica ATCC20362 strains with
co-expression of MhFAR and AbWS via chromosomal integration grown on glucose, oleic acid,
soybean oil or waste cooking oil as a carbon source for 120 h in a shaking flask. ScFAR from S.
chinensis; MmFAR from Mus musculus; MhFAR from M. hydrocarbonoclasticus strain VT8; AbWS
from A. baylyi ADP1; LB, Lipid bodies
in shaking flask scale (Zhang et al. 2019). A recent report also showed that Y.
lipolytica possessing two genes coding for MhFAR achieved at titers of 5.8 g/L
fatty alcohol production in a minimal glucose media under fed-batch fermentation
(Cordova et al. 2020). Also, Y. lipolytica was engineered to produce fatty acid
methyl esters (FAME) and fatty acid ethyl esters (FAEE) by introducing wax ester
synthase (Gao et al. 2018; Xu et al. 2016).
Cellular Engineering of Yarrowia lipolytica … 79
Recently, several studies have been conducted to achieve significantly greater cell
growth rate and higher production yield of microbial-derived biodiesel (Blazeck
et al. 2014; Darvishi et al. 2017). Other research efforts include metabolic engi-
neering of microorganisms to enhance the conversion of hydrophobic substrates
into value-added, lipid-derived products (Sabirova et al. 2011; Xie 2017; Xue et al.
2013; Yang et al. 2019). In Y. lipolytica, co-expression of MhFAR (M. hydrocarbon-
oclasticus strain VT8) and the AbWS (A. baylyi ADP1) via plasmid transformation
produced the most fatty alcohols and wax esters as compared to the strains co-
expressing the ScFAR (S. chinensis) or the MmFAR (Mus musculus) together with
the AbWS (Fig. 6b). Oil-based substrates have been employed for the improving
production of lipid-derived products by Y. lipolytica (Enshaeieh et al. 2014; Li
et al. 2008). In particular, oleic acid (C18:1) is the major fatty acid present in
Y. lipolytica, whereas palmitic (C16:0) and linoleic (C18:2) acids have also been
detected in high content in the cells (Dobrowolski et al. 2016; Magdouli, Guedri
et al. 2020).
In our recent study (Soong et al. 2021), as shown in Fig. 6c, the enlarged
lipid bodies were produced when the engineered Y. lipolytica (integrating the
MhFAR and AbWS into the chromosome) cells grown on the oleic acid or oil-
containing media, indicating that the accumulation of lipid-derived products could
be improved. In addition, by switching the main carbon source from glucose to
waste cooking oil, the wax ester titer was increased by 70-fold, reaching a maxi-
mum of 7.58 g/L after 120 h of cultivation in a shaking flask (Fig. 6d). Thus, in
the presence of exogenous oleic acid or oil-based substrates (soybean oil or waste
cooking oil) to the culture medium, the engineered Y. lipolytica strain tends to
rely on an external source of fatty acids rather than endogenous de novo fatty acid
biosynthesis for wax ester production. This work opens a door toward the econom-
ical production of lipids, lipid-derived, and lipid-assisted products for applications
as fuels, chemicals, nutraceuticals, and pharmaceuticals.
Transporters are key components for efficient import and export activities. The
transport of the substrates, such as sugars and fatty acids, across the cytoplasmic
membrane into the cell, which links extracellular substrates utilization and intra-
cellular metabolic pathways, is a critical step for consolidated yeast bioprocessing.
Furthermore, the transport of the produced products out of the cell is also critical
for maintaining a high production rate (Hara et al. 2017). This section summa-
rizes the engineering of specific transporters that are used in yeast to enhance
bioproduction.
80 N. Liu et al.
YALI0B00396p, have been identified in Y. lipolytica (Ryu and Trinh 2018). The
yeast S. cerevisiae lacks pentose-specific transporters; it is known to uptake D-
xylose and L-arabinose via endogenous hexose/glucose transporters. The efficiency
of hexose transporters for D-xylose uptake was as follows: HXT7 > HXT5 >
GAL2 HXT1 > HXT4 (Sedlak and Ho 2004). Overexpression of endogenous
hexose transporters and/or heterologous xylose transporters have effectively facil-
itated D-xylose uptake into engineered S. cerevisiae strains (Hector et al. 2008).
Because D-xylose uptake is competitively inhibited by D-glucose, cofermentation
of D-glucose and D-xylose, such as in cellulosic and hemicellulose hydrolysates,
is not cost-efficient (Farwick et al. 2014). Many studies have focused on effec-
tive D-xylose fermentation from mixed sugars by developing more xylose-specific
transporters. Farwick and coworkers (Farwick et al. 2014) engineered hexose
transporters HXT7p and GAL2p and have achieved glucose-insensitive xylose
transporters. These engineering efforts have facilitated the economic production
of value-added products from renewable feedstocks.
Cells cannot directly uptake oils or fats in the format of triacylglycerides (TAGs),
as no known TAG transporters have been identified (Beopoulos et al. 2009). After
hydrolysis of TAGs catalyzed by extracellular lipases, the released free fatty acids
(FFAs), usually are long-chain fatty acids (LCFAs), can be transported into cells by
membrane-bound FA-transporters. In S. cerevisiae, the transport system of LCFAs
has been well characterized. In this yeast, exogenous LCFAs traverse the mem-
brane via the FA-transporter ScFAT1p. Once transported across the membrane,
LCFAs are activated by conversion into a fatty acyl-CoA by the fatty acyl-CoA
synthetases ScFAA1p and ScFAA4p (DiRusso and Black 1999). Y. lipolytica has
LCFAs transport and activation proteins similar to those of S. cerevisiae, but the
transporter YlFAT1p had a different function than ScFat1p, and it is also involved
in the export of FAs from lipid bodies after the TAGs hydrolysis under the catalyza-
tion of intracellular lipases (Dulermo et al. 2014). Once FAs have been exported
from intracellular lipid bodies and activated, they can enter the peroxisomes and
be degraded as a result of β-oxidation cycle. In S. cerevisiae, the membrane-bound
transport heterodimer ScPXA1p/ScPXA2p is responsible for transporting fatty
acyl-CoA into peroxisomes (Shani and Valle 1996). Disruption of the peroxisomal
fatty acyl-CoA transporter ScPXA1p increased the intracellular TAGs accumula-
tion by 14% (Ferreira et al. 2018). In Y. lipolytica, there is a membrane-bound
transport protein pair YlPXA1p (YALI0A06655) and YlPXA2p (YALI0D04246)
corresponding to transporters in S. cerevisiae, which were responsible for the trans-
port of fatty acyl-CoA into peroxisome (Dulermo et al. 2015). In Y. lipolytica,
the utilization of FAs is largely dependent on transporters YlFAT1p, YlPXA1p,
and YlPXA2p. The deletion of YlPXA1and YlPXA2 in Y. lipolytica has increased
the accumulation of FAs when cells were grown in an oleate (C18:1) medium
(Dulermo et al. 2015).
82 N. Liu et al.
7 Conclusions
The GRAS status of Y. lipolytica and its metabolic traits, such as the ability to uti-
lize diverse hydrophobic substrates, high flux through acetyl-CoA, has made the
oleaginous yeast an important host for the production of fuels, commodity chem-
icals, nutraceuticals, and pharmaceuticals that can be derived from acetyl-CoA,
FAs, and lipids. Plant oils and animal fats, especially the low-cost waste oils and
fats, can be the preferred substrates for the biomanufacturing of these products
at low cost and high yield. However, due to the insolubility of oils and fats in
water, poor oil–water mixing and mass transfer is one of the major challenges to
be addressed for the fermentation processes using plant oils or animal fats as sub-
strates. CFD model and simulation were applied to analyze the distribution of oil
droplets in water in a stirred bioreactor, which further guided the bioreactor design
and optimization of operating conditions to significantly improve the fermentation
performance. Controlling cell morphology in a yeast-like form via morphology
engineering enhanced the cell growth on oils/fats and improved the production
of lipid-derived products. However, compared to the hydrophilic substrates, more
work in both cellular engineering and bioreaction engineering of Y. lipolytica
should be conducted in the future to facilitate the fatty acid transport for more effi-
cient extracellular substrate uptake and intracellular bioconversion. In a summary,
with advances in metabolic engineering in various promising microorganisms, oils
Cellular Engineering of Yarrowia lipolytica … 83
or fats can be used as a great substitute for sugars for biomanufacturing a series
of high-value products.
Acknowledgements This research work presented in this chapter was supported by NSF
(#1911480) and UML-WPI seed grant (2019–2020). The authors would also like to thank Dr.
Carl Lawton and Massachusetts Biomanufacturing Center for providing experimental facilities and
technical support.
References
Abghari A, Chen S (2014) Yarrowia lipolytica as an oleaginous cell factory platform for production
of fatty acid-based biofuel and bioproducts. Front Energy Res 2. https://doi.org/10.3389/fenrg.
2014.00021
Alvarez HM (2016) Triacylglycerol and wax ester-accumulating machinery in prokaryotes.
Biochimie 120:28–39
Arsène F, Tomoyasu T, Bukau B (2000) The heat shock response of Escherichia coli. Int J Food
Microbiol 55(1–3):3–9
Bailey DA, Feldmann P, Bovey M, Gow N, Brown A (1996) The Candida albicans HYR1 gene,
which is activated in response to hyphal development, belongs to a gene family encoding yeast
cell wall proteins. J Bacteriol 178(18):5353–5360
Bankar A, Zinjarde S, Telmore A, Walke A, Kumar AR (2018) Morphological response of
Yarrowia lipolytica under stress of heavy metals. Can J Microbiol 64(8):559–566. https://doi.
org/10.1139/cjm-2018-0050
Barria C, Malecki M, Arraiano C (2013) Bacterial adaptation to cold. Microbiology
159(Pt_12):2437–2443
Bartholomew CR, Hardy CF (2009) p21-activated kinases Cla4 and Ste20 regulate vacuole inher-
itance in Saccharomyces cerevisiae. Eukaryot Cell 8(4):560–572
Beopoulos A, Chardot T, Nicaud JM (2009) Yarrowia lipolytica: A model and a tool to understand
the mechanisms implicated in lipid accumulation. Biochimie 91(6):692–696. https://doi.org/10.
1016/j.biochi.2009.02.004
Berman J, Sudbery PE (2002) Candida Albicans: a molecular revolution built on lessons from
budding yeast. Nat Rev Genet 3(12):918–930. https://doi.org/10.1038/nrg948
Biswas S, Van Dijck P, Datta A (2007) Environmental sensing and signal transduction path-
ways regulating morphopathogenic determinants of Candida albicans. Microbiol Mol Biol Rev
71(2):348–376
Blazeck J, Hill A, Liu L, Knight R, Miller J, Pan A, Alper HS (2014) Harnessing Yarrowia lipoly-
tica lipogenesis to create a platform for lipid and biofuel production. Nat Commun 5:3131.
https://doi.org/10.1038/ncomms4131
Braga A, Gomes N, Belo I (2012) Lipase induction in Yarrowia lipolytica for castor oil hydrolysis
and its effect on γ-decalactone production. J Am Oil Chem Soc 89(6):1041–1047. https://doi.
org/10.1007/s11746-011-1987-5
Cao C, Wu M, Bing J, Tao L, Ding X, Liu X, Huang G (2017) Global regulatory roles of the
cAMP/PKA pathway revealed by phenotypic, transcriptomic and phosphoproteomic analyses
in a null mutant of the PKA catalytic subunit in Candida albicans. Mol Microbiol 105(1):46–64
Cervantes-Chávez JA, Ruiz-Herrera J (2006) STE11 disruption reveals the central role of a MAPK
pathway in dimorphism and mating in Yarrowia lipolytica. FEMS Yeast Res 6(5):801–815
Cervantes-Chávez JA, Ruiz-Herrera J (2007) The regulatory subunit of protein kinase a promotes
hyphal growth and plays an essential role in Yarrowia lipolytica. FEMS Yeast Res 7(6):929–940
Cervantes-Chávez JA, Kronberg F, Passeron S, Ruiz-Herrera J (2009) Regulatory role of the PKA
pathway in dimorphism and mating in Yarrowia lipolytica. Fungal Genet Biol 46(5):390–399
Chaleff DT, Tatchell K (1985) Molecular cloning and characterization of the STE7 and STE11
genes of Saccharomyces cerevisiae. Mol Cell Biol 5(8):1878–1886
84 N. Liu et al.
Chen B, Ling H, Chang MW (2013) Transporter engineering for improved tolerance against alkane
biofuels in Saccharomyces cerevisiae. Biotechnol Biofuels 6(1):21–21. https://doi.org/10.1186/
1754-6834-6-21
Chen Y, Siewers V, Nielsen J (2012) Profiling of cytosolic and peroxisomal acetyl-CoA
metabolism in Saccharomyces cerevisiae. PLoS One 7(8):e42475. https://doi.org/10.1371/jou
rnal.pone.0042475
Cordova LT, Butler J, Alper HS (2020) Direct production of fatty alcohols from glucose using
engineered strains of Yarrowia lipolytica. Metabol Eng Commun 10:e00105
Costa M, Borges CL, Bailao AM, Meirelles GV, Mendonça YA, Dantas SF, Mendes-Giannini MJ
(2007) Transcriptome profiling of Paracoccidioides brasiliensis yeast-phase cells recovered
from infected mice brings new insights into fungal response upon host interaction. Microbi-
ology 153(12):4194–4207
Cuevas B, Abell A, Johnson G (2007) Role of mitogen-activated protein kinase kinase kinases in
signal integration. Oncogene 26(22):3159–3171
Darbani B, Stovicek V, van der Hoek SA, Borodina I (2019) Engineering energetically effi-
cient transport of dicarboxylic acids in yeast Saccharomyces cerevisiae. Proc Natl Acad Sci
116(39):19415. https://doi.org/10.1073/pnas.1900287116
Darvishi F, Fathi Z, Ariana M, Moradi H (2017) Yarrowia lipolytica as a workhorse for biofuel
production. Biochem Eng J 127:87–96
Deive FJ, Sanromán MA, Longo MA (2010) A comprehensive study of lipase production by
Yarrowia lipolytica CECT 1240 (ATCC 18942): from shake flask to continuous bioreactor. J
Chem Technol Biotechnol 85(2):258–266. https://doi.org/10.1002/jctb.2301
DiRusso CC, Black PN (1999) Long-chain fatty acid transport in bacteria and yeast. Paradigms
for defining the mechanism underlying this protein-mediated process. Mol Cell Biochem
192(1):41–52. https://doi.org/10.1023/A:1006823831984
Doan CD, To CM, De Vrieze M, Lynen F, Danthine S, Brown A, Patel AR (2017) Chemical pro-
filing of the major components in natural waxes to elucidate their role in liquid oil structuring.
Food Chem 214:717–725
Dobrowolski A, Mituła P, Rymowicz W, Mirończuk AM (2016) Efficient conversion of crude glyc-
erol from various industrial wastes into single cell oil by yeast Yarrowia lipolytica. Biores
Technol 207:237–243
Dominguez A, Ferminan E, Gaillardin C (2000) Yarrowia lipolytica: an organism amenable to
genetic manipulation as a model for analyzing dimorphism in fungi. Contrib Microbiol 5:151–
172. Retrieved from https://www.ncbi.nlm.nih.gov/pubmed/10863671
D’Souza CA, Heitman J (2001) Conserved cAMP signaling cascades regulate fungal development
and virulence. FEMS Microbiol Rev 25(3):349–364
Dulermo R, Gamboa-Melendez H, Dulermo T, Thevenieau F, Nicaud JM (2014) The fatty acid
transport protein Fat1p is involved in the export of fatty acids from lipid bodies in Yarrowia
lipolytica. FEMS Yeast Res 14(6):883–896. https://doi.org/10.1111/1567-1364.12177
Dulermo R, Gamboa-Meléndez H, Ledesma-Amaro R, Thévenieau F, Nicaud JM (2015) Unrav-
eling fatty acid transport and activation mechanisms in Yarrowia lipolytica. Biochim Biophys
Acta 1851(9):1202–1217. https://doi.org/10.1016/j.bbalip.2015.04.004
Enshaeieh M, Nahvi I, Madani M (2014) Improving microbial oil production with standard and
native oleaginous yeasts by using Taguchi design. Int J Environ Sci Technol 11(3):597–604
Farwick A, Bruder S, Schadeweg V, Oreb M, Boles E (2014) Engineering of yeast hexose trans-
porters to transport D-xylose without inhibition by D-glucose. Proc Natl Acad Sci USA
111(14):5159–5164. https://doi.org/10.1073/pnas.1323464111
Ferreira R, Teixeira PG, Gossing M, David F, Siewers V, Nielsen J (2018) Metabolic engineering of
Saccharomyces cerevisiae for overproduction of triacylglycerols. Metabol Eng Commun 6:22–
27. https://doi.org/10.1016/j.meteno.2018.01.002
Fickers P, Nicaud JM, Destain J, Thonart P (2003) Overproduction of lipase by Yarrowia lipoly-
tica mutants. Appl Microbiol Biotechnol 63(2):136–142. https://doi.org/10.1007/s00253-003-
1342-3
Cellular Engineering of Yarrowia lipolytica … 85
Fickers P, Benetti PH, Wache Y, Marty A, Mauersberger S, Smit MS, Nicaud JM (2005a)
Hydrophobic substrate utilisation by the yeast Yarrowia lipolytica, and its potential applica-
tions. FEMS Yeast Res 5(6–7):527–543. https://doi.org/10.1016/j.femsyr.2004.09.004
Fickers P, Fudalej F, Nicaud JM, Destain J, Thonart P (2005b) Selection of new over-producing
derivatives for the improvement of extracellular lipase production by the non-conventional
yeast Yarrowia lipolytica. J Biotechnol 115(4):379–386. https://doi.org/10.1016/j.jbiotec.2004.
09.014
Fiume MM, Heldreth BA, Bergfeld WF, Belsito DV, Hill RA, Klaassen CD, Slaga TJ (2015) Safety
Assessment of alkyl esters as used in cosmetics. Int J Toxicol 34(2_suppl):5S-69S
Gao Q, Cao X, Huang Y-Y, Yang J-L, Chen J, Wei LJ, Hua Q (2018) Overproduction of fatty
acid ethyl esters by the oleaginous yeast Yarrowia lipolytica through metabolic engineering and
process optimization. ACS Synth Biol 7(5):1371–1380
Goncalves FA, Colen G, Takahashi JA (2014) Yarrowia lipolytica and its multiple applications in
the biotechnological industry. Sci World J 476207. https://doi.org/10.1155/2014/476207
González-López CI, Ortiz-Castellanos L, Ruiz-Herrera J (2006) The ambient pH response Rim
pathway in Yarrowia lipolytica: identification of YlRIM9 and characterization of its role in
dimorphism. Curr Microbiol 53(1):8–12
Guan N, Li J, Shin H-D, Du G, Chen J, Liu L (2017) Microbial response to environmental
stresses: from fundamental mechanisms to practical applications. Appl Microbiol Biotechnol
101(10):3991–4008
Guevara-Olvera L, Calvo-Mendez C, Ruiz-Herrera J (1993) The role of polyamine metabolism in
dimorphism of Yarrowia lipolytica. Microbiology 139(3):485–493
Hara KY, Kobayashi J, Yamada R, Sasaki D, Kuriya Y, Hirono-Hara Y, Kondo A (2017) Trans-
porter engineering in biomass utilization by yeast. FEMS Yeast Res 17(7).https://doi.org/10.
1093/femsyr/fox061
Hector RE, Qureshi N, Hughes SR, Cotta MA (2008) Expression of a heterologous xylose trans-
porter in a Saccharomyces cerevisiae strain engineered to utilize xylose improves aerobic
xylose consumption. Appl Microbiol Biotechnol 80(4):675–684. https://doi.org/10.1007/s00
253-008-1583-2
Hogan DA, Sundstrom P (2009) The Ras/cAMP/PKA signaling pathway and virulence in Candida
albicans. Future Microbiol 4(10):1263–1270
Hu Y, Zhu Z, Nielsen J, Siewers V (2018) Heterologous transporter expression for improved fatty
alcohol secretion in yeast. Metab Eng 45:51–58. https://doi.org/10.1016/j.ymben.2017.11.008
Huo K, Zhao F, Zhang F, Liu R, Yang C (2020) Morphology engineering: a new strategy to
construct microbial cell factories. World J Microbiol Biotechnol 36(9):1–15
Hurtado CA, Rachubinski RA (2002) YlBMH1 encodes a 14-3-3 protein that promotes filamentous
growth in the dimorphic yeast Yarrowia lipolytica. Microbiology 148(11):3725–3735
Hurtado CA, Beckerich J-M, Gaillardin C, Rachubinski RA (2000) A rac homolog is required for
induction of hyphal growth in the dimorphic yeast Yarrowia lipolytica. J Bacteriol 182(9):2376–
2386
Hutmacher DW, Singh H (2008) Computational fluid dynamics for improved bioreactor design and
3D culture. Trends Biotechnol 26(4):166–172. https://doi.org/10.1016/j.tibtech.2007.11.012
Ishige T, Tani A, Takabe K, Kawasaki K, Sakai Y, Kato N (2002) Wax ester production from n-
alkanes by Acinetobacter sp. strain M-1: ultrastructure of cellular inclusions and role of acyl
coenzyme a reductase. Appl Environ Microbiol 68(3):1192–1195
Jetter R, Kunst L (2008) Plant surface lipid biosynthetic pathways and their utility for metabolic
engineering of waxes and hydrocarbon biofuels. Plant J 54(4):670–683
Jiang X-R, Chen G-Q (2016) Morphology engineering of bacteria for bio-production. Biotechnol
Adv 34(4):435–440
Jiménez-Bremont JF, Rodriguez-Hernandez AA, Rodriguez-Kessler M, Ruiz-Herrera J (2012)
Development and dimorphism of the yeast Yarrowia lipolytica. In: ER-H. José, Dimorphic
fungi. their importance as models for differentiation and fungal pathogenesis, pp 58–66
Johnson DA, Akamine P, Radzio-Andzelm E, Madhusudan A, Taylor SS (2001) Dynamics of
cAMP-dependent protein kinase. Chem Rev 101(8):2243–2270
86 N. Liu et al.
Liu N, Soong Y-HV, Mirzaee I et al (2021) Biomanufacturing of value-added products from oils or
fats: a case study on cellular and fermentation engineering of Yarrowia lipolytica. Biotechnol
Bioeng 1–16
Luyten K, Riou C, Blondin B (2002) The hexose transporters of Saccharomyces cerevisiae play
different roles during enological fermentation. 19(8):713–726
Madzak C (2018) Engineering Yarrowia lipolytica for use in biotechnological applications: a
review of major achievements and recent innovations. Mol Biotechnol 60(8):621–635
Magdouli S, Guedri T, Rouissi T, Brar SK, Blais J-F (2020) Sync between leucine, biotin and
citric acid to improve lipid production by Yarrowia lipolytica on crude glycerol-based media.
Biomass Bioenergy 142:105764
Marchler G, Schüller C, Adam G, Ruis H (1993) A saccharomyces cerevisiae UAS element con-
trolled by protein kinase a activates transcription in response to a variety of stress conditions.
EMBO J 12(5):1997–2003
Martínez-Soto D, Ruiz-Herrera J (2017) Functional analysis of the MAPK pathways in fungi. Rev
Iberoamericana De Mycologia 34(4):192–202
Martinez-Vazquez A, Gonzalez-Hernandez A, Domínguez Á, Rachubinski R, Riquelme M,
Cuellar-Mata P, Guzman JCT (2013) Identification of the transcription factor Znc1p, which
regulates the yeast-to-hypha transition in the dimorphic yeast Yarrowia lipolytica. PLoS One
8(6):e66790
Mcdaniel R, Behrouzian B, Clark L, Hattendorf D, Valle F (2011) Production of fatty alcohols with
fatty alcohol forming acyl-coa reductases (far). In: Google patents
Miwa TK (1971) Jojoba oil wax esters and derived fatty acids and alcohols: gas chromatographic
analyses. J Am Oil Chem Soc 48(6):259–264
Mlickova K, Roux E, Athenstaedt K, d’Andrea S, Daum G, Chardot T, Nicaud JM (2004) Lipid
accumulation, lipid body formation, and acyl coenzyme a oxidases of the yeast Yarrowia lipoly-
tica. Appl Environ Microbiol 70(7):3918–3924. https://doi.org/10.1128/AEM.70.7.3918-3924.
2004
Morales-Vargas AT, Dominguez A, Ruiz-Herrera J (2012) Identification of dimorphism-involved
genes of Yarrowia lipolytica by means of microarray analysis. Res Microbiol 163(5):378–387.
https://doi.org/10.1016/j.resmic.2012.03.002
Najjar A, Robert S, Guerin C, Violet-Asther M, Carriere F (2011) Quantitative study of lipase
secretion, extracellular lipolysis, and lipid storage in the yeast Yarrowia lipolytica grown
in the presence of olive oil: analogies with lipolysis in humans. Appl Microbiol Biotechnol
89(6):1947–1962. https://doi.org/10.1007/s00253-010-2993-5
Okano T, Yamada N, Okuhara M, Sakai H, Sakurai Y (1995) Mechanism of cell detachment from
temperature-modulated, hydrophilic–hydrophobic polymer surfaces. In: The biomaterials: sil-
ver jubilee compendium, Elsevier, pp 109–115
Palecek SP, Parikh AS, Kron SJ (2002) Sensing, signalling and integrating physical processes dur-
ing Saccharomyces cerevisiae invasive and filamentous growth. Microbiology 148(Pt 4):893–
907. https://doi.org/10.1099/00221287-148-4-893
Papanikolaou S, Diamantopoulou P, Blanchard F, Lambrinea E, Chevalot I, Stoforos NG, Rondags
E (2020) Physiological characterization of a novel wild-type Yarrowia lipolytica strain grown
on glycerol: effects of cultivation conditions and mode on polyols and citric acid production.
Appl Sci 10(20):7373
Papanikolaou S, Chevalot I, Komaitis M, Marc I, Aggelis G (2002) Single cell oil production by
Yarrowia lipolytica growing on an industrial derivative of animal fat in batch cultures. Appl
Microbiol Biotechnol 58:308-312. lipid-derived metabolites
Papanikolaou S, Chevalot I, Galiotou-Panayotou M, Komaitis M, Marc I, Aggelis G (2007) Indus-
trial derivative of tallow: a promising renewable substrate for microbial lipid, single-cell protein
and lipase production by Yarrowia lipolytica. Electron J Biotechnol 10(3):425–435.https://doi.
org/10.2225/vol10-issue3-fulltext-8
Pérez-Campo FM, Domínguez A (2001) Factors affecting the morphogenetic switch in Yarrowia
lipolytica. Curr Microbiol 43(6):429–433
88 N. Liu et al.
Petersson AE, Gustafsson LM, Nordblad M, Börjesson P, Mattiasson B, Adlercreutz P (2005) Wax
esters produced by solvent-free energy-efficient enzymatic synthesis and their applicability as
wood coatings. Green Chem 7(12):837–843
Pomraning KR, Bredeweg EL, Kerkhoven EJ, Barry K, Haridas S, Hundley H, Baker SE (2018)
Regulation of yeast-to-hyphae transition in Yarrowia lipolytica. mSphere 3(6):e00541–00518.
https://doi.org/10.1128/mSphere.00541-18
Reifenberger E, Boles E, Ciriacy M (1997) Kinetic characterization of individual hexose trans-
porters of Saccharomyces cerevisiae and their relation to the triggering mechanisms of glucose
repression. Eur J Biochem 245(2):324–333. https://doi.org/10.1111/j.1432-1033.1997.00324.x
Rintala E, Wiebe MG, Tamminen A, Ruohonen L, Penttilä M (2008) Transcription of hexose trans-
porters of Saccharomyces cerevisiae affected by change in oxygen provision. BMC Microbiol
8(1):53. https://doi.org/10.1186/1471-2180-8-53
Röttig A, Strittmatter CS, Schauer J, Hiessl S, Poehlein A, Daniel R, Steinbüchel A (2016) Role of
wax ester synthase/acyl coenzyme a: diacylglycerol acyltransferase in oleaginous Streptomyces
sp. strain G25. Appl Environ Microbiol 82(19):5969–5981
Round JW, Roccor R, Eltis LD (2019) A biocatalyst for sustainable wax ester production: re-
wiring lipid accumulation in Rhodococcus to yield high-value oleochemicals. Green Chem
21(23):6468–6482
Ruiz-Herrera J, Sentandreu R (2002) Different effectors of dimorphism in Yarrowia lipolytica.
Arch Microbiol 178(6):477–483
Rupp S, Summers E, Lo HJ, Madhani H, Fink G (1999) MAP kinase and cAMP filamentation sig-
naling pathways converge on the unusually large promoter of the yeast FLO11 gene. EMBO J
18(5):1257–1269
Ryu S, Trinh CT (2018) Understanding functional roles of native pentose-specific transporters for
activating dormant pentose metabolism in Yarrowia lipolytica. Appl Environ Microbiol 84(3).
https://doi.org/10.1128/aem.02146-17
Sabirova JS, Haddouche R, Van Bogaert I, Mulaa F, Verstraete W, Timmis K, Soetaert W (2011)
The ‘LipoYeasts’ project: using the oleaginous yeast Yarrowia lipolytica in combination with
specific bacterial genes for the bioconversion of lipids, fats and oils into high-value products.
Microb Biotechnol 4(1):47–54
Santala S, Efimova E, Koskinen P, Karp MT, Santala V (2014) Rewiring the wax ester production
pathway of Acinetobacter baylyi ADP1. ACS Synth Biol 3(3):145–151
Sedlak M, Ho NW (2004) Characterization of the effectiveness of hexose transporters for transport-
ing xylose during glucose and xylose co-fermentation by a recombinant Saccharomyces yeast.
Yeast 21(8):671–684. https://doi.org/10.1002/yea.1060
Shani N, Valle D (1996) A Saccharomyces cerevisiae homolog of the human adrenoleukodystrophy
transporter is a heterodimer of two half ATP-binding cassette transporters. Proc Natl Acad Sci
USA 93(21):11901–11906. https://doi.org/10.1073/pnas.93.21.11901
Sirakova TD, Deb C, Daniel J, Singh HD, Maamar H, Dubey VS, Kolattukudy PE (2012) Wax
ester synthesis is required for Mycobacterium tuberculosis to enter in vitro dormancy. PLoS
One 7(12)
Soong YHV, Liu N, Yoon S, Lawton C, Xie D (2019) Cellular and metabolic engineering of oleagi-
nous yeast Yarrowia lipolytica for bioconversion of hydrophobic substrates into high-value
products. Eng Life Sci 19(6):423–443
Soong Y, Zhao L, Liu N, Yu P, Lopez C, Olson A, Wong H, Shao Z, Xie D (2021) Microbial
Synthesis of Wax Esters. Metab Eng 67:428–442. https://doi.org/10.1016/j.ymben.2021.08.002
Spagnuolo M, Shabbir Hussain M, Gambill L, Blenner M (2018) Alternative substrate metabolism
in Yarrowia lipolytica. Front Microbiol 9:1077. https://doi.org/10.3389/fmicb.2018.01077
Su C, Yu J, Lu Y (2018) Hyphal development in Candida albicans from different cell states. Curr
Genet 64(6):1239–1243
Szabo R (2001) Cla4 protein kinase is essential for filament formation and invasive growth of
Yarrowia lipolytica. Mol Genet Genomics 265(1):172–179
Cellular Engineering of Yarrowia lipolytica … 89
Szabo R, Štofanı́ková V (2002) Presence of organic sources of nitrogen is critical for filament
formation and pH-dependent morphogenesis in Yarrowia lipolytica. FEMS Microbiol Lett
206(1):45–50
Taylor S, Yang J, Wu J, Haste N, Radzio-Andzelm E, Anand G (2004) PKA: a portrait of protein
kinase dynamics. Biochim Biophys Acta (BBA)-Proteins Proteomics 1697(1–2):259–269
Timoumi A, Guillouet SE, Molina-Jouve C, Fillaudeau L, Gorret N (2018) Impacts of environmen-
tal conditions on product formation and morphology of Yarrowia lipolytica. Appl Microbiol
Biotechnol 102(9):3831–3848
Toda T, Cameron S, Sass P, Zoller M, Wigler M (1987) Three different genes in S. cerevisiae
encode the catalytic subunits of the cAMP-dependent protein kinase. Cell 50(2):277–287
Tomiyama T, Kurihara K, Ogawa T, Maruta T, Ogawa T, Ohta D, Ishikawa T (2017) Wax ester
synthase/diacylglycerol acyltransferase isoenzymes play a pivotal role in wax ester biosynthesis
in Euglena gracilis. Sci Rep 7(1):1–13
Torres-Guzman JC, Domínguez A (1997) HOY1, a homeo gene required for hyphal formation in
Yarrowia lipolytica. Mol Cell Biol 17(11):6283–6293
Treger JM, Magee TR, McEntee K (1998) Functional analysis of the stress response element and its
role in the multistress response of Saccharomyces cerevisiae. Biochem Biophys Res Commun
243(1):13–19
van der Walt JP, von Arx JA (1980) The yeast genus Yarrowia gen. nov. Antonie Van Leeuwenhoek
46(6):517–521. Retrieved from https://www.ncbi.nlm.nih.gov/pubmed/7195185
Wältermann M, Hinz A, Robenek H, Troyer D, Reichelt R, Malkus U, Von Landenberg P (2005)
Mechanism of lipid-body formation in prokaryotes: how bacteria fatten up. Mol Microbiol
55(3):750–763
Wang HJ, Le Dall MT, Wach Y, Laroche C, Belin JM, Gaillardin C, Nicaud JM (1999) Evalua-
tion of acyl coenzyme A oxidase (Aox) isozyme function in the n-alkane-assimilating yeast
Yarrowia lipolytica. J Bacteriol 181(17):5140–5148. https://doi.org/10.1128/jb.181.17.5140-
5148.1999
Wang H, Jia X, Wang X, Zhou Z, Wen J, Zhang J (2014) CFD modeling of hydrodynamic charac-
teristics of a gas–liquid two-phase stirred tank. Appl Math Model 38(1):63–92. https://doi.org/
10.1016/j.apm.2013.05.032
Wang G, Li D, Miao Z, Zhang S, Liang W, Liu L (2018) Comparative transcriptome analysis
reveals multiple functions for Mhy1p in lipid biosynthesis in the oleaginous yeast Yarrowia
lipolytica. Biochim Biophys Acta (BBA)-Mol Cell Biol Lipids 1863(1):81–90
Weeks G, Spiegelman GB (2003) Roles played by Ras subfamily proteins in the cell and develop-
mental biology of microorganisms. Cell Signal 15(10):901–909
Wieczorke R, Krampe S, Weierstall T, Freidel K, Hollenberg CP, Boles E (1999) Concurrent
knock-out of at least 20 transporter genes is required to block uptake of hexoses in Saccha-
romyces cerevisiae. FEBS Lett 464(3):123–128. https://doi.org/10.1016/S0014-5793(99)016
98-1
Willis RM, Wahlen BD, Seefeldt LC, Barney BM (2011) Characterization of a fatty acyl-CoA
reductase from Marinobacter aquaeolei VT8: a bacterial enzyme catalyzing the reduction of
fatty acyl-CoA to fatty alcohol. Biochemistry 50(48):10550–10558
Wu H, Shu T, Mao Y-S, Gao XD (2020) Characterization of the promoter, downstream target genes
and recognition DNA sequence of Mhy1, a key filamentation-promoting transcription factor in
the dimorphic yeast Yarrowia lipolytica. Curr Genet 66(1):245–261
Xie D (2017) Integrating cellular and bioprocess engineering in the non-conventional yeast
Yarrowia lipolytica for biodiesel production: a review. Front Bioeng Biotechnol 5:65
Xu P, Qiao K, Ahn WS, Stephanopoulos G (2016) Engineering Yarrowia lipolytica as a plat-
form for synthesis of drop-in transportation fuels and oleochemicals. Proc Natl Acad Sci
113(39):10848–10853
Xu C, Liu R, Zhang Q, Chen X, Qian Y, Fang W (2017) The diversification of evolutionarily con-
served MAPK cascades correlates with the evolution of fungal species and development of
lifestyles. Genome Biol Evol 9(2):311–322
90 N. Liu et al.
Xue Z, Sharpe PL, Hong S-P, Yadav NS, Xie D, Short DR, Wang J (2013) Production of omega-
3 eicosapentaenoic acid by metabolic engineering of Yarrowia lipolytica. Nat Biotechnol
31(8):734–740
Yang K, Qiao Y, Li F, Xu Y, Yan Y, Madzak C, Yan J (2019) Subcellular engineering of lipase
dependent pathways directed towards lipid related organelles for highly effectively compart-
mentalized biosynthesis of triacylglycerol derived products in Yarrowia lipolytica. Metab Eng
55:231–238
Young EM, Tong A, Bui H, Spofford C, Alper HS (2014) Rewiring yeast sugar transporter prefer-
ence through modifying a conserved protein motif. Proc Natl Acad Sci 111(1):131. https://doi.
org/10.1073/pnas.1311970111
Yuzbashev TV, Yuzbasheva EY, Sobolevskaya TI, Laptev IA, Vybornaya TV, Larina AS, Sineoky
SP (2010) Production of succinic acid at low pH by a recombinant strain of the aerobic yeast
Yarrowia lipolytica. Biotechnol Bioeng 107(4):673–682
Zakhartsev M, Reuss M (2018) Cell size and morphological properties of yeast Saccharomyces
cerevisiae in relation to growth temperature. FEMS yeast Res 18(6):foy052
Zeng SY, Liu HH, Shi TQ, Song P, Ren LJ, Huang H, Ji XJ (2018) Recent advances in
metabolic engineering of Yarrowia lipolytica for lipid overproduction. Eur J Lipid Sci Technol
120(3):1700352
Zhang W, Liu HT (2002) MAPK signal pathways in the regulation of cell proliferation in mam-
malian cells. Cell Res 12(1):9–18
Zhang J-L, Cao Y-X, Peng Y-Z, Jin C-C, Bai Q-Y, Zhang R-S, Yuan Y-J (2019) High produc-
tion of fatty alcohols in Yarrowia lipolytica by coordination with glycolysis. Sci China Chem
62(8):1007–1016
Zinjarde SS, Kale BV, Vishwasrao PV, Kumar AR (2008) Morphogenetic behavior of tropical
marine yeast Yarrowia lipolytica in response to hydrophobic substrates. J Microbiol Biotechnol
18(9):1522e1528
Whole Cell Yeast-Based Biosensors
Abstract
Analyte detection is a major component of fieldwork, environmental surveil-
lance, and health protection, but lack of resources or access to instruments can
create challenges in technology-limited environments such as remote research
sites or low- and middle-income countries (LMICs). Whole-cell yeast-based
biosensors provide potential solutions to these barriers. Here, we discuss the
components of whole-cell yeast-based biosensors, emphasizing the process by
which they are designed for their analyte of interest, approaches for harnessing
them to function in particular research environments, and their advantages and
disadvantages relative to other analytical tools. We provide examples of the var-
ious purposes for which existing whole-cell yeast-based biosensors have been
used, focusing most on those appropriate for detecting externally-generated ana-
lytes in technology-limited settings. The further development of field-friendly
whole-cell yeast-based biosensors still faces challenges, including the need to
reduce the time from contact with the analyte to signal readout of the biosen-
sor. Regardless, the inexpensive, robust, portable, environmentally friendly,
and highly modular nature of yeast-based biosensors suggests that they could
become useful tools for a range of analytical tasks.
1 Introduction
Sensors are tools used for analyte detection, analyte quantification, and/or gather-
ing information relating to biological activity. One subset of sensors is biosensors.
Though this term has varied meanings, here we define a biosensor as an analytical
tool that utilizes biological components (e.g., proteins, nucleic acids, whole cells
or even animals) to detect the analyte(s) of interest (Ostrov et al. 2017). Whole-cell
biosensors are a subset of biosensors that utilize living cells for sensing tasks. One
advantage of whole-cell biosensors relative to many other types of sensors is that
whole-cell biosensors can provide information on the biological relevance of the
analyte. This information is particularly important in cases where the sensing task
is performed in a complex environment (e.g., wastewater) or where compounds of
interest have varied composition (such as with endocrine disruptors) or unknown
identity (e.g., as might occur with environmental mutagens). Additionally, whole-
cell biosensors have the potential to provide detection with both high sensitivity
and high specificity without need for expensive equipment or analytical-grade stan-
dards. These attributes suggest that whole-cell biosensor detection systems could
be a useful alternative to spectral and mass-based detection systems in areas that
lack sufficient equipment or resources (Martin-Yken 2020; Miller et al. 2020a, b).
Like most other sensors (biological and not), whole-cell biosensors are mod-
ular devices that can be viewed as being composed of standard components. As
typically described, the fundamental parts of a whole-cell biosensor are the recep-
tor (an element that interacts in a specific way with the analyte), the reporter (an
element that produces a detectable signal in response to the receptor-analyte inter-
action), and the transducer (an element or process that connects the state of the
receptor to the state of the reporter, sometimes with signal amplification) (Conroy
et al. 2009). However, in considering the full spectrum of whole-cell biosensors
and thinking about how they could be designed through synthetic biology, we
suggest that it is useful to view these systems as being composed of five com-
ponents: an analyte, a detection process (which often but not always involves
a specific receptor), a transduction process (often with signal amplification), a
response process (which often but not always involves production of an identifiable
reporter), and a readout process (which involves quantifying or otherwise detecting
the output of the response process) (Fig. 1, Table 1). In addition, most whole-cell
biosensors also require a sixth component: some type of external system housing
to contain the cells both during exposure to the analyte and periods of extended
storage (Fig. 1). The reason for breaking the full process down in this way is that
it enables a better understanding of how different modules can be rearranged,
especially for cases that don’t fit neatly into the “receptor-transducer-reporter”
paradigm.
Whole Cell Yeast-Based Biosensors 93
System Housing
Fig. 1 Elements of whole-cell yeast-based biosensors. Whole cell biosensors are made up of five
key components: analyte, detection process, transduction process, response process, and readout
process. Here we have provided examples of each of these components featured in this review.
These five components are then contained within the system housing, which is selected based on
the environment in which the biosensor is intended to be operated. In technology-limited settings,
paper-based assays or tubes read by a portable reader or the human senses (eyes, nose) would be
ideal. In some settings, preloaded microtiter plates could work depending on the availability of
suitable plate-reading technology
While the list above seems straightforward, people who are familiar with this
or similar systems will realize that there are some additional steps implicit in
the description above, and it is important to recognize these additional complica-
tions in considering how to design or improve particular whole-cell biosensors.
For example, the transduction process for induction of transcription in response to
estrogen is not as simple as implied above: it appears to involve ligand-mediated
release from the estrogen receptor of the cellular component HSP90, with atten-
dant translocation to the nucleus, though aspects of this process remain uncertain
(Fuentes and Silveyra 2019). Moreover, there are implicit amplification steps in
the process above. These amplification steps are important because there is no
way that a human nose would be able to detect the signal if there were a 1:1
relationship between analyte (estrogen) and response (produced odorant).
The involvement of additional (and potentially unknown) eukaryote-specific
cellular components and/or processes during transduction (as illustrated by the
estrogen sensor above) is a challenging aspect of harnessing exogenous receptors
and is a major reason why eukaryotic whole-cell biosensors are preferable to anal-
ogous prokaryotic systems for many applications. An additional issue with using
prokaryotes for whole-cell biosensors is that they are often poorly suited to express
eukaryotic receptors including the aforementioned estrogen receptor because they
lack many protein folding factors, post-translational modification enzymes, and
internal organelles specific to eukaryotic cells (Zhang et al. 2000; Vieira Gomes
et al. 2018). Biosensors based on mammalian cells provide a potential solution to
these problems, but they are expensive to maintain and require specialized equip-
ment to do so, as they are sensitive to environmental conditions (Jarque et al.
2016a).
96 H. A. M. Shepherd et al.
yeast molecular biology, see (Gardner and Jaspersen 2014). For information about
recent advances in yeast genetic engineering that may be relevant to biosensor
engineering, see reviews including (Malcı et al. 2020; Schindler 2020) and the
other chapters in this book.
In theory, yeast can be harnessed to make a biosensor suitable for any analyte to
which yeast naturally respond (e.g., nutrients, chemical toxins, metals, mutagens,
temperature, generic stress) or anything for which an exogenous detection system
can be imported from another organism (e.g., medicines, human hormones, light).
In order to maintain a living sensor (and thus maintain the cell machinery), the
analyte of interest must be within a concentration range compatible with yeast via-
bility. For the purposes of detecting general hazards, one could use a sensor based
on cell death. Whole-cell yeast-based biosensors can be particularly appropriate
for situations where a biologically-active analyte (e.g., a drug or toxin) is present
in low concentrations under conditions where detection by typical analytical tech-
niques (e.g., mass spectrometry) is not feasible; they can also be particularly useful
when the goal is to detect a biological activity (e.g., mutagenicity) instead of a par-
ticular chemical, or when the analyte is of unknown or of mixed composition. Note
that these last two applications stretch the typical definition of “analyte”.
In trying to design a whole-cell yeast-based biosensor, a key consideration is
where the analyte is expected to be when it is detected: on the cell surface, in
the cytoplasm, or in some other cellular compartment? It is important to recog-
nize that some analytes can cross the membrane spontaneously (e.g., hydrophobic
molecules such as steroid hormones), while others cross by dedicated transporters
(for example, nutrients such as sugars), and still others become spontaneously con-
centrated in organelles such as the mitochondria (e.g., hydrophobic cationic dye
molecules). However, many remain outside the cell or at best sequestered in inter-
nal vesicles because they can’t cross membranes spontaneously. Considering the
expected localization of the analyte is essential for proper design of any whole-
cell biosensor because one needs to make sure that the necessary detection system
components are in the same compartment as the analyte.
In our discussion of detection systems, we provide some examples of specific
analytes for which whole-cell yeast-based biosensors have been developed. For a
more comprehensive list, see (Adeniran et al. 2015; Jarque et al. 2016a; Martin-
Yken 2020).
98 H. A. M. Shepherd et al.
as the ability to induce DNA damage (Paetkau et al. 1994; Billinton et al. 1998;
Burrill and Silver 2011; Lu et al. 2015; Bui et al. 2015), cause cell stress (Hollis
et al. 2000; Hung et al. 2018; Gong et al. 2020), or serve as nutrients (Shepherd
et al. 2021; Trentman et al. 2021).
Most whole-cell biosensors developed thus far detect analytes that can make their
way into the cytoplasm. Many of these utilize specific receptors that interface
with the transcription/translation system for the transduction part of the sensing
process. Thus, these receptors typically have an analyte binding domain and a
DNA binding domain; those that work as activators (not repressors) generally also
have a transcription activator domain. Generally speaking, binding of analytes to
these receptors changes their ability to bind DNA, thus altering the amount of
transcription of a reporter gene and ultimately the amount of reporter protein.
An example of an endogenous DNA-binding receptor used in biosensors is
provided by the copper-response system; the yeast CUP1 promoter is activated
in response to copper through the action of endogenous machinery including the
DNA-binding Ace1p protein (Dameron et al. 1991; Smith et al. 2017). One par-
ticularly well-characterized exogenous DNA-binding intracellular receptor is the
bacterial tetracycline receptor (TetR), which can either activate or repress tran-
scription in yeast in response to binding tetracycline; the specific effect depends
on the details of how the system is set up (Baron and Bujard 2000). TetR is a
very “portable” type of gene response element (i.e., it can work in diverse organ-
isms) because it needs nothing other than itself and a target sequence placed near
a promoter to cause production of a reporter in response to TetR.
Another example of an imported DNA-binding receptor is provided by the
human estrogen receptor (ER), mentioned earlier. However, while the DNA bind-
ing ability of Ace1 and TetR are directly regulated by binding to their ligands
(copper and tetracycline), the effect of estrogen on the DNA-binding activity of
ER is more indirect and involves the participation of some additional cellular fac-
tors. Ultimately, binding of estrogen to ER causes increased transcription from
promoters controlled by “estrogen-response elements” (EREs), similar to what
happens with TetR (Gruber et al. 2004). However, the change in transcription in
response to estrogen cannot be explained simply by differential binding to EREs
in the presence of estrogen. Instead, it appears to involve regulation of ER nuclear
localization through differential binding to chaperone HSP90 and also involves
the participation of co-receptors (Smith and Toft 1993). When such additional
elements are involved in the downstream transduction process, the detection ele-
ment is much less portable, and use of compatible eukaryotic systems that contain
these downstream components (either naturally or by engineering) can be essen-
tial. Indeed, the compatibility of yeast with the estrogen receptor (as well as other
steroid receptors, see (Ito-Harashima et al. 2020)) is an advantage of yeast over
bacteria for whole-cell biosensor applications (Routledge and Sumpter 1996).
100 H. A. M. Shepherd et al.
Using a related set of ideas, researchers have developed yeast-based assays for
stress-inducing agents. Though these agents can overlap chemically and biologi-
cally with toxins, the transduction process in these assays occurs through induction
of cellular stress-response pathways. A typical response for these stress-detecting
sensors would be an accumulation of reporters placed under control of stress
response promoters (Hartner and Glieder 2006; Adeniran et al. 2015; Zhao et al.
2019). Similarly, there are yeast-based sensors for mutagens in which the yeast
genome is the detector, the transduction process is provided by the machinery
that senses and responds to DNA damage, and the response is increased tran-
scription/translation of reporters placed under control of DNA-damage-induced
promoters (Burrill and Silver 2011). An even more sensitive example of a muta-
gen detector is provided by a biosensor where the detector is a yeast CAN1 gene.
Mutations in CAN1 make yeast resistant to the toxin canavanine; the response to
mutagen is thus an increase in the number of yeast cells that are resistant to cana-
vanine, and the readout process is to count the number of canavanine-resistant
colonies (Ong et al. 2021).
Fig. 2 Examples of biosensor pathways for extracellular analytes, where endogenous signaling
pathways are typically used as part of the transduction process. The response can involve activa-
tion of transcription of a reporter gene or activation of other enzymes. a Analytes can either diffuse
across the membrane (red) or remain outside and bind to extracellular receptors (orange). Ligands
that can freely diffuse across the membrane typically bind to a receptor protein in the cytoplasm,
enabling the complex to enter the nucleus and alter transcription. For analytes that remain extra-
cellular, potential detection mechanisms include kinase receptors and second-messenger systems;
these bind ligands (orange) and then can either activate transcription, alter the activity of other
enzymes, or both. b Specific example of a medium-chain fatty acid biosensor detected through a
GPCR. A heterologous GPCR (yellow) detects a medium-chain fatty acid in the culture medium,
transmits this chemical signal to the yeast mating-pathway (mustard), which relays it to a synthetic
transcription factor (STF, orange). The STF activates transcription of GFP. The image for b was
reprinted with permission of the American Chemical Society (ACS). The original file can be found
at https://doi.org/10.1021/sb500365m. Further permissions regarding this image should be directed
to the ACS
receptor across the membrane to the intracellular part. The typical answer is that
the information crosses the membrane in the form of a conformation change. While
that is a reasonable answer for GPCRs (which cross the membrane seven times,
providing a 3-D structure by which to convey allosteric change), the mechanism
of this information transfer is less clear for proteins like receptor kinases, which
generally cross the membrane only one time (Fig. 2). One plausible explanation is
that binding of the receptor to analyte induces dimerization, and it is the dimeriza-
tion event that turns on downstream signalling (Nakamura et al. 2016). Thus, this
dimerization response could be considered the first step in the transduction pro-
cess. This point suggests that regulation of the level of expression of the receptor
itself can be very important for functionality of a sensing system: if the level is
too low or too high, performance of a system that relies on regulated dimerization
could be severely compromised.
5.3 Amplification
For example, one response system commonly used in conjunction with aptamers
is FRET (Fluorescence/Förster Resonance Energy Transfer). With FRET-based
aptamer-sensing molecules, the probes are placed such that FRET increases (or
decreases) upon binding to analyte (reviewed by (Yoon et al. 2012)). FRET
reporters can indeed be very effective. However, the ~ 1:1 ratio between detec-
tion (binding event) and response means that the concentration of analytes needs
to be relatively high for this method to be effective, or that the detection system
needs to be very sensitive. Thus, unless one can design a way to amplify the signal
from the aptamer (or similar molecule) and do so in a way that takes advantage
of the cellular environment, it might be more effective to stick with a non-cellular
context for using these tools to detect extracellular analytes if the goal is to do so
in a technology-limited environment.
lower protein expression levels and the longer replication times of yeast compared
to other potential whole-cell biosensors (e.g., E. coli), signals produced by these
color-based systems are relatively weak and can take a long time (e.g., >24 h) to
develop (Chen et al. 2018).
The reporters discussed thus far are light-based: they produce light or
color during the readout process. A more recently developed reporter for
transcription/translation-based transduction processes in yeast is an enzyme that
produces a scent: “scentsor” assays involve the induction of a reporter enzyme
that produces banana smell (Miller et al. 2020b). Aromatic compounds and their
properties have been paired with yeast for years through bioengineering and fer-
mentation (Van Wyk et al. 2018), making diverse olfactory reporters in whole-cell
yeast-based biosensors a realistic possibility. The benefit of these assays is that the
readout requires no equipment other than the human nose and minimal training
or resources outside of the biosensor itself. However, it is difficult to quantify a
scent, limiting scentsor-based assays to yes/no assessments. Despite these chal-
lenges, the ease of use suggests that olfactory-based reporter systems could be
ideal for technology-limited environments (Miller et al. 2020b).
108 H. A. M. Shepherd et al.
As noted above, cell growth and/or metabolism can be used as reporters for biosen-
sors used to detect nutrients (e.g., (Shepherd et al. 2021; Trentman et al. 2021)) or
toxins (Hung et al. 2018).
In addition, while the readouts of the reporters above follow logically from
the activities of the reporter proteins themselves, it is possible to have situations
where the responses, reporters, and readouts are less directly connected and/or
have additional layers. For example, Lehmann et al. (2000) engineered a yeast
strain in which a lactose-metabolizing enzyme (LacZ) was placed under control
of a copper-responsive promoter. Because yeast cannot otherwise metabolize lac-
tose, this created cells that require copper to grow on lactose. Thus, with this
strain, one can use growth on media lacking a carbon source other than lactose
to detect the presence of copper. In this case, LacZ is a reporter, and induction of
LacZ is one level of response, but cell growth resulting from the copper-induced
ability to metabolize lactose is another level of response. The overall readout for
detection of copper would be whatever method is convenient to assess cell growth
and/or metabolism; Lehmann and colleagues actually used amperometry to mea-
sure oxygen consumption. Using a similar strain in which LacZ was induced by
copper, Tag et al. (2007) measured changes in lactose concentration in response to
heavy metals. This lactose-based method works in combination with flow injection
analysis to allow for semi-continuous measurements and to increase the sensor’s
capability as an in-line sensor for heavy metal waste produced by manufacturers.
The text thus far focuses on the characteristics of the yeast cells used for whole-
cell biosensor work. To make a useful whole-cell biosensor-based device, one
generally needs to design a type of "housing" to hold the cells, expose them to
a potential analyte, and develop the signal. Also needed is a reading device to
interpret (recognize and/or quantify) the signal, though in some cases the senses
of the biosensor-users themselves (e.g., eyes to detect color, nose to detect smell)
can provide this functionality. An additional issue is that one also needs a way to
distribute the biosensor cells. Because this review focuses on whole-cell biosen-
sors for use in technology-limited settings, we concentrate here on reading and
housing devices that require minimal skills or access to specialized equipment for
use; such devices are generally based on large populations of cells. However, it is
important to recognize that yeast-based whole-cell biosensors have a long history
of use in laboratory settings (see e.g., (Routledge and Sumpter 1996)), and appli-
cations based on small populations (as in microtiter plates) or individual cells (as
in microfluidic devices) have been proliferating. We discuss some microtiter plate-
based approaches below, but for a deeper discussion of lab-based yeast whole-cell
biosensor-devices, see (Adeniran et al. 2015; Jarque et al. 2016a; Martin-Yken
2020).
Whole Cell Yeast-Based Biosensors 109
8 Long-Term Storage
Whole-cell yeast-based biosensors must be stored in a way that the biosensors can
be transported and used at a later time. When biosensor yeast are used in paper-
based assays, yeast cultures are typically stored using a sugar-based hydrogel to
protect the yeast until the biosensor is ready for use (Fine et al. 2006; Weaver
et al. 2015; Jarque et al. 2016b; Miller et al. 2020a). As noted above, bioPADs
constructed with yeast in hydrogel have been shown to be shelf-stable for > 1 year
at -20 °C, > six months at 4 °C, and at least 56 days at 37 °C (Weaver et al. 2015;
Miller et al. 2020a). Storage of yeast spotted onto paper-based "dipstick" biosen-
sors in argon has also been reported (Ostrov et al. 2017). For long-term storage,
drying methods such as lyophilization (freeze-drying) or storage in hydrogel have
been effective. Depending on when the biosensors will need to be used, freeze-
dried yeast can be stored for 2 months at 4 °C or 10 months at -18 °C and still
retain normal activity and sensitivity in assays (Jarque et al. 2016b). Immobiliza-
tion in polyvinyl alcohol hydrogel particles increased yeast storage time to 1 year
(Herkommerová et al. 2018). More high-tech treatments may also have utility in
enhancing long-term storage. For example, the application of different coatings
(e.g. polyelectrolytes, biomolecules, metal nanoparticles, or oxides) has been found
to improve reproducibility and cell sensitivity, viability, and stability as compared
Whole Cell Yeast-Based Biosensors 111
to bare cells (Bittner et al. 2015; Dai et al. 2018). In particular, bio-silica sol-gels
can act as protective shells against environmental factors. These gels have even
been found to protect cells from exposure to heavy metals and UV radiation for
up to 28 days (Ponamoreva et al. 2015).
9 Conclusions
References
Adachi T, Nakamura Y (2019) Aptamers: a review of their chemical properties and modifications
for therapeutic application. Molecules 23:4229. https://doi.org/10.3390/molecules24234229
Adeniran A, Sherer M, Tyo KEJ (2015) Yeast-based biosensors: design and applications. FEMS
Yeast Res 15:1–15. https://doi.org/10.1111/1567-1364.12203
Adeniran A, Stainbrook S, Bostick JW, Tyo KEJ (2018) Correction to “Detection of a peptide
biomarker by engineered yeast receptors.” ACS Synth Biol 7:1973–1973. https://doi.org/10.
1021/acssynbio.8b00291
Ahmad R, Dalziel JE (2020) G protein-coupled receptors in taste physiology and pharmacology.
Front Pharmacol 11:587664. https://doi.org/10.3389/fphar.2020.587664
Allard STM (2008) Bioluminescent reporter genes. https://www.promega.com/resources/pubhub/
enotes/bioluminescent-reporter-genes/
Aranda-Díaz A, Mace K, Zuleta I et al (2017) Robust synthetic circuits for two-dimensional control
of gene expression in yeast. ACS Synth Biol 6:545–554. https://doi.org/10.1021/acssynbio.6b0
0251
112 H. A. M. Shepherd et al.
Ballard ZS, Joung H-A, Goncharov A et al (2020) Deep learning-enabled point-of-care sensing
using multiplexed paper-based sensors. NPJ Digital Medicine 3:1–8. https://doi.org/10.1038/
s41746-020-0274-y
Baron U, Bujard H (2000) Tet repressor-based system for regulated gene expression in eukaryotic
cells: principles and advances. Methods Enzymol 327:401–421. https://doi.org/10.1016/s0076-
6879(00)27292-3
Berg B, Cortazar B, Tseng D et al (2015) Cell phone-based hand-held microplate reader for point-
of-care testing of enzyme-linked immunosorbent assays. ACS Nano 9:7857–7866. https://doi.
org/10.1021/acsnano.5b03203
Billinton N, Barker MG, Michel CE et al (1998) Development of a green fluorescent protein
reporter for a yeast genotoxicity biosensor. Biosens Bioelectron 13:831–838. https://doi.org/
10.1016/s0956-5663(98)00049-9
Bittner M, Jarque S, Hilscherová K (2015) Polymer-immobilized ready-to-use recombinant yeast
assays for the detection of endocrine disruptive compounds. Chemosphere 132:56–62. https://
doi.org/10.1016/j.chemosphere.2015.02.063
Borse V, Patil AS, Srivastava R (2017) Development and testing of portable fluorescence reader
(PorFloRTM ). In: 2017 9th international conference on communication systems and networks
(COMSNETS). pp 498–501
Bui VN, Nguyen TTH, Bettarel Y et al (2015) Genotoxicity of chemical compounds identification
and assessment by yeast cells transformed with GFP reporter constructs regulated by the PLM2
or DIN7 promoter. Int J Toxicol 34:31–43. https://doi.org/10.1177/1091581814566870
Burrill DR, Silver PA (2011) Synthetic circuit identifies subpopulations with sustained memory of
DNA damage. Genes Dev 25:434–439. https://doi.org/10.1101/gad.1994911
Cevenini L, Lopreside A, Calabretta MM et al (2018) A novel bioluminescent NanoLuc yeast-
estrogen screen biosensor (nanoYES) with a compact wireless camera for effect-based detec-
tion of endocrine-disrupting chemicals. Anal Bioanal Chem 410:1237–1246. https://doi.org/10.
1007/s00216-017-0661-7
Chen B, Lee HL, Heng YC et al (2018) Synthetic biology toolkits and applications in Saccha-
romyces cerevisiae. Biotechnol Adv 36:1870–1881. https://doi.org/10.1016/j.biotechadv.2018.
07.005
Conroy PJ, Hearty S, Leonard P, O’Kennedy RJ (2009) Antibody production, design and use for
biosensor-based applications. Semin Cell Dev Biol 20:10–26. https://doi.org/10.1016/j.semcdb.
2009.01.010
Dai B, Wang L, Wang Y et al (2018) Single-cell nanometric coating towards whole-cell-
based biodevices and biosensors. ChemistrySelect 3:7208–7221. https://doi.org/10.1002/slct.
201800963
Dameron CT, Winge DR, George GN et al (1991) A copper-thiolate polynuclear cluster in the
ACE1 transcription factor. PNAS 88:6127–6131. https://doi.org/10.1073/pnas.88.14.6127
Emr SD, Schauer I, Hansen W et al (1984) Invertase beta-galactosidase hybrid proteins fail to
be transported from the endoplasmic reticulum in Saccharomyces cerevisiae. Mol Cell Biol
4:2347–2355. https://doi.org/10.1128/MCB.4.11.2347
Fine T, Leskinen P, Isobe T et al (2006) Luminescent yeast cells entrapped in hydrogels for estro-
genic endocrine disrupting chemical biodetection. Biosens Bioelectron 21:2263–2269. https://
doi.org/10.1016/j.bios.2005.11.004
Fuentes N, Silveyra P (2019) Estrogen receptor signaling mechanisms. Adv Protein Chem Struct
Biol 116:135–170. https://doi.org/10.1016/bs.apcsb.2019.01.001
Gao CY, Pinkham JL (2000) Tightly regulated, β-estradiol dose-dependent expression system for
yeast. Biotechniques 29:1226–1231. https://doi.org/10.2144/00296st02
Gardner JM, Jaspersen SL (2014) Manipulating the yeast genome: deletion, mutation, and tagging
by PCR. Methods Mol Biol 1205:45–78. https://doi.org/10.1007/978-1-4939-1363-3_5
Gnügge R, Rudolf F (2017) Saccharomyces cerevisiae shuttle vectors. Yeast 34:205–221. https://
doi.org/10.1002/yea.3228
Whole Cell Yeast-Based Biosensors 113
Malcı K, Walls LE, Rios-Solis L (2020) Multiplex genome engineering methods for yeast cell fac-
tory development. Front Bioeng Biotechnol 8:1264. https://doi.org/10.3389/fbioe.2020.589468
Martin-Yken H (2020) Yeast-based biosensors: current applications and new developments.
Biosensors 10:51. https://doi.org/10.3390/bios10050051
Miller RA, Brown G, Barron E et al (2020a) Development of a paper-immobilized yeast biosensor
for the detection of physiological concentrations of doxycycline in technology-limited settings.
Anal Methods 12:2123–2132. https://doi.org/10.1039/D0AY00001A
Miller RA, Lee S, Fridmanski EJ et al (2020b) “Scentsor”: a whole-cell yeast biosensor with an
olfactory reporter for low-cost and equipment-free detection of pharmaceuticals. ACS Sensors
5:3025–3030. https://doi.org/10.1021/acssensors.0c01344
Möckli N, Auerbach D (2004) Quantitative β-galactosidase assay suitable for high-throughput
applications in the yeast two-hybrid system. Biotechniques 36:872–876. https://doi.org/10.
2144/04365PT03
Mukherjee K, Bhattacharyya S, Peralta-Yahya P (2015) GPCR-based chemical biosensors for
medium-chain fatty acids. ACS Synth Biol 4:1261–1269. https://doi.org/10.1021/sb500365m
Nakamura Y, Hashimoto T, Ishii J, Kondo A (2016) Dual-color reporter switching system to
discern dimer formations of G-protein-coupled receptors using Cre/loxP site-specific recom-
bination in yeast. Biotechnol Bioeng 113:2178–2190. https://doi.org/10.1002/bit.25974
Nelson DL, Cox MM (2017) Lehninger principles of biochemistry. In: Freeman WH, 7th edn. New
York, NY
Ong JY, Pence JT, Molik DC et al (2021) Yeast grown in continuous culture systems can detect
mutagens with improved sensitivity relative to the Ames test. PLoS ONE 16:e0235303. https://
doi.org/10.1371/journal.pone.0235303
Ostrov N, Jimenez M, Billerbeck S et al (2017) A modular yeast biosensor for low-cost point-of-
care pathogen detection. Sci Adv 3:e1603221. https://doi.org/10.1126/sciadv.1603221
Paetkau DW, Riese JA, MacMorran WS et al (1994) Interaction of the yeast RAD7 and SIR3 pro-
teins: implications for DNA repair and chromatin structure. Genes Dev 8:2035–2045. https://
doi.org/10.1101/gad.8.17.2035
Peltomaa R, Benito-Peña E, Barderas R, Moreno-Bondi MC (2019) Phage display in the quest
for new selective recognition elements for biosensors. ACS Omega 4:11569–11580. https://doi.
org/10.1021/acsomega.9b01206
Ponamoreva ON, Kamanina OA, Alferov VA et al (2015) Yeast-based self-organized hybrid bio-
silica sol–gels for the design of biosensors. Biosens Bioelectron 67:321–326. https://doi.org/
10.1016/j.bios.2014.08.045
Redden H, Morse N, Alper HS (2015) The synthetic biology toolbox for tuning gene expression
in yeast. FEMS Yeast Res 15:1–10. https://doi.org/10.1111/1567-1364.12188
Renneberg T, Kwan RCH, Chan C et al (2004) A salt-tolerant yeast-based microbial sensor for
24 hour community wastewater monitoring in coastal regions. Microchim Acta 148:235–240.
https://doi.org/10.1007/s00604-004-0266-7
Roney IJ, Rudner AD, Couture J-F, Kærn M (2016) Improvement of the reverse tetracycline
transactivator by single amino acid substitutions that reduce leaky target gene expression to
undetectable levels. Sci Rep 6:27697. https://doi.org/10.1038/srep27697
Routledge EJ, Sumpter JP (1996) Estrogenic activity of surfactants and some of their degrada-
tion products assessed using a recombinant yeast screen. Environ Toxicol Chem 15:241–248.
https://doi.org/10.1002/etc.5620150303
Routledge SJ, Mikaliunaite L, Patel A et al (2016) The synthesis of recombinant membrane pro-
teins in yeast for structural studies. Methods 95:26–37. https://doi.org/10.1016/j.ymeth.2015.
09.027
Schindler D (2020) Genetic engineering and synthetic genomics in yeast to understand life and
boost biotechnology. Bioengineering 7:137. https://doi.org/10.3390/bioengineering7040137
Seo KS, Choo KH, Chang HN, Park JK (2009) A flow injection analysis system with encapsulated
high-density Saccharomyces cerevisiae cells for rapid determination of biochemical oxygen
demand. Appl Microbiol Biotechnol 83:217–223. https://doi.org/10.1007/s00253-008-1852-0
Whole Cell Yeast-Based Biosensors 115
Shaw WM, Yamauchi H, Mead J et al (2019) Engineering a model cell for rational tuning of GPCR
signaling. Cell 177:782-796.e27. https://doi.org/10.1016/j.cell.2019.02.023
Shepherd HAM, Trentman MT, Tank JL, et al (2021) Development of a yeast-based assay for
bioavailable phosphorous. bioRxiv 28 Feb 2021. 433264. https://doi.org/10.1101/2021.02.28.
433264
Smith DF, Toft DO (1993) Steroid receptors and their associated proteins. Mol Endocrinol 7:4–11.
https://doi.org/10.1210/mend.7.1.8446107
Smith AD, Logeman BL, Thiele DJ (2017) Copper acquisition and utilization in fungi. Annu Rev
Microbiol 71:597–623. https://doi.org/10.1146/annurev-micro-030117-020444
Stainbrook SC, Yu JS, Reddick MP et al (2017) Modulating and evaluating receptor promiscuity
through directed evolution and modeling. Protein Eng Des Sel 30:455–465. https://doi.org/10.
1093/protein/gzx018
Tamai KT, Liu X, Silar P et al (1994) Heat shock transcription factor activates yeast metal-
lothionein gene expression in response to heat and glucose starvation via distinct signalling
pathways. Mol Cell Biol 14:8155–8165
Tag K, Riedel K, Bauer H-J, Hanke G, Baronian KHR, Kunze G (2007) Amperometric detec-
tion of Cu2+ by yeast biosensors using flow injection analysis (FIA). Sensors and Actuators
B: Chemical 122(2):403–409. https://doi.org/10.1016/j.snb.2006.06.007
Trentman MT, Tank JL, Shepherd HAM, et al (2021) Characterizing bioavailable phosphorus
concentrations in an agricultural stream during hydrologic and streambed disturbances. Bio-
geochemistry. https://doi.org/10.1007/s10533-021-00803-w
Urlinger S, Baron U, Thellmann M et al (2000) Exploring the sequence space for tetracycline-
dependent transcriptional activators: novel mutations yield expanded range and sensitivity.
PNAS 97:7963–7968. https://doi.org/10.1073/pnas.130192197
Van Wyk N, Kroukamp H, Pretorius IS (2018) The smell of synthetic biology: engineering strate-
gies for aroma compound production in yeast. Fermentation 4:54. https://doi.org/10.3390/fer
mentation4030054
Vashist SK, Mudanyali O, Schneider EM et al (2014) Cellphone-based devices for bioanalytical
sciences. Anal Bioanal Chem 406:3263–3277. https://doi.org/10.1007/s00216-013-7473-1
Venkatesh AG, Sun A, Brickner H et al (2015) Yeast dual-affinity biobricks: progress towards
renewable whole-cell biosensors. Biosens Bioelectron 70:462–468. https://doi.org/10.1016/j.
bios.2015.03.044
Versele M, Lemaire K, Thevelein JM (2001) Sex and sugar in yeast: two distinct GPCR systems.
EMBO Rep 2:574–579. https://doi.org/10.1093/embo-reports/kve132
Vieira Gomes AM, Souza Carmo T, Silva Carvalho L, et al (2018) Comparison of yeasts as hosts
for recombinant protein production. Microorganisms 6:38. https://doi.org/10.3390/microorga
nisms6020038
Vopálenská I, Váchová L, Palková Z (2015) New biosensor for detection of copper ions in
water based on immobilized genetically modified yeast cells. Biosens Bioelectron 72:160–167.
https://doi.org/10.1016/j.bios.2015.05.006
Wang L-J, Naudé N, Demissie M et al (2018) Analytical validation of an ultra low-cost mobile
phone microplate reader for infectious disease testing. Clin Chim Acta 482:21–26. https://doi.
org/10.1016/j.cca.2018.03.013
Wang J, Yang D, Guo X et al (2020) A novel RNA aptamer-modified riboswitch as chemical sensor.
Anal Chim Acta 1100:240–249. https://doi.org/10.1016/j.aca.2019.11.071
Weaver AA, Halweg S, Joyce M et al (2015) Incorporating yeast biosensors into paper-based ana-
lytical tools for pharmaceutical analysis. Anal Bioanal Chem 407:615–619. https://doi.org/10.
1007/s00216-014-8280-z
Wei Q, Qi H, Luo W et al (2013) Fluorescent imaging of single nanoparticles and viruses on a
smart phone. ACS Nano 7:9147–9155. https://doi.org/10.1021/nn4037706
Yoon HK, Jung ST, Kim J-H, Yoo TH (2012) Recent development of highly sensitive protease
assay methods: signal amplification through enzyme cascades. Biotechnol Bioproc E 17:1113–
1119. https://doi.org/10.1007/s12257-012-0545-9
116 H. A. M. Shepherd et al.
Yu D, Liao L, Zhang J et al (2018) A novel, easy and rapid method for constructing yeast two-
hybrid vectors using in-fusion technology. Biotechniques 64:219–224. https://doi.org/10.2144/
btn-2018-0007
Yudina NY, Arlyapov VA, Chepurnova MA et al (2015) A yeast co-culture-based biosensor for
determination of waste water contamination levels. Enzyme Microb Technol 78:46–53. https://
doi.org/10.1016/j.enzmictec.2015.06.008
Zhang CC, Glenn KA, Kuntz MA, Shapiro DJ (2000) High level expression of full-length estrogen
receptor in Escherichia coli is facilitated by the uncoupler of oxidative phosphorylation, CCCP.
J Steroid Biochem Mol Biol 74:169–178. https://doi.org/10.1016/s0960-0760(00)00120-5
Zhang J, Jensen MK, Keasling JD (2015) Development of biosensors and their application
in metabolic engineering. Curr Opin Chem Biol 28:1–8. https://doi.org/10.1016/j.cbpa.2015.
05.013
Zhao H, Zhang Y, Pan M et al (2019) Dynamic imaging of cellular pH and redox homeostasis with
a genetically encoded dual-functional biosensor, pHaROS, in yeast. J Biol Chem 294:15768–
15780. https://doi.org/10.1074/jbc.RA119.007557
Zhou X, Vink M, Klaver B et al (2006) Optimization of the Tet-On system for regulated gene
expression through viral evolution. Gene Ther 13:1382–1390. https://doi.org/10.1038/sj.gt.330
2780
Applications of Yeast Synthetic Biology
Yeast Synthetic Biology Approaches
for the Production of Valuable
Polyphenolic Compounds
Abstract
Polyphenols are secondary metabolites isolated from plants that are known
for their biological and therapeutical properties. However, the extraction from
plants renders low yields, besides being expensive and not environmentally
friendly. Therefore, the use of heterologous microorganisms, so-called micro-
bial chassis, became an interesting alternative approach to produce polyphenols.
With the advances in the metabolic engineering and synthetic biology fields, the
development of such microbial chassis able to produce these compounds with
higher yields and productivities became easier. Several yeast species, such as
Saccharomyces cerevisiae, Yarrowia lipolytica, and Pichia pastoris, have already
been engineered to produce some polyphenols. However, there is still a long
way to go before these compounds can be produced by heterologous organisms
at an industrial scale. In this chapter, we review the recent advances in the pro-
duction of polyphenolic compounds using yeasts as heterologous hosts, as well
as several synthetic biology approaches used to improve such production.
1 Introduction
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 119
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_5
120 D. Gomes et al.
2 Polyphenolic Compounds
3 Biological Activities
3.1 Antioxidant
that naringenin pretreatment improves the memory and learning capacities in Aβ-
injected rat model. These beneficial effects were achieved due to the decrease
in lipid peroxidation and apoptosis in the hippocampus. Moreover, curcumin was
found to protect biomembranes by scavenging oxygen free radicals like hydroxyl
radicals and superoxide anions responsible for initiation of lipid peroxidation in
liver and kidneys of diabetic db/db mice. These results showed that curcumin could
be used to prevent oxidative stress in diabetes patients (Soto-Urquieta et al. 2014).
3.2 Anticancer
Cancer is considered one of the most frequent causes of death worldwide, and
it is considered a health problem due to its high incidence in society. In 2018,
the World Health Organization estimated that among 18.1 million cases of cancer,
9.6 million cases would result in death (World Health Organization 2018).
To reduce cancer mortality, the discovery of new therapeutic compounds
became a priority. Polyphenols have been reported as potential therapeutic com-
pounds for several types of cancer. These compounds can interfere in cell
proliferation, tumor initiation and proliferation, metastasis, angiogenesis, and
apoptosis (Niedzwiecki et al. 2016). These effects are attributed to the ability of
polyphenols to interact with transcription factors, signaling molecules, growth and
apoptotic regulators, and adhesion molecules (Avtanski and Poretsky 2018; Khan
et al. 2020; Park 2015). For example, it was shown that curcumin could decrease
the Warburg effect in different cancer cell lines. The Warburg effect is character-
ized by the high productions of lactate and high glucose uptake by tumor cells
even when oxygen is available. The reduction of the Warburg effect was achieved
due to the ability of curcumin to downregulate pyruvate kinase M2 expression
by regulating mammalian target of rapamycin (mTOR)/ hypoxia-inducible factor
1–α (HIF1α) pathway. Additionally, it was also demonstrated that the viability of
cancer cells was reduced after curcumin treatment (Siddiqui et al. 2018). Eriodyc-
tiol also showed anticancer effect against human lung cancer cell line A549. This
effect was characterized by downregulating the expression of the apoptotic protein
Bcl-2 and upregulating the expression of Bax protein. Consequently, apoptosis
was induced. Additionally, phosphatidylinositol 3 kinase (PI3K)/protein kinase B
(AKT)/mTOR signaling pathway was also inhibited. All these modifications in the
protein expression have contributed to the inhibition of the cancer cell’s growth
(Zhang et al. 2020). Another polyphenolic compound that has been widely stud-
ied as an anticancer agent is quercetin. This flavonoid showed activity against
several types of cancer cell lines. For example, Hashemzaei et al. (2017) have
studied the effect of quercetin treatment in nine different tumor cell lines. In all
the tested tumor cell lines, it was concluded that quercetin induces apoptosis.
Additionally, the effect of quercetin in vivo was also assessed in mice bear-
ing estrogen receptor-positive breast cancer MCF-7 and colon carcinoma CT-26
tumors. After the treatment with quercetin, a significant reduction in the tumor
volume was found. The authors concluded that quercetin has the potential to be
Yeast Synthetic Biology Approaches for the Production of Valuable … 123
used as an anticancer agent due to its in vitro and in vivo effects. The anticancer
properties of resveratrol have been also studied. For example, Zeng et al. (2017)
studied the effect of resveratrol in human colon cancer cells. The authors have
found that resveratrol inhibits the proliferation and induces the apoptosis of these
cells. Moreover, it was also reported that resveratrol upregulates the activity of
bone morphogenetic protein 7 (BMP7). BMP7 is considered an effective tumor
suppressor that interacts with the PI3K/AKT signaling pathway. This signaling
pathway was inactivated by resveratrol, and its inactivation was potentiated by
BMP7, thus suggesting that resveratrol could be applied in the treatment of colon
cancer. In addition to these studies, there are many more that demonstrate the abil-
ity of polyphenols to act on several cancer types (Abbaszadeh et al. 2019). For this
reason, these compounds have a lot of potentials to be used as anticancer drugs.
3.3 Antiviral
Beyond the antioxidant and anticancer activities, polyphenols have also been
reported as natural compounds with antiviral properties due to the presence of
hydroxyl and ester groups (Kamboj et al. 2012). Several studies have reported the
antiviral activity of polyphenols against a large variety of virus, such as herpes
simplex virus (HSV), dengue virus (DENV), zika virus, human immunodeficiency
virus (HIV), severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2),
among others (Ali and Banerjea 2016; Chen et al. 2012; Jasso-Miranda et al.
2019; Lee et al. 2017; Paraiso et al. 2020; Vázquez-Calvo et al. 2017). For exam-
ple, the in vitro effect of a mixture of polyphenols extracted from almond skin in
HSV type I was tested. The treatment with 0.4 mg/mL of the mixture contain-
ing catechin, naringenin-7-O-glucoside, kaempferol-3-O-glucoside, epicatechin,
isorhamnetin-3-O-rutinoside, and isorhamnetin-3-O-glucoside led to a reduction
in the expression of viral proteins ICP0, UL42, and Us11. Moreover, the viral
DNA accumulation was lower compared to the control (infected cells without any
treatment). The results showed that the mixture of polyphenols tested is efficient in
the treatment of HSV type I infected cells (Musarra-Pizzo et al. 2019). Quercetin
has also shown activity against HSV type I. In this study, Raw 264.7 cells were
infected with HSV type I in the presence of quercetin at different concentrations.
Quercetin reduced the expression of HSV-1 glycoprotein D that is essential for
viral entry into the host cells and ICP0 that is expressed in the HSV-1 replica-
tion cycle. The effect on the viral proteins responsible for the virus replication
was also studied. The levels of expression of ICP0, UL13, and UL52 were signifi-
cantly reduced in cells treated with quercetin. With these results, it was possible to
conclude that quercetin affects the virus entrance in the host cells and its further
replication (Lee et al. 2017). Another compound with reported antiviral proper-
ties is curcumin. For example, this compound has shown activity against HIV by
the degradation of the Tat protein that is responsible for viral replication. More-
over, the treatment of infected cells with this compound resulted in significant
inhibition of virus production (Ali and Banerjea 2016). Polyphenols, like fisetin
124 D. Gomes et al.
and quercetin, were also studied as efficient antiviral compounds in DENV infec-
tion. Fisetin and quercetin were found to inhibit DENV-2 and DENV-3 infections.
Both compounds exhibited an effect in the inflammatory response by decreas-
ing the production and secretion of tumor necrosis factor-alpha (TNF-α) that is
related to severe DENV infections (Jasso-Miranda et al. 2019). Recently, polyphe-
nols have also shown potential activity against SARS-CoV2 (Paraiso et al. 2020).
It was found that angiotensin-converting enzyme-2 (ACE2) cellular receptor is
downregulated during SARS-Cov2 infection, and this downregulation could be
the reason for the pathogenesis and progression of SARS-CoV2. Horne and Vohl
(2020) have suggested that resveratrol has the potential to control the pathogenesis
of SARS-Cov2 by upregulating the ACE2 cellular receptor.
Polyphenols have also wound healing properties. The wound healing process
involves three different phases that are the inflammatory phase, proliferative phase,
and remodeling phase (Ibrahim et al. 2018). For example, curcumin has been found
to accelerate the wound closure in mice with excisional wound healing. This effect
is related to a reduction in the expression of TNF-α and matrix metallopeptidase
9, which are up-regulated in the presence of wounds, by inhibiting the NF-kB
signaling pathway. Moreover, this compound was also able to increase cell pro-
liferation and collagen synthesis (Yen et al. 2018). Quercetin has shown positive
effects on the wound healing process in pressure ulcer lesions. This compound was
also found to accelerate wound closure. Moreover, the production of TNF-α and
IL-1β was also significantly reduced through suppression of the MAPK signaling
pathway (Yin et al. 2018).
4 Biosynthetic Pathway
Fig. 2 Schematic representation of the shikimate pathway involved in the synthesis of the
aromatic amino acids L-tyrosine and L-phenylalanine. AAT, aromatic amino acid aminotrans-
ferase; ADH, arogenate dehydrogenase; ADT, arogenate dehydratase; CM, chorismate mutase;
CS, chorismate synthase; DAHP, 3-deoxy-D-arabino-heptulosonic acid 7-phosphate; DAHPS,
3-deoxy-D-arabino-heptulosonic acid 7-phosphate synthase; DHQ, 3-dehydroquinate; DHQD,
3-dehydroquinate dehydratase; DHQS, 3-dehydroquinate synthase; E4P, erythrose-4-phosphate;
EPSP, 5-enolpyruvylshikimate 3-phosphate; EPSPS, 5-enolpyruvylshikimate 3-phosphate syn-
thase; PAT, prephenate aminotransferase; PDH, prephenate dehydrogenase; PDT, prephenate dehy-
dratase; PEP, phosphoenol pyruvate; S3P, shikimate 3-phosphate; SDH, shikimate dehydrogenase;
SK, shikimate kinase
and arogenate by the action of aromatic amino acid aminotransferase (AAT) and
arogenate dehydrogenase (ADH), respectively (Tzin and Galili 2010).
L-tyrosine and L-phenylalanine are converted into the hydroxycinnamic acids
through the phenylpropanoid pathway (Fig. 3). When L-phenylalanine is used as a
substrate, the first enzyme of this biosynthetic pathway is phenylalanine ammonia-
lyase (PAL). This enzyme performs the deamination of L-phenylalanine to produce
cinnamic acid. Afterward, cinnamic acid is hydroxylated by cinnamic acid 4-
hydroxylase (C4H). In this step, p-coumaric acid is produced (Fraser and Chapple
2011). Alternatively, p-coumaric acid could be produced in a single enzymatic
step from L-tyrosine. This alternative route uses tyrosine ammonia lyase (TAL) to
perform the deamination of L-tyrosine (Watts et al. 2006).
Subsequently, p-coumaric acid could be used to produce other hydroxycin-
namic acids. By the action of 4-coumarate 3-hydroxylase (C3H), p-coumaric acid
is converted into caffeic acid. Moreover, caffeic acid could be converted into
ferulic acid by caffeic acid 3-O-methyltransferase (COMT). These hydroxycin-
namic acids could be activated to the correspondent coenzyme A (CoA) esters by
Yeast Synthetic Biology Approaches for the Production of Valuable … 127
Fig. 3 Schematic representation of the biosynthetic pathway responsible for the production of
polyphenols (curcuminoids, resveratrol, and naringenin chalcone) using aromatic amino acids as
substrates. The enzymes involved in the phenylpropanoid pathway are phenylalanine ammonia-
lyase (PAL), tyrosine ammonia-lyase (TAL), cinnamic acid 4-hydroxylase (C4H), 4-coumarate
3-hydroxylase (C3H), caffeic acid 3-O-methyltransferase (COMT), and 4-coumarate-CoA lig-
ase (4CL). p-coumaroyl 5-O-shikimate 3’-hydroxylase (CS3’H), p-coumaroyl shikimate trans-
ferase (CST), and caffeoyl-CoA 3-O methyltransferase (CCoAOMT) catalyze the conversion
of p-coumaric acid into other hydroxycinnamic acids. The Type III polyketide synthase (PKS)
enzymes responsible for curcuminoids, stilbenoids, and flavonoids scaffold production, which are
underlined, are diketide-CoA synthase (DCS), curcumin synthase (CURS), curcuminoids synthase
(CUS), chalcone synthase (CHS), and stilbene synthase (STS)
Polyphenols, like other plant secondary metabolites, are produced and accumu-
lated in very low amounts in plants. Due to the low amounts produced in plants,
the extraction of polyphenols from their natural resources is very difficult and the
yields are considered low (Krivoruchko and Nielsen 2015; Rainha et al. 2020a;
Rodrigues et al. 2015c; Rodrigues and Rodrigues 2020). Moreover, the amounts
of these compounds are affected by climatic and seasonal variations. The chemi-
cal synthesis could be an alternative to produce some polyphenols. However, some
of these compounds have complex structures being hard to chemically synthesize.
Additionally, chemical synthesis involves the use of toxic compounds and expen-
sive substrates. For this reason, the process is considered non-environmentally
friendly and expensive (Liu et al. 2017). Due to their potential applications and to
try to satisfy their industrial demand, there has been an increased interest in the
heterologous production of polyphenols.
The heterologous production in plants has the advantage of only requiring the
introduction of one or two genes of the biosynthetic pathway (Rodrigues et al.
2015c). The other genes from the pathway, such as the ones responsible for the
phenylpropanoid pathway, are already present in the plant. However, heterologous
production in plants has also disadvantages. One of these disadvantages is the
long period that the plant takes to grow. The growth is also dependent on the
climacteric conditions and fertile land occupation. Additionally, the production
yields are usually very low and variable due to cellular heterogeneity. Beyond
this, the downstream purification process is more difficult since several similar
compounds are also produced in plants (Ebrahimi and Mokhtari 2017; Wilson and
Roberts 2012).
Yeast Synthetic Biology Approaches for the Production of Valuable … 129
et al. 2020). Due to these unique features, S. cerevisiae, Y. lipolytica, and P. pas-
toris were exploited as a microbial chassis for the heterologous production of
several compounds, including polyphenols.
Table 1 (continued)
Class Compound Strain Genesa Modifications in the Substrate Production (mg/L) References
chassis
Flavonoids Naringenin S. cerevisiae BY4741 FjTAL, At4CL, Construction of Glucose (20 g/L) 220 Lyu et al. (2019)
HaCHS, and aromatic amino acids
PhCHI overproducing strain
Kaempferol S. cerevisiae BY4741 FjTAL, At4CL, Construction of Glucose (20 g/L) 168.1 Du et al. (2020)
AtCHS, AtCHI, aromatic amino acids
NtF3H, and and malonyl-CoA
AtFLS overproducing strain
Quercetin FjTAL, At4CL, 154.2
AtCHS, AtCHI,
NtF3H, PhF3’H,
and AtFLS
Myrcetin FjTAL, At4CL, 145
AtCHS, AtCHI,
NtF3H, PhF3’H,
SlF3 5’H, and
AtFLS
Delphinidin FjTAL, At4CL, 26.1
AtCHS, AtCHI,
NtF3H, PhF3’H,
SlF3 5’H,
AaDFR, and
GeANS
Pelargonidin FjTAL, At4CL, 33.3
AtCHS, AtCHI,
NtF3H, AaDFR,
and GeANS
Cyanidin FjTAL, At4CL, 31.7
AtCHS, AtCHI,
NtF3H, PhF3’H,
AaDFR, and
GeANS
(continued)
D. Gomes et al.
Table 1 (continued)
Class Compound Strain Genesa Modifications in the Substrate Production (mg/L) References
chassis
Flavonoids Naringenin Y. lipolytica Po1f SeTAL, Nt4CL, Construction of Glucose (80 g/L) 898 Palmer et al.
and HsCHS aromatic amino acids (2020)
and malonyl-CoA
overproducing strain
Eriodictyol Y. lipolytica Po1f RtTAL, Pc4CL, Construction of a Glucose (40 g/L) 134.2 Lv et al. (2019b)
PhCHS, MsCHI, strain with improved
GhF3’H, and chorismate and
CrCPR malonyl-CoA
Taxifolin RtTAL, Pc4CL, biosynthesis 110.5
PhCHS, MsCHI,
GhF3’H, CrCPR,
and SlF3H
Liquiritigenin Y. lipolytica (ATCC201249) ZmPAL, PcC4H, – Glucose (20 g/L) + 62.4 Akram et al.
Pc4CL, PhCHS, p-Coumaric acid (2020)
MsCHR, and (100 mg/L)
MsCHI
8-hydroxydaidzein P. pastoris X-33 AoCYP57B3 and – Daidzein (25.4 mg/L) 0.58 Chang et al.
3 -hydroxydaidzein BmBM3R 0.23 (2013)
6-hydroxydaidzein 9.1
3’-hydroxygenistein P. pastoris X-33 AoCYP57B3 and Strain modified by Genistein (135 mg/L) 20.3 Wang et al.
ScCPR treatment with (2016)
hydrogen peroxide
to induce oxidative
Yeast Synthetic Biology Approaches for the Production of Valuable …
stress and,
consequently, select
the most resistant
strains
(continued)
133
134
Table 1 (continued)
Class Compound Strain Genesa Modifications in the Substrate Production (mg/L) References
chassis
Stilbenoids Resveratrol S. cerevisiae AtPAL, AtC4H, Construction of Glucose (88 g/L) 800 Li et al. (2016)
CEN.PK102-5B At4CL, and aromatic amino acids
VvVST and malonyl-CoA
Pterostilbene AtPAL, AtC4H, overproducing strain 34.9
At4CL, VvVST,
and VvROMT
Pinostilbene AtPAL, AtC4H, 5.52
At4CL, VvVST,
and SbROMT
Resveratrol Y. lipolytica W29 FjTAL, At4CL, Construction of Glucose (40 g/L) 12,400 Sáez-Sáez et al.
and VvSTS aromatic amino acids (2020)
and malonyl-CoA
overproducing strain
(continued)
D. Gomes et al.
Table 1 (continued)
Class Compound Strain Genesa Modifications in the Substrate Production (mg/L) References
chassis
Coumarins Scopoletin S. cerevisiae BY4742 Pc4CL, AtF6 H1, – Lignin hydrolysate 4.79 Zhao et al.
PaHpaB, and containing p-coumaric (2020)
EnHpaC acid (2.3 g/L) and ferulic
acid (0.5 g/L)
Curcuminoids Bisdemethoxycurcumin Y. lipolytica Po1f Nt4CL and Construction of p-Coumaric acid 0.17 Palmer et al.
OsCUS aromatic amino acids (0.33 g/L) (2020)
and malonyl-CoA
overproducing strain
Polyphenolic p-Coumaroyl-3-hydroxyanthranilic S. cerevisiae CENPK113-5d Nt4CL and – p-Coumaric acid 120 Moglia et al.
amides acid ccHCT (462 mg/L) and (2015)
3-Hydroxyanthranilic
acid (77 mg/L)
Caffeoyl-3-hydroxyanthranilic acid Caffeic acid (540 mg/L) 22
and 3-Hydroxyanthranilic
acid (77 mg/L)
a Aa—Anthurium andraeanum, At—Arabidopsis thaliana, Ao—Aspergillus oryzae, Bm—Bacillus megaterium, Cc—Cynara cardunculus, Cr—Catharanthus roseus, En—Enterobacteriaceae, Fj—Flavobac-
terium johnsoniaeu, Ge—Gerbera specie, Gh—Gerbera hybrid, Ha—Hypericum androsaemum, Hs—Huperiza serrata, Ms—Mendicago sativa, Nt—Nicotiana tabacum, Os—Oryza sativa, Pa—Pseudomonas
aeruginosa, Pc—Petroselinum crispum, Ph—Populus hybrid, Rc—Rhodobacter capsulatus, Rt—Rhodosporidium toruloides, Sb—Sorghum bicolor, Sc—Saccharomyces cerevisiae, Se—Salmonella enterica,
Sl—Solanum lycopersicum, Tf—Thermobifida fusca NTU22, Vv—Vitis vinifera, Zm—Zea mays; 4CL—4-coumarate-CoA ligase, ANS—anthocyanidin synthase, AXE—thermostable esterase, BM3R—reductase
domain of the CYP102A1 gene, C3H—4-coumarate 3-hydroxylase, C4H— cinnamic acid 4-hydroxylase, CHR—chalcone reductase, CHI—chalcone isomerase, CHS—chalcone synthase, CPR- cytochrome
P450 reductase, CUS—curcuminoid synthase, CYP57B3—cytochrome P450 monooxygenase, DFR—dihydroflavonol 4-reductase, F3H—flavanone 3-hydroxylase, F3’H—flavonoid 3’-hydroxylase, F3´5´H—
flavonoid 3’,5’-hydroxylase, F6’H1—feruloyl-CoA 6’-hydroxylase, FLS—flavonol synthase, FSI—flavone synthase I, HpaB and HpaC—4-hydroxyphenylacetate 3-hydroxylase, PAL— phenylalanine ammonia
lyase, ROMT—resveratrol O-methyltransferase, STS—stilbene synthase, TAL—tyrosine ammonia lyase, VST—resveratrol synthase; b Corncob biomass (2%) contains feruloyl-polysaccharide that is AXE
substrate
Yeast Synthetic Biology Approaches for the Production of Valuable …
135
136 D. Gomes et al.
Fig. 4 Structure of the hydroxycinnamic compounds p-coumaric acid, caffeic acid, and ferulic
acid
acids since they are present in limited amounts in the cells. The constructed strain
holds a knockout in the pyruvate decarboxylase (PDC) genes, PDC5 and Aro10.
These genes are responsible for the conversion of the precursor of phenylalanine,
that is, phenylpyruvate, into phenylacetaldehyde contributing to the pathway devi-
ation. Moreover, DAHP synthase and chorismate mutase genes were replaced by a
feedback-insensitive version to prevent the inhibition of aromatic amino acids syn-
thesis since these genes are strongly affected by the amounts of aromatic amino
acids produced. Moreover, the authors have overexpressed a shikimate kinase from
E. coli since this could be a limiting step in the production of higher amounts of
p-coumaric acid. TAL from Flavobacterium johnsoniae was also expressed in this
engineered strain that was able to produce almost 2000 mg/L of p-coumaric acid in
fed-batch fermentation (Rodriguez et al. 2015). Later, the same group studied the
influence of the production of p-coumaric acid in the metabolism of S. cerevisiae-
producing strains to construct more efficient and robust strains (Rodriguez et al.
2017a). After transcriptomic and metabolomic analysis, the authors have found
that genes involved in the sugars and amino acids transport were downregulated
in the producer strain. When these downregulated genes were knocked out, it was
found that the production of p-coumaric acid was improved. The highest produc-
tion of p-coumaric acid (2400 mg/L) was obtained in the S. cerevisiae CEN.PK
strain that holds the deletion of the tyrosine and tryptophan amino acid transporter
1 (TAT1), which is probably involved in the transport of tyrosine to the outside
of the cells being less available for p-coumaric acid production (Rodriguez et al.
2017a). More recently, Liu et al. (2019a) have also constructed a S. cerevisiae strain
able to produce higher amounts of p-coumaric acid. In this strain, PAL and C4H
genes from Arabidopsis thaliana were expressed to produce p-coumaric acid. To
increase the cytochrome P450 activity, cytochrome B5 (CYB5) from S. cerevisiae
and cytochrome P450 reductase (CPR1) from A. thaliana were overexpressed.
Moreover, the strain was also modified to overexpress a feedback-insensitive ver-
sion of the DAHP synthase and chorismate mutase genes. The TAL gene from F.
johnsoniae was also expressed, and the production of p-coumaric acid was verified
using both tyrosine and phenylalanine branches. Prephenate dehydrogenase (PDH)
genes from different microorganisms were also tested, and it was found that PDH
from Medicago truncatula significantly improves the shikimate pathway flux and,
consequently, the production of p-coumaric acid. The production of E4P was also
improved by expressing a heterologous phosphoketolase pathway in conjugation
Yeast Synthetic Biology Approaches for the Production of Valuable … 137
Within polyphenolic compounds, flavonoids are the class with more variety of
structures. These compounds are responsible for organoleptic properties and for the
pigments that confer color to vegetables, fruits, and herbs. Due to their medical and
industrial applications, the market value of these compounds is estimated to reach
USD 1.06 billion by 2025 (Grand View Research 2016). The flavonoid skeleton
is characterized by two benzene rings that are connected by one pyrene ring that
contains an oxygen molecule (Fig. 5). Depending on the differences in the struc-
ture of these compounds, flavonoids could be grouped into different subclasses.
These subclasses are flavonols, flavones, flavanones, flavonols, isoflavonoids, and
anthocyanins (Panche et al. 2016).
138 D. Gomes et al.
Fig. 5 Basic structure of flavonoids compounds. A and B correspond to two benzene rings and C
corresponds to a pyrene ring containing an oxygen molecule
In the last years, several research efforts have been made to construct yeast cell
factories with the ability to produce flavonoids with high titers and yields. S. cere-
visiae has been the most explored yeast to produce these compounds. Several S.
cerevisiae strains were constructed to produce flavonoids using p-coumaric acid or
amino acids as precursors (Jiang et al. 2005; Lyu et al. 2017; Trantas et al. 2009;
Yan et al. 2005). However, these substrates are more expensive, and the industrial
production process was unfeasible (Santos et al. 2011). For this reason, de novo
biosynthesis of these compounds from cheap carbon sources became a priority.
Koopman et al. (2012) have constructed a S. cerevisiae strain able to produce narin-
genin from glucose. This strain harbors the genes responsible for the naringenin
biosynthetic pathway (PAL1, C4H, CPR1, 4CL3, CHS3, and chalcone isomerase
(CHI1)) from A. thaliana. The production of naringenin was evaluated in shake
flasks fermentations, and it was found that low amounts of naringenin were pro-
duced (1.47 mg/L). The authors have modified the S. cerevisiae strain to construct a
new yeast chassis able to overproduce amino acids and, consequently, optimize the
production yields and titers. In this strain, the DAHP synthase gene was replaced
for a feedback-insensitive version of the gene. PDC gene was also deleted to pre-
vent the pathway deviation. Additionally, the number of copies of CHS gene was
also increased to improve its catalytic efficiency. To improve the production of
p-coumaric acid, TAL from R. capsulatus was also expressed. These modifications
have resulted in an improvement of naringenin production reaching 54.5 mg/L in
shake flask cultures and 108.9 mg/L in aerated, pH-controlled batch reactors from
20 g/L of glucose (Koopman et al. 2012). Rodriguez et al. (2017b) have also con-
structed S. cerevisiae strains able to produce several flavonoids using glucose as a
substrate. In this work, the genes responsible for the biosynthesis of naringenin,
liquiritigenin, kaempferol, resokaempferol, quercetin, and fisetin were integrated
into the yeast genome. These strains contained modifications to improve the flux
toward amino acids, namely the knockout of the PDC5 and Aro10 genes and the
overexpression of DAHP synthase and chorismate mutase. The highest productions
obtained in this study were 26.57 mg/L of kaempferol and 20.38 mg/L of quercetin
(Rodriguez et al. 2017b). Duan et al. (2017) also constructed a S. cerevisiae strain
to produce kaempferol de novo. Since malonyl-CoA low availability represents
a limiting step in the high-level production of polyphenols, the authors overex-
pressed the endogenous genes responsible for the synthesis of malonyl-CoA from
Yeast Synthetic Biology Approaches for the Production of Valuable … 139
ethanol to increase its availability. Using a fed-batch fermentation strategy, the pro-
duction of kaempferol has reached 66.29 mg/L. More recently, Lyu et al. (2019)
also optimized a S. cerevisiae strain to improve the flux toward amino acids and
prevent the pathway deviation. As performed by other research groups, the authors
constructed a strain with a knockout in the PDC5 and Aro10 genes. Additionally,
the genes responsible for the biosynthetic pathway were overexpressed and the
promotors were replaced by the glucose-inducible strong promoter pGAL1. This
strain was able to produce 86 mg/L of kaempferol. Moreover, this strain was also
able to produce 220 mg/L of naringenin when Tween 80, that is a surfactant that
could increase the enzymatic activity and cell biomass growth, was added to the
fermentation medium. This is the highest de novo production of naringenin using
S. cerevisiae as chassis reported so far (Lyu et al. 2019).
The biosynthetic pathway responsible for the synthesis of anthocyanins has
been firstly engineered in S. cerevisiae by Eichenberger et al. (2018). In this study,
the genes responsible for the biosynthetic pathway were integrated into a single
copy in the yeast genome. The constructed strains were able to produce 0.85 mg/L
of pelargonidin-3-O-glucosidase, 1.55 mg/L of cyanidin-3-O-glucosidase and
1.86 mg/L of delphinidin-3-O-glucosidase (Eichenberger et al. 2018). Afterward,
a co-culture strategy was designed by Du et al. (2020) to produce naringenin,
kaempferol, quercetin, myricetin, delphinidin, pelargonidin, and cyanidin. In this
co-culture strategy, one strain harbors the biosynthetic pathway responsible for the
production of naringenin, and the other strains harbor the other genes responsi-
ble for the production of the three flavonols and the three anthocyanidins from
naringenin. The first strain contains several modifications to improve the pro-
duction of aromatic amino acids and malonyl-CoA and to prevent the deviation
of the pathway for competing pathways. The first strain was able to produce
144.1 mg/L of naringenin. The co-culture of both strains resulted in the pro-
duction of 168.1 mg/L of kaempferol, 154.2 mg/L of quercetin, 145 mg/L of
myrcetin, 26.1 mg/L delphinidin, 33.3 mg/L of pelargonidin, and 31.7 mg/L of
cyanidin. This study demonstrated the importance of using co-culture strategies to
reduce the metabolic burden caused by the presence of large biosynthetic pathways
(Du et al. 2020; Rodrigues et al. 2020). Additionally, these studies strengthened
the relevance of improving the synthesis of aromatic amino acids and the exten-
der substrates availability to significantly increase the production of the desired
compounds.
Although Y. lipolytica has been less used than S. cerevisiae, it has been recently
explored as a yeast chassis for the production of flavonoids. Lv et al. (2019a) have
constructed the biosynthetic pathways responsible for the production of naringenin,
eriodyctiol, and taxifolin. The authors have developed a method for integration
of the genes responsible for the biosynthetic pathway in multiple copies in ran-
dom sites of the Y. lipolytica genome by combining 26 s ribosomal DNA (rDNA)
multi-copy integration with Cre-loxP system. For the integration into the yeast
genome, this methodology uses the sequence of the repetitive 26 s rDNA sites of
Y. lipolytica genome as regions of homology for the cassette integration. Beyond
the genes of the biosynthetic pathway, the cassette also contains the selection
140 D. Gomes et al.
marker URA3 with loxP flanking sequences. After integration in the genome and
selection of the transformants, Cre recombinase is expressed and recognizes the
loxP sequences and, consequently, the selection marker is removed. The devel-
opment of this method of integration was important because this microorganism
has few auxotrophic markers available. Moreover, the use of plasmids can lead to
genomic instability as well as copy number variations. After integrating the genes
responsible for the biosynthetic pathways, the authors have evaluated the abil-
ity of the constructed strains to produce naringenin, eriodyctiol, and taxifolin in
Yeast Extract–Peptone–Dextrose (YPD)-rich medium. The best performing strains
were able to produce 71.2 mg/L, 54.2 mg/L, and 48.1 mg/L of naringenin, eri-
odyctiol, and taxifolin, respectively (Lv et al. 2019a). Later, the same research
group has performed several steps to optimize the production of these compounds.
The authors have optimized the copy number of the CHS and CPR genes that
are responsible for limiting steps in the production of these flavonoids. Moreover,
Lv et al. (2019b) also enhanced the availability of the chorismate and malonyl-
CoA precursors by overexpressing the genes responsible for the synthesis of both
compounds. Moreover, fermentation parameters such as pH and C/N ratio were
also optimized. In the best conditions, the optimized strains were able to produce
252.4 mg/L, 134.2 mg/L, and 110.5 mg/L of naringenin, eriodictyol, and taxi-
folin, respectively. Comparing with the previous study, these optimizations led to
3.5-fold, 2.5-fold, and 2.3-fold improvement in the production of naringenin, eri-
odictyol, and taxifolin, respectively (Lv et al. 2019b). More recently, a mixture
of glucose and xylose was used as a substrate to produce naringenin (Wei et al.
2020). Due to the need to produce flavonoids in an inexpensive way, the search for
low-cost substrates, such as agricultural residues, has become a priority to reduce
the production costs of these compounds at an industrial scale (Gudiña et al. 2020).
Since xylose is one of the sugars that is present in a higher amount in agricultural
residues, Wei et al. (2020) engineered the Y. lipolytica strain toward the use of
xylose as a substrate to produce naringenin since the wild-type strain is not able
to metabolize xylose. This strain was constructed by integration into the genome
of a xylose-inducible module for activation of the expression of the genes respon-
sible for xylose utilization. By activating the expression of these genes, xylose is
metabolized and originates E4P which is a precursor of the shikimate pathway.
Moreover, the genes responsible for the naringenin biosynthetic pathway were
also integrated into the genome by Clustered Regularly Interspaced Short Palin-
dromic Repeats—associated Cas9 endonuclease (CRISPR-Cas9). The production
of naringenin was evaluated in the strain that only contains the naringenin biosyn-
thetic pathway and in the strain that contains the naringenin biosynthetic pathway
and the xylose-inducible utilization module. The first strain was able to produce
239.1 mg/L of naringenin in YPD fermentation medium. The strain that also con-
tains the xylose-inducible utilization module was able to produce 715.3 mg/L of
naringenin in YPD fermentation medium containing 60 g/L of xylose. This study
showed that the xylose utilization pathway is a good alternative to be explored to
increase the precursor availability and, consequently, improve the production of
the desired polyphenolic compound. Moreover, this study opens doors to explore
Yeast Synthetic Biology Approaches for the Production of Valuable … 141
the flux toward the aromatic amino acids production was improved by replacing
DAHP synthase and chorismate mutase by feedback-resistant versions of these
genes. Malonyl-CoA synthesis was also improved by expressing the acetyl-CoA
synthase from Salmonella enterica and the mutated version of the ACC gene.
Beyond these modifications, the pathway deviation was also prevented by delet-
ing the PDC gene. In this study, two stilbenoids that are derived from resveratrol
were also produced. By expressing resveratrol O-methyltransferase (ROMT) from
Sorghum bicolor, the constructed strain produced 5.52 mg/L of pinostilbene. By
expressing ROMT from V. vinifera, the strain produced 34.92 mg/L of pterostil-
bene (Li et al. 2016). Pterostilbene was also previously produced using p-coumaric
acid as a substrate in S. cerevisiae expressing 4CL, STS, and ROMT (Wang et al.
2015a). However, the production was even lower (2.2 mg/L) than the one reported
by Li et al. (2016).
The heterologous production of resveratrol was also achieved using Y. lipolytica
as microbial chassis. Huang et al. (2006) reported for the first time the production
of resveratrol in Y. lipolytica. The strain expressing PAL, C4H, 4CL, and STS
was able to produce 1.46 mg/L of resveratrol from 36.2 mg/L of tyrosine. More
recently, there have been many advances to produce larger amounts of resveratrol
in Y. lipolytica. Beyond the construction of the p-coumaric acid biosynthetic path-
way, Gu et al. (2020) also constructed the pathway responsible for the resveratrol
production. The authors expressed 4CL and STS in the strain previously con-
structed to produce p-coumaric acid. Although this strain has several modifications
to improve the flux toward the synthesis of aromatic amino acids and the shikimate
pathway, only 12.67 mg/L of resveratrol were produced from 40 g/L of glucose.
Palmer et al. (2020) have also reported the construction of a Y. lipolytica strain
able to de novo produce resveratrol. This strain also contains modifications to
improve the synthesis of malonyl-CoA and amino acids, and it was able to produce
48.7 mg/L of resveratrol from 20 g/L of glucose and 32.8 mg/L of p-coumaric acid.
Afterward, the heterologous production of resveratrol in Y. lipolytica was optimized
by Sáez-Sáez et al. (2020). In this study, TAL from Flavobacterium johnsoniae,
4CL from A. thaliana and STS from V. vinifera were successfully integrated into
the Y. lipolytica genome using integrative plasmids. DAHP synthase and choris-
mate mutase were replaced by feedback-insensitive versions. Moreover, PCD5 and
Aro10 were also deleted to prevent the pathway deviation. Two mutations were
introduced in the ACC gene to prevent the phosphorylation and inactivation of
malonyl-CoA synthesis. This strain was able to produce 85 mg/L of resveratrol.
In order to improve the resveratrol titer, five copies of the genes responsible for
the biosynthetic pathway were introduced using integrative plasmids in the pre-
viously constructed strain reaching the production of 409 mg/L of resveratrol. In
fed-batch fermentation, this strain was able to produce 12.400 mg/L of resvera-
trol from glucose. This is the highest production of resveratrol reported so far in
any microorganism (Sáez-Sáez et al. 2020). More recently, He et al. (2020) also
constructed a strain for de novo production of resveratrol. The authors studied
the production using the tyrosine or the phenylalanine routes or the conjugation
of both routes. The biosynthetic pathways were firstly integrated into a single
Yeast Synthetic Biology Approaches for the Production of Valuable … 145
copy in the Y. lipolytica genome. Afterward, one extra copy of the biosynthetic
pathways was also integrated to increase the metabolic flux toward the produc-
tion of resveratrol. When glucose was used as a substrate, the strain carrying two
copies of the tyrosine route was the best resveratrol producer (85 mg/L). Addi-
tionally, the performance of the three strains was also tested using glycerol as an
alternative substrate. In these experiments, the strain carrying two copies of both
tyrosine and phenylalanine routes showed the best results of resveratrol production.
In bioreactor fermentation, this strain was able to produce 430 mg/L of resveratrol
from 100 g/L of glycerol without any supplementation of amino acids (He et al.
2020). All these studies show the great potential to use Y. lipolytica as a chassis to
industrially produce resveratrol and other resveratrol-derived compounds.
Coumarins are polyphenolic compounds that are largely found in many plant
species being fundamentally present in essential oils and fruits (Jain and Joshi
2012). These compounds act in the defense mechanism of plants against stress
factors such as fungal infections (Stringlis et al. 2019). The core structure of
coumarins is composed of one benzene ring and one α-pyrone ring (Fig. 7). These
compounds are divided into two main classes that are simple coumarins and com-
plex coumarins. In the group of complex coumarins, the compounds are classified
as furanocoumarins, pyranocoumarins, phenylcoumarins, and bicoumarins (Matos
et al. 2015).
Relatively to the flavonoids and stilbenoids, the heterologous production of
coumarins has been less explored. Simple coumarins, such as umbelliferone,
scopoletin, and esculetin, were already produced in E. coli (Lin et al. 2013; Yang
et al. 2015; Zhao et al. 2019). However, the use of yeasts as microbial chassis
to produce these compounds practically has not been explored. Until now, only
scopoletin was already produced in S. cerevisiae (Zhao et al. 2020). The genes
4CL from Petroselinum crispum and feruloyl-CoA 6 -hydroxylase (F6 H1) from
A. thaliana responsible for the synthesis of scopoletin from ferulic acid were
expressed into S. cerevisiae BY4742. This strain was able to produce 0.364 mg/L
of scopoletin. Since these results of production were very low, the authors fused
4CL and F6’H1 with different linkers. The best production result was obtained
when the genes were fused with the linker (GGGGS)4 , resulting in a 3-fold
improvement in the production of scopoletin. After integrating the fusion gene
into the S. cerevisiae genome, 3.42 mg/L of scopoletin were produced using fer-
ulic acid as substrate. The authors also tested the production of scopoletin using
p-coumaric acid as substrate. HpaB from P. aeruginosa, HpaC from Enterobacte-
riaceae and COMT from A. thaliana were integrated into the genome of the strain
carrying the fusion gene 4CL-F6’H1. This strain was able to produce 4.98 mg/L of
scopoletin using p-coumaric acid as substrate. Due to the interest of using alterna-
tive substrates to produce compounds of interest, the production was also evaluated
using lignin hydrolysate as substrate (Gudiña et al. 2020). Since ferulic acid and
p-coumaric acid are monomers of lignin, this strain produced 4.79 mg/L of scopo-
letin (Zhao et al. 2020). As far as we know, this is the only report of heterologous
production of coumarins in yeasts. Although the production titers are lower com-
pared with the productions obtained in E. coli, we believe that this production could
be successfully improved and reach higher levels. Moreover, the construction of the
biosynthetic pathway to produce the other simple and complex coumarins should
also be explored using S. cerevisiae and other yeasts as chassis.
Curcuminoids are natural polyphenolic compounds that can be found in the rhi-
zome of the plant Curcuma longa which is widely known as turmeric. The
rhizome of turmeric is mainly composed of curcumin, demethoxycurcumin, and
bisdemethoxycurcumin. These compounds have been largely used as a food addi-
tive, as well as therapeutic agents due to their several recognized therapeutical
activities (Amalraj et al. 2017). For this reason, the market value of curcumin,
which is present in higher amounts in turmeric and exhibits the most potent bio-
logical activities, is expected to reach USD 151.9 million by 2027 (Hewlings and
Kalman 2017; Grand View Research 2020). The core structure of curcuminoids
is composed of a C6-C7-C6 structure being classified as diarylheptanoids (Fig. 8)
(Rodrigues et al. 2015c).
Fig. 8 Structure of the three curcuminoids compounds: curcumin, demethoxycurcumin, and bis-
demethoxycurcumin
Yeast Synthetic Biology Approaches for the Production of Valuable … 147
7 Conclusions
Acknowledgements This study was supported by the Portuguese Foundation for Science and
Technology (FCT) under the scope of the strategic funding of UIDB/BIO/04469/2020 unit and
BioTecNorte operation (NORTE-01-0145- FEDER-000004) funded by the European Regional
Development Fund (ERDF) under the scope of Norte2020—North Portugal Regional Pro-
gram. DG and JR are recipients of a fellowship supported by a doctoral advanced training
(SFRH/BD/04433/2020 and SFRH/BD/138325/2018, respectively) funded by FCT.
References
Abbaszadeh H, Keikhaei B, Mottaghi S (2019) A review of molecular mechanisms involved
in anticancer and antiangiogenic effects of natural polyphenolic compounds. Phytother Res
33:2002–2014
Abramovič H (2015) Antioxidant properties of hydroxycinnamic acid derivatives: a focus on bio-
chemistry, physicochemical parameters, reactive species, and biomolecular interactions. In:
Preedy VR (ed) Coffee in health and disease prevention. Elsevier Inc, pp 843–852
Akram M, Rasool A, An T, Feng X, Li C (2020) Metabolic engineering of Yarrowia lipolytica for
liquiritigenin production. Chem Eng Sci 230(2021):116177
Ali A, Banerjea AC (2016) Curcumin inhibits HIV-1 by promoting tat protein degradation. Sci Rep
6(27539):1–9
Amalraj A, Pius A, Gopi S, Gopi S (2017) Biological activities of curcuminoids, other
biomolecules from turmeric and their derivatives—a review. J Tradit Complement Med
7(2):205–233
Avtanski D, Poretsky L (2018) Phyto-polyphenols as potential inhibitors of breast cancer metasta-
sis. Mol Med 24(1):1–17
Becker JVW, Armstrong GO, Van Der Merwe MJ, Lambrechts MG, Vivier MA, Pretorius IS
(2003) Metabolic engineering of Saccharomyces cerevisiae for the synthesis of the wine-related
antioxidant resveratrol. FEMS Yeast Res 4(1):79–85
Beekwilder J, Wolswinkel R, Jonker H, Hall R, De Rie Vos CH, Bovy A (2006) Production of
resveratrol in recombinant microorganisms. Appl Environ Microbiol 72(8):5670–5672
Braga A, Ferreira P, Oliveira J, Rocha I, Faria N (2018) Heterologous production of resveratrol in
bacterial hosts: current status and perspectives. World J Microbiol Biotechnol 34(8):1–11
Braga A, Rocha I, Faria N (2019) Microbial hosts as a promising platform for polyphenol pro-
duction. In: Akhtar M, Swamy M, Sinniah U (eds) Natural bio-active compounds: volume 1:
production and applications vol 1. Springer, pp 71–103
Cassidy L, Fernandez F, Johnson JB, Naiker M, Owoola AG (2020) Broszczak DA (2020) Oxida-
tive stress in Alzheimer’s disease: a review on emergent natural polyphenolic therapeutics.
Complement Ther Med 49:102294
Chang TS, Chao SY, Chen YC (2013) Production of ortho-hydroxydaidzein derivatives by a
recombinant strain of Pichia pastoris harboring a cytochrome P450 fusion gene. Process
Biochem 48(3):426–429
Chen X, Qiao H, Liu T, Yang Z, Xu L, Xu Y, Ge HM, Tan RX, Li E (2012) Inhibition of her-
pes simplex virus infection by oligomeric stilbenoids through ROS generation. Antiviral Res
95(1):30–36
Chen L, Deng H, Cui H, Fang J, Zuo Z, Deng J, Li Y, Wang X, Zhao L (2018) Inflammatory
responses and inflammation-associated diseases in organs. Oncotarget 9(6):7204–7218
Chen R, Yang S, Zhang L, Zhou YJ (2020) Advanced Strategies for Production of Natural Products
in Yeast. Cell Press Reviews 23(3):100879
Couto MR, Rodrigues JL, Rodrigues LR (2017) Optimization of fermentation conditions for the
production of curcumin by engineered Escherichia coli. J R Soc Interface 14(133):20170470
150 D. Gomes et al.
Dayem AA, Choi HY, Yang GM, Kim K, Saha SK, Cho SG (2016) The anti-cancer effect of
polyphenols against breast cancer and cancer stem cells: molecular mechanisms. Nutrients
8(9):1–37
de Araújo FF, de Paulo Farias D, Neri-Numa IA, Pastore GM (2020) Polyphenols and their appli-
cations: an approach in food chemistry and innovation potential. Food Chem 338(2021):1–15
Ding L, Hofius D, Hajirezaei MR, Fernie AR, Börnke F, Sonnewald U (2007) Functional analysis
of the essential bifunctional tobacco enzyme 3-dehydroquinate dehydratase/shikimate dehydro-
genase in transgenic tobacco plants. J Exp Bot 58(8):2053–2067
Ding HY, Chiang CM, Tzeng WM, Chang TS (2015) Identification of 3 -hydroxygenistein as
a potent melanogenesis inhibitor from biotransformation of genistein by recombinant Pichia
pastoris. Process Biochem 50(10):1614–1617
Du Y, Yang B, Yi Z, Hu L, Li M (2020) Engineering Saccharomyces cerevisiae coculture platform
for the production of flavonoids. J Agric Food Chem 68(7):2146–2154
Duan L, Ding W, Liu X, Cheng X, Cai J, Hua E, Jiang H (2017) Biosynthesis and engineering of
kaempferol in Saccharomyces cerevisiae. Microb Cell Fact 16(1):1–10
Ebrahimi M, Mokhtari A (2017) Engineering of secondary metabolites in tissue and cell culture of
medicinal plants: an alternative to produce beneficial compounds using bioreactor technologies.
In: Abdullah SNA, Chai-Ling H, Wagstaff C (eds) Crop improvement. Springer, Cham, pp 137–
167
Eichenberger M, Hansson A, Fischer D, Dürr L, Naesby M (2018) De novo biosynthesis of antho-
cyanins in Saccharomyces cerevisiae. FEMS Yeast Res 18(4):1–13
Fang Z, Jones JA, Zhou J, Koffas MAG (2018) Engineering Escherichia coli co-cultures for
production of curcuminoids from glucose. Biotechnol J 1700576(13):1–8
Fraser CM, Chapple C (2011) The phenylpropanoid pathway in arabidopsis. The Arabidopsis
Book, vol 9. p e0152
Ghofrani S, Joghataei MT, Mohseni S, Baluchnejadmojarad T, Bagheri M, Khamse S, Roghani M
(2015) Naringenin improves learning and memory in an Alzheimer’s disease rat model: insights
into the underlying mechanisms. Eur J Pharmacol 764(2015):195–201
Grand View Research (2016) Flavonoids market size worth $1.06 billion by 2025. https://www.gra
ndviewresearch.com/press-release/global-flavonoids-market
Grand View Research (2020) Curcumin Market Size Worth $151.9 Million By 2027. https://www.
grandviewresearch.com/press-release/curcumin-market
Gu Y, Ma J, Zhu Y, Ding X, Xu P (2020) Engineering Yarrowia lipolytica as a chassis for de novo
synthesis of five aromatic-derived natural products and chemicals. ACS Synth Biol 9(8):2096–
2106
Gudiña EJ, Amorim C, Braga A, Costa Â, Rodrigues JL, Rodrigues SS, Rodrigues LR (2020)
Biotech green approaches to unravel the potential of residues into valuable products. In: Ina-
muddin (ed) Green chemistry for the sustainable development of chemical industry. Springer,
pp 97–150
Hashemzaei M, Far AD, Yari A, Heravi RE, Tabrizian K, Taghdisi SM, Sadegh SE, Tsarouhas K,
Kouretas D (2017) Anticancer and apoptosis—inducing effects of quercetin in vitro and in vivo.
Oncol Rep 38:819–828
He Q, Szczepańska P, Yuzbashev T, Lazar Z, Ledesma-Amaro R (2020) De novo production of
resveratrol from glycerol by engineering different metabolic pathways in Yarrowia lipolytica.
Metab Eng Commun 11:0–5
Hewlings S, Kalman D (2017) Curcumin: a review of its effects on human health. Foods 6(10):92
Horne JR, Vohl M-C (2020) Biological plausibility for interactions between dietary fat, resveratrol,
ACE2, and SARS-CoV illness severity. Am J Physiol Endocrinol Metab 318:830–833
Huang YC, Chen YF, Chen CY, Chen WL, Ciou YP, Liu WH, Yang CH (2011) Production of
ferulic acid from lignocellulolytic agricultural biomass by Thermobifida fusca thermostable
esterase produced in Yarrowia lipolytica transformant. Biores Technol 102(17):8117–8122
Huang ZQ, Chen P, Su WW, Wang YG, Wu H, Peng W, Li PB (2018) Antioxidant activity and
hepatoprotective potential of quercetin 7-rhamnoside in vitro and in vivo. Molecules 23(5):1–
13
Yeast Synthetic Biology Approaches for the Production of Valuable … 151
Huang LL, Xue Z, Zhu QQ (2006) Method for the production of resveratrol in a recombinant
oleaginous microorganism (Patent No. PCT/US2006/019085)
Hussain T, Tan B, Yin Y, Blachier F, Tossou MCB, Rahu N (2016) Oxidative stress and inflamma-
tion: what polyphenols can do for us? Oxidative Med Cell Longevity 2016
Ibrahim NI, Wong SK, Mohamed IN, Mohamed N, Chin KY, Ima-Nirwana S, Shuid AN (2018)
Wound healing properties of selected natural products. Int J Environ Res Public Health
15(11):1–23
Jain PK, Joshi H (2012) Coumarin: chemical and pharmacological profile. J Appl Pharm Sci
2(6):236–240
Jasso-Miranda C, Herrera-Camacho I, Flores-Mendoza LK, Dominguez F, Vallejo-Ruiz V,
Sanchez-Burgos GG, Pando-Robles V, Santos-Lopez G, Reyes-Leyva J (2019) Antiviral
and immunomodulatory effects of polyphenols on macrophages infected with dengue virus
serotypes 2 and 3 enhanced or not with antibodies. Infect Drug Resist 2019(12):1833–1852
Jiang H, Wood KV, Morgan JA (2005) Metabolic engineering of the phenylpropanoid pathway in
Saccharomyces cerevisiae. Appl Environ Microbiol 71(6):2962–2969
Jin L, Zeng W, Zhang F, Zhang C, Liang W (2017) Naringenin ameliorates acute inflammation by
regulating intracellular cytokine degradation. J Immunol 199(10):3466–3477
Kamboj A, Saluja AK, Kumar M, Atri P (2012) Antiviral activity of plant polyphenols. J Pharm
Res 5(5):2402–2412
Karbalaei M, Rezaee SA, Farsiani H (2020) Pichia pastoris: a highly successful expression system
for optimal synthesis of heterologous proteins. J Cell Physiol 235(9):5867–5881
Katsuyama Y, Matsuzawa M, Funa N, Horinouchi S (2008) Production of curcuminoids by
Escherichia coli carrying an artificial biosynthesis pathway. Microbiology 154(9):2620–2628
Katsuyama Y, Kita T, Horinouchi S (2009) Identification and characterization of multiple curcumin
synthases from the herb Curcuma longa. FEBS Lett 583(17):2799–2803
Katsuyama Y, Hirose Y, Funa N, Ohnishi Y, Horinouchi S (2010) precursor-directed biosynthesis
of curcumin analogs in Escherichia coli. Biosci Biotechnol Biochem 74(3):641–645
Khan H, Reale M, Ullah H, Sureda A, Tejada S, Wang Y, Zhang ZJ, Xiao J (2020) Anti-cancer
effects of polyphenols via targeting p53 signaling pathway: updates and future directions.
Biotechnol Advan 38:107385
Kim EJ, Cha MN, Kim BG, Ahn JH (2017) Production of curcuminoids in engineered Escherichia
coli. J Microbiol Biotechnol 27(5):975–982
Koopman F, Beekwilder J, Crimi B, van Houwelingen A, Hall RD, Bosch D, van Maris AJA,
Pronk JT, Daran JM (2012) De novo production of the flavonoid naringenin in engineered
Saccharomyces cerevisiae. Microb Cell Fact 11:1–15
Krivoruchko A, Nielsen J (2015) Production of natural products through metabolic engineering of
Saccharomyces cerevisiae. Curr Opin Biotechnol 35:7–15
Lee S, Lee HH, Shin YS, Kang H, Cho H (2017) The anti-HSV-1 effect of quercetin is dependent
on the suppression of TLR-3 in Raw 264.7 cells. Arch Pharm Res 40(5):623–630
Leema G, Tamizhselvi R (2018) Protective effect of scopoletin against cerulein-induced acute
pancreatitis and associated lung injury in mice. Pancreas 47(5):577–585
Li M, Kildegaard KR, Chen Y, Rodriguez A, Borodina I, Nielsen J (2015) De novo production
of resveratrol from glucose or ethanol by engineered Saccharomyces cerevisiae. Metab Eng
32(2015):1–11
Li M, Schneider K, Kristensen M, Borodina I, Nielsen J (2016) Engineering yeast for high-level
production of stilbenoid antioxidants. Sci Rep 6(36827):1–8
Lin Y, Sun X, Yuan Q, Yan Y (2013) Combinatorial biosynthesis of plant-specific coumarins in
bacteria. Metab Eng 18(2013):69–77
Liou G, Chiang YC, Wang Y, Weng JK (2018) Mechanistic basis for the evolution of chalcone
synthase catalytic cysteine reactivity in land plants. J Biol Chem 293(48):18601–18612
Liu X, Ding W, Jiang H (2017) Engineering microbial cell factories for the production of plant nat-
ural products: from design principles to industrial-scale production. Microb Cell Fact 16(1):1–9
152 D. Gomes et al.
Liu L, Liu H, Zhang W, Yao M, Li B, Liu D, Yuan Y (2019a) Engineering the biosynthesis of
caffeic acid in saccharomyces cerevisiae with heterologous enzyme combinations. Engineering
5(2):287–295
Liu Q, Yu T, Li X, Chen Y, Campbell K, Nielsen J, Chen Y (2019b) Rewiring carbon metabolism
in yeast for high level production of aromatic chemicals. Nat Commun 10(1):1–13
Lv Y, Edwards H, Zhou J, Xu P (2019a) Combining 26s rDNA and the Cre-loxP system for iter-
ative gene integration and efficient marker curation in Yarrowia lipolytica. ACS Synth Biol
8(3):568–576
Lv Y, Marsafari M, Zhou J, Xu P (2019b) Optimizing oleaginous yeast cell factories for flavonoids
and hydroxylated flavonoids biosynthesis. ACS Synth Biol 8:2514–2523
Lyu X, Ng KR, Lee JL, Mark R, Chen WN (2017) Enhancement of naringenin biosynthesis
from tyrosine by metabolic engineering of Saccharomyces cerevisiae. J Agric Food Chem
65(31):6638–6646
Lyu X, Zhao G, Ng KR, Mark R, Chen WN (2019) Metabolic engineering of Saccharomyces
cerevisiae for de Novo production of kaempferol. J Agric Food Chem 67(19):5596–5606
Marchiosi R, dos Santos WD, Constantin RP, de Lima RB, Soares AR, Finger-Teixeira A, Mota
TR, de Oliveira DM, Foletto-Felipe MP, Abrahão J, Ferrarese-Filho O (2020) Biosynthesis and
metabolic actions of simple phenolic acids in plants. Phytochem Rev 19(4):865–906
Market Watch (2020) Global trans resveratrol market 2020—indepth analysis of current market
trends including industry share, size, manufacturers and future prospects. https://www.market
watch.com/press-release/global-trans-resveratrol-market-2020—sindepth-analysis-of-current-
market-trends-including-industry-share-size-manufacturers-and-future-prospects-2020-09-22?
tesla=y
Martillanes S, Rocha-Pimienta J, Cabrera-Bañegil M, Martín-Vertedor D, Delgado-Adámez J
(2017) Application of phenolic compounds for food preservation: food additive and active
packaging. In: Soto-Hernández M (ed) Phenolic compounds—biological activity. InTechOpen,
pp 39–56
Matos MJ, Santana L, Uriarte E, Abreu OA, Molina E, Yordi EG (2015) Coumarins—an important
class of phytochemicals. Rao V (eds) Phytochemicals—isolation, characterisation and role in
human health. InTechOpen, pp 114–140
Meydani M (2009) Potential health benefits of avenanthramides of oats. Nutr Rev 67(12):731–735
Milke L, Aschenbrenner J, Marienhagen J, Kallscheuer N (2018) Production of plant-derived
polyphenols in microorganisms: current state and perspectives. Appl Microbiol Biotechnol
102(4):1575–1585
Moglia A, Goitre L, Gianoglio S, Baldini E, Trapani E, Genre A, Scattina A, Dondo G, Trabalzini
L, Beekwilder J, Retta SF (2015) Evaluation of the bioactive properties of avenanthramide
analogs produced in recombinant yeast. BioFactors 41(1):15–27
Munack S, Roderer K, Ökvist M, Kamarauskaite J, Sasso S, Van Eerde A, Kast P, Krengel U (2016)
Remote control by inter-enzyme allostery: a novel paradigm for regulation of the shikimate
pathway. J Mol Biol 428(6):1237–1255
Musarra-Pizzo M, Ginestra G, Smeriglio A, Pennisi R, Sciortino MT, Mandalari G (2019) The
antimicrobial and antiviral activity of polyphenols from almond (Prunus dulcis L.) skin. Nutri-
ents 11(10):1–11
Niedzwiecki A, Roomi MW, Kalinovsky T, Rath M (2016) Anticancer efficacy of polyphenols and
their combinations. Nutrients 8(9):1–17
Palmer CM, Miller KK, Nguyen A, Alper HS (2020) Engineering 4-coumaroyl-CoA derived
polyketide production in Yarrowia lipolytica through a β-oxidation mediated strategy. Metab
Eng 57(2020):174–181
Panche AN, Diwan AD, Chandra SR (2016) Flavonoids: an overview. J Nutr Sci 5(47):1–15
Panda AK, Chakraborty D, Sarkar I, Khan T, Sa G (2017) New insights into therapeutic activity
and anticancer properties of curcumin. J Exp Pharmacol 9:31–45
Parage C, Tavares R, Réty S, Baltenweck-Guyot R, Poutaraud A, Renault L, Heintz D, Lugan R,
Marais GAB, Aubourg S, Hugueney P (2012) Structural, functional, and evolutionary analysis
Yeast Synthetic Biology Approaches for the Production of Valuable … 153
of the unusually large stilbene synthase gene family in grapevine. Plant Physiol 160(3):1407–
1419
Paraiso IL, Revel JS, Stevens JF (2020) Potential use of polyphenols in the battle against COVID-
19. Curr Opin Food Sci 32:149–155
Park S (2015) Polyphenol compound as a transcription factor inhibitor. Nutrients 7(11):8987–9004
Parthasarathy A, Cross PJ, Dobson RCJ, Adams LE, Savka MA, Hudson AO (2018) A three-ring
circus: metabolism of the three proteogenic aromatic amino acids and their role in the health
of plants and animals. Front Mol Biosci 5:1–30
Pecyna P, Wargula J, Murias M, Kucinska M (2020) More than resveratrol: new insights into
stilbene-based compounds. Biomolecules 10(8):1–40
Perrelli A, Goitre L, Salzano AM, Moglia A, Scaloni A, Retta SF (2018) Biological activities,
health benefits, and therapeutic properties of avenanthramides: from skin protection to preven-
tion and treatment of cerebrovascular diseases. Oxid Med Cell Longev 2018(6015351):1–17
Proficient Market (2020). Global polyphenols market will register a CAGR of around 9.0% by
2027. https://proficientmarket.com/press-release/1225/global-polyphenols-market
Rainha J, Gomes D, Rodrigues LR, Rodrigues JL (2020a) Synthetic biology approaches to engi-
neer Saccharomyces cerevisiae towards the industrial production of valuable polyphenolic
compounds. Life 10(56):1–26
Rainha J, Rodrigues JL, Rodrigues LR (2020b) CRISPR-Cas9: a powerful tool to efficiently engi-
neer Saccharomyces cerevisiae. Life 11(13):1–16
Rashmi R, Magesh SB, Ramkumar KM, Suryanarayanan S, SubbaRao MV (2017) Antioxidant
potential of naringenin helps to protect liver tissue from streptozotocin-induced damage. R
Biochem Mol Biol 7(1):76–84
Rivière C, Pawlus AD, Mérillon JM (2012) Natural stilbenoids: distribution in the plant kingdom
and chemotaxonomic interest in Vitaceae. Nat Prod Rep 29(11):1317–1333
Rodrigues JL, Rodrigues LR (2017) Synthetic biology: perspectives in industrial biotechnology.
In: Pandey A, Teixeira J (eds) Current developments in biotechnology and bioengineering:
foundations of biotechnology and bioengineering. Elsevier, pp 239–269
Rodrigues JL, Prather KLJ, Kluskens LD, Rodrigues LR (2015a) Heterologous production of
curcuminoids. Microbiol Mol Biol Rev 79(1):39–60
Rodrigues JL, Araújo RG, Prather KLJ, Kluskens LD (2015b) Production of curcuminoids from
tyrosine by a metabolically engineered Escherichia coli using caffeic acid as an intermediate.
Biotechnol J 10(4):1–27
Rodrigues JL, Araújo RG, Prather KLJ, Kluskens LD, Rodrigues LR (2015c) Heterologous pro-
duction of caffeic acid from tyrosine in Escherichia coli. Enzyme Microb Technol 71:36–44
Rodrigues JL, Couto MR, Araújo RG, Prather KLJ, Kluskens L, Rodrigues LR (2017a) Hydrox-
ycinnamic acids and curcumin production in engineered Escherichia coli using heat shock
promoters. Biochem Eng J 125:41–49
Rodrigues JL, Ferreira D, Rodrigues LR (2017b) Synthetic biology strategies towards the devel-
opment of new bioinspired technologies for medical applications. In: Rodrigues LR, Mota M
(eds) Bioinspired materials for medical applications. Elsevier, pp 451–497
Rodrigues JL, Gomes D, Rodrigues LR (2020) A combinatorial approach to optimize the produc-
tion of curcuminoids from tyrosine in Escherichia coli. Front Bioeng Biotechnol 8(59):1–15
Rodrigues JL, Rodrigues LR (2020) Biosynthesis and heterologous production of
furanocoumarins: perspectives and current challenges. Natural Product Rep
Rodriguez A, Kildegaard KR, Li M, Borodina I, Nielsen J (2015) Establishment of a yeast platform
strain for production of p-coumaric acid through metabolic engineering of aromatic amino acid
biosynthesis. Metab Eng 31(2015):181–188
Rodriguez A, Chen Y, Khoomrung S, Özdemir E, Borodina I, Nielsen J (2017a) Comparison of
the metabolic response to over-production of p-coumaric acid in two yeast strains. Metab Eng
44(2017):265–272
Rodriguez A, Strucko T, Stahlhut SG, Kristensen M, Svenssen DK, Forster J, Nielsen J, Borod-
ina I (2017b) Metabolic engineering of yeast for fermentative production of flavonoids. Biores
Technol 245:1645–1654
154 D. Gomes et al.
Rüfer CE, Kulling SE (2006) Antioxidant activity of isoflavones and their major metabolites using
different in vitro assays. J Agric Food Chem 54(8):2926–2931
Sáez-Sáez J, Wang G, Marella ER, Sudarsan S, Cernuda Pastor M, Borodina I (2020) Engineer-
ing the oleaginous yeast Yarrowia lipolytica for high-level resveratrol production. Metab Eng
62(2020):51–61
Santos CNS, Koffas M, Stephanopoulos G (2011) Optimization of a heterologous pathway for the
production of flavonoids from glucose. Metab Eng 13(4):392–400
Sharma A, Shahzad B, Rehman A, Bhardwaj R, Landi M, Zheng B (2019) Response of phenyl-
propanoid pathway and the role of polyphenols in plants under abiotic stress. Molecules
24(13):1–22
Shen T, Zie C-F, Wang X-N, Lou H-X (2013) Stilbenoids. In: Ramawat KG, Mérillon J-M (eds)
Natural products: phytochemistry, botany and metabolism of alkaloids, phenolics and terpenes.
Springer, pp 1902–1948
Shin SY, Han NS, Park YC, Kim MD, Seo JH (2011) Production of resveratrol from p-coumaric
acid in recombinant Saccharomyces cerevisiae expressing 4-coumarate:coenzyme A ligase and
stilbene synthase genes. Enzyme Microb Technol 48(1):48–53
Shin SY, Jung SM, Kim MD, Han NS, Seo JH (2012) Production of resveratrol from tyrosine in
metabolically engineered Saccharomyces cerevisiae. Enzyme Microb Technol 51(4):211–216
Siddiqui FA, Prakasam G, Chattopadhyay S, Rehman AU, Padder RA, Ansari MA, Irshad R, Man-
galhara K, Bamezai RNK, Husain M, Ali SM, Iqbal MA (2018) Curcumin decreases warburg
effect in cancer cells by down-regulating pyruvate kinase M2 via mTOR-HIF1α inhibition. Sci
Rep 8(1):2–10
Singla RK, Dubey AK, Garg A, Sharma RK, Fiorino M, Ameen SM, Haddad MA, Al-Hiary M
(2019) Natural polyphenols: chemical classification, definition of classes, subcategories, and
structures. J AOAC Int 102(5):1397–1400
Sobhani M, Farzaei MH, Kiani S, Khodarahmi R (2020) Immunomodulatory; anti-
inflammatory/antioxidant effects of polyphenols: A comparative review on the parental
compounds and their metabolites. Food Rev Intl 00(00):1–53
Sökmen M, Akram Khan M (2016) The antioxidant activity of some curcuminoids and chalcones.
Inflammopharmacology 24(2–3):81–86
Soto-Urquieta MG, López-Briones S, Pérez-Vázquez V, Saavedra-Molina A, González-Hernández
GA, Ramírez-Emiliano J (2014) Curcumin restores mitochondrial functions and decreases lipid
peroxidation in liver and kidneys of diabetic db/db mice. Biol Res 47(74):1–8
Stringlis IA, De Jonge R, Pieterse CMJ (2019) The age of coumarins in plant-microbe interactions.
Plant Cell Physiol 60(7):1405–1419
Sydor T, Schaffer S, Boles E (2010) Considerable increase in resveratrol production by recombi-
nant industrial yeast strains with use of rich medium. Appl Environ Microbiol 76(10):3361–
3363
Tanase C, Bujor O-C, Popa VI (2019) Phenolic natural compounds and their influence on physi-
ological processes in plants. In: Watson RR (ed) Polyphenols in plants, 2nd edn. Elsevier Inc,
pp 45–58
Taofiq O, González-Paramás AM, Barreiro MF, Ferreira ICFR, McPhee DJ (2017) Hydroxycin-
namic acids and their derivatives: cosmeceutical significance, challenges and future perspec-
tives, a review. Molecules 22(2):1–24
Trantas E, Panopoulos N, Ververidis F (2009) Metabolic engineering of the complete pathway
leading to heterologous biosynthesis of various flavonoids and stilbenoids in Saccharomyces
cerevisiae. Metab Eng 11(6):355–366
Tsao R (2010) Chemistry and biochemistry of dietary polyphenols. Nutrients 2(12):1231–1246
Tzin V, Galili G (2010) New Insights into the shikimate and aromatic amino acids biosynthesis
pathways in plants. Mol Plant 3(6):956–972
Van de Velde F, Esposito D, Grace MH, Pirovani ME, Lila MA (2019) Anti-inflammatory and
wound healing properties of polyphenolic extracts from strawberry and blackberry fruits. Food
Res Int 121(2019):453–462
Yeast Synthetic Biology Approaches for the Production of Valuable … 155
Zhao Y, Jian X, Wu J, Huang W, Huang C, Luo J, Kong L (2019) Elucidation of the biosynthesis
pathway and heterologous construction of a sustainable route for producing umbelliferone. J
Biol Eng 13(1):1–13
Zhao CH, Zhang RK, Qiao B, Li BZ, Yuan YJ (2020) Engineering budding yeast for the production
of coumarins from lignin. Biochem Eng J 160(May):107634
Zeng YH, Zhou LY, Chen QZ, Li Y, Shao Y, Ren WY, Liao YP, Wang H, Zhu JH, Huang M, He F,
Wang J, Wu K, He BC (2017) Resveratrol inactivates PI3K/Akt signaling through upregulating
BMP7 in human colon cancer cells. Oncol Rep 38(1):456–464
Yeast Synthetic Biology
for Production of Artemisinin
as an Antimalarial Drug
Abstract
Artemisinin, a sesquiterpene endoperoxide lactone derived from Artemisia
annua, is highly effective against malaria parasite Plasmodium falciparum.
Artemisinin and its derivatives (collectively termed artemisinins) demonstrate
additional anticancer, anti-inflammation, antiviral, and anti-SARS-CoV-2 activ-
ity. Because of the expensive medicine of artemisinins and the low content
of artemisinin biosynthesis in native host, researchers have been endeav-
ored to produce artemisinin via alternative approaches. Previous attempts give
attention to increase artemisinin biosynthesis via common plant breeding tech-
niques and on engineering cultivated plants for wide production. However,
current trends are focusing on the bioengineering of artemisinin biosynthetic
pathway in heterologous expression platforms. To date, in planta artemisinin
production in two model plants, tobacco (Nicotiana benthamiana) and moss
(Physcomitrella patens), has been reported successfully. Nevertheless, to meet
the large-scale demands, de novo reconstituting of artemisinin biosynthetic
pathway in heterologous microorganisms such as Escherichia coli and Sac-
charomyces cerevisiae has had documentary achievements. Since there are
challenges for the expression of cytochrome P450s (CYPs) from the eukary-
otic origin (such as CYP71AV1) in E. coli, S. cerevisiae has been proposed as a
heterologous host for production of artemisinin precursors such as amorphadi-
ene, artemisinic acid, and dihydro artemisinic acid. Moreover, the presence of
the MVA biosynthetic pathway for production of universal sesquiterpenes pre-
cursor, farnesyl pyrophosphate (FPP), and relative compatibility of S. cerevisiae
cell environment for the expression of plant-origin genes (enzymes) have made
this microorganism as a favorable platform for artemisinins production.
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 157
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_6
158 A. Beyraghdar Kashkooli et al.
1 Introduction
Terpenoids (isoprenoids) as the oldest and most diverse group of natural molecules
represent the first instance of natural bioactive compounds (Pateraki et al. 2015).
Most terpenes are isolated from plants which play an important role in some plant’s
metabolism (Davis and Croteau 2000). In plants, terpenes are involved in the
biosynthesis of the phytohormones gibberellic acid, abscisic acid, and cytokinins
and serve as precursors of steroid hormones. They also act as biosynthetic precur-
sors for carotenoid pigments and phytol side chains in chlorophyll and components
of electron-carrying coenzymes such as quinone, ubiquinone, and plastoquinone.
The terpenes–environment interactions are more important as a chemical defense
Yeast Synthetic Biology for Production of Artemisinin … 159
system against herbivores and pathogens, but they also have other roles such as
absorption of pollinators and allopathic properties. Apart from their potential roles
in plants, bioactive terpenoids have raised considerable interest as pharmaceuticals,
nutraceuticals, and flavors and fragrances (Pateraki et al. 2015).
Structurally, terpenoids are various polymers created from the fundamental C5
block, the isoprene unit. They are organized and grouped based on the number of
isoprene units. Therefore, hemiterpenes, monoterpenes, sesquiterpenes, triterpenes,
diterpenes, and carotenoids contain 5, 10, 15, 20, 30, and 40 isoprene units, respec-
tively. The long-chain polymers possess thousands of units and comprise natural
products such as rubber. Regardless of their diversity, all terpenoids are biosynthe-
sized from the universal precursors dimethylallyl pyrophosphate (DMAPP) and
isopentenyl pyrophosphate (IPP, isopentenyl diphosphate, or IDP). This occurs
via two different pathways, the methyl erythritol 4-phosphate (MEP; mevalonate-
independent, or non-mevalonate) pathway and the mevalonate pathway (MVA)
(Capell and Christou 2004).
Biosynthesis of terpenes can be divided into four stages. The first step involves
the production of isopentenyl diphosphate (IPP) and its isomer, dimethylal-
lyl pyrophosphate (DMAPP), which is carried out by isopentenyl diphosphate
isomerase (IPP isomerase). In the second stage, IPP and DMAPP are struc-
turally elongated to form more complex isoprenoid backbones called geranyl
pyrophosphate (GPP C10), farnesyl pyrophosphate (FPP, C15), geranylgeranyl
pyrophosphate (GGPP, C20). The third stage involves the conversion of GPP,
FPP, and GGPP to the corresponding terpene groups producing monoterpenes,
sesquiterpenes, and diterpenes, respectively. The final step in the biosynthesis of
terpenes occurs in the cytosol (e.g., cyclization, hydroxylation, and carboxylation)
which produces derivatives of the basic terpene groups (Davis and Croteau 2000;
Moses et al. 2013).
In general, the major pathway for the biosynthesis of monoterpenes is within
plastids. Monoterpenes can also be metabolically engineered to be produced in the
cytosol, although these pathways can also produce intermediates (Bouwmeester
2006). In plants, two distinct but interacting pathways have been identified for IPP
biosynthesis: the mevalonate (MVA) or cytosolic pathway, the methylerythritol
4-phosphate (MEP) pathway or the plastid pathway.
In parallel to the MEP pathway, the mevalonate pathway (MVA) is one of the
most important cellular metabolic pathways in all eukaryotes and many bacte-
ria. This pathway aims to produce dimethylallyl diphosphate (DMAPP) as well
as isopentenyl diphosphate (IPP), both of which act as a basis for molecular
biosynthesis in a variety of pathways, including the synthesis of terpenoids, cell
membrane preservation and regeneration, and many other cellular hormones. In the
MVA pathway, two molecules of acetyl-CoA first combine to form acetoacetyl-
CoA. Then, by adding another molecule of acetyl-CoA to acetoacetyl-CoA, and
then by performing a series of other steps, mevalonate is obtained. Mevalonate
is then converted to IPP under the influence of other steps and enzymes, and
IPP is converted to DMAPP under the influence of isopentenyl diphosphate iso-
merase. IPP is elongated by the enzyme prenyltransferase, and geranyl diphosphate
is produced by the enzyme geranyl diphosphate synthase, which is also called
the precursor (monoterpene synthases) produces different types of monoterpenes
(Vranova et al. 2012; Moses et al. 2013) (Fig. 1).
It is important to deeply understand the key biosynthetic pathway routes and genes
involved in artemisinin biosynthesis in order to be able to enhance artemisinin
yield in the natural artemisinin producing plant, A. annua, as well as other heterol-
ogous host platforms. artemisinin is a sesquiterpene lactone, containing 15 carbon
atoms with lipophilic properties (Lv et al. 2017). Both artemisinin and arteannuin
B are produced via a biosynthetic pathway in the glandular trichomes (GTs) in
leaves and oval of the A. annua (Wang et al. 2016). However, some reports have
shown that the highest GTs densities are detected on flowers, buds, and young
leaves of this plant (Olofsson et al. 2011). The biosynthetic pathway of GTs occurs
via the terpenoids biosynthetic pathway. Isoprene units of terpenoids are obtained
from two pathways: the MVA pathway, which occurs in the cytosol, MEP pathway
in the plastids (Dewick 2009; Vranová et al. 2013). The artemisinin biosynthetic
pathway starts with the conversion of 5-carbon units of IPP and MAPP units to FPP
(Fuentes et al. 2016). A study on the artemisinin biosynthetic pathway shows that
the first and key step in biosynthesis of this compound is the conversion of FPP to
amorpha 4, 11-diene (amorphadiene), which is catalyzed by a well-known terpene
cyclase, the amorphadiene synthase (ADS) (Bertea et al. 2005; Mercke et al. 2000).
The cytochrome P450 hydroxylase (CYP71AV1) (Teoh et al. 2006) then converts
amorphadiene to artemisinic alcohol. CYP71AV1 also oxidizes artemisinic alcohol
to artemisinic aldehyde and artemisinic acid, respectively. Artemisinic aldehyde
reductase (DBR2) and aldehyde dehydrogenase 1 (ALDH1) (Zhang et al. 2008),
as the branching point, converts artemisinic aldehyde to dihydroartemisinic alde-
hyde (Bertea et al. 2005; Schramek et al. 2010). Bifurcation points of artemisinic
aldehyde biosynthetic pathway can synthesize artemisinic acid by CYP71AV1
and ALDH1. Also, due to the non-enzymatic reaction, dihydroartemisinic acid
and artemisinic acid are converted to artemisinin and arteannuin B, respectively
(Brown 2010; Han et al. 2014; Ikram and Simonsen 2017). ALDH1 then pro-
duces dihydroartemisinic acid (DHAA) from dihydroartemisinic aldehyde. DHAA
is considered as a precursor to the artemisinin (bio)synthesis (Tang et al. 2014),
and finally its conversion to artemisinin is considered as a non-enzymatic reaction
which occurs under the influence of UV light and oxygen (Bertea et al. 2005;
Ikram et al. 2019; Schramek et al. 2010; Teoh et al. 2006) (Fig. 2).
Studies show that all genes responsible for biosynthesis of artemisinin are
fully identified (Ikram et al. 2019). Identification of the genes involved in the
biosynthetic pathway of artemisinin offers an opportunity to increase artemisinin
production and subtle reduction in its production costs through plant breeding
and/or the engineering of other putative hosts (Zhang et al. 2008).
Yeast Synthetic Biology for Production of Artemisinin … 163
4.1 Antimalarial
Malaria has been a worldwide outbreak health threat since old times. Scientific and
hospital reports show that many patients die every year due to malaria infections.
Artemisinins and artemisinin-based combination therapies (ACTs) are frontline
antimalarial agents accepted for their potency and low toxicity (Yang et al. 2020).
In the early 1970s, a group of Chinese scientists led by Professor Tu Youyou
(winner of Nobel prize in physiology or medicine 2015) succeeded in obtaining
the effective results of A. annua plant extracts on malaria (Liu and Liu 2016;
Su and Miller 2015). By testing the natural extract from the A. annua, Professor
Tu found that the compound could have 100% inhibitory effect on malaria para-
sites in vivo experiments (Tu 2011). In the late 1970s, artemisinin was effectively
recognized as an anti-malarial drug (Klayman 1985; Tu 2011; Xie et al. 2016).
So far, several mechanisms of action such as interfering with the heme detoxifica-
tion pathway, creating protein damage and inhibiting parasite proteasome function,
inducing the alkylation of PfTCTP and inhibiting PfATPase, and membrane depo-
larizing of mitochondria via ROS production have been reported for artemisinin to
fight malaria parasites (Plasmodium falciparum and Plasmodium vivax).
4.2 Anti-inflammation
Artemisinin prevents the secretion and the mRNA amount of TNF-α, interleukin
(IL)-1ß, and IL-6 in a dose-dependent manner in PMA-induced THP-1 human
monocytes. It also inhibits the phosphorylation of IKKα/ß, phosphorylation and
degradation of IκBα, and the nuclear translocation of the NF-κB p65 subunit
(Wang et al. 2011).
4.3 Anticancer
Beyond the anti-malarial activity of artemisinins, they have also been proven to
contain anticancer effects, indicating cytotoxic effects in combating numerous
cancer types. These activities look to be done by artemisinin-induced modifica-
tions in several signaling pathways, interfering with various hallmarks of cancer at
the same time. Considerable successes have been obtained to distinguish these
pathways and to display their anticancer mechanisms of action (Wong et al.
2017). Recent investigations have documented the strong anticancer activity of
artemisinins (Slezakova and Ruda-Kucerova 2017). The main mechanisms of
action are reported for artemisinin in renal (Yu et al. 2019), breast (Cao et al.
2019; Li et al. 2019), ovarian (Liang et al. 2019), prostate (Willoughby et al.
2009), colon (Riganti et al. 2009), cervical (Mondal and Chatterji 2015), ishikawa
endometrial (Tran et al. 2014), neuroblastoma (Zhu et al. 2014), nasopharyngeal
(Wu et al. 2011), gastric (Zhang et al. 2014), leukemia (Ohgami et al. 2010),
lung (Li et al. 2018), fibrosarcoma tumors (Singh and Lai 2004), for artesunate
in cancer stem cells (Subedi et al. 2016), prostate (Nunes et al. 2017), head and
neck (Roh et al. 2017), cervical (Thanaketpaisarn et al. 2011), colorectal (Chen
et al. 2017), bladder (Zhao et al. 2020; Zhou et al. 2020), leukemia (Zhou et al.
2007), skin (Jiang et al. 2012), liver (Yin et al. 2020), laryngeal (Singh and Verma
2002), ovarian (McDowell et al. 2021), glioblastoma (Berdelle et al. 2011), rhab-
domyosarcoma (Beccafico et al. 2015), breast (Pirali et al. 2020), endometrial (Yin
et al. 2021), lung (Ma et al. 2011), colon (Hao et al. 2020), non-small cell lung
(Wang et al. 2020), pancreatic (Wang et al. 2019), gastric (Zhang et al. 2015),
beta-cell lymphoma (Våtsveen et al. 2018), and esophageal (Fei et al. 2018), for
dihydroartemisinin in colorectal (Yao et al. 2018), cervical (Disbrow et al. 2005),
esophageal (Li et al. 2018a, b), breast (Mao et al. 2013), hepatocellular (Wu et al.
2019), ovarian (Liu et al. 2018), cholangiocarcinoma and hepatocarcinoma (Chai-
jaroenkul et al. 2011), pancreatic (Chen et al. 2009), non-small-cell lung (Jiang
et al. 2016), for artemisone in melanoma, breast, colon, and pancreatic (Gravett
et al. 2011; Dwivedi et al. 2015), and for artemether in gastric (Alcântara et al.
2013) cancer cells.
Yeast Synthetic Biology for Production of Artemisinin … 165
steps which the pathway starts with the acetyl-CoA precursor. Also, newly rec-
ognized genes, such as cytochrome P450 and its related reductase, have been
indicated to catalyze several phases in the artemisinin biochemical network. This
has the promising results to create a semi-synthetic strategy to production that is
both feasible and profitable. Artemisinin precursors de novo biosynthesis in engi-
neered yeast is about many orders of the content above field-grown A. annua plant
(Arsenault et al. 2008). Because cytochromes P450 enzymes (CYPs) from eukary-
otic origin (such as CYP71AV1) cannot be perfectly expressed in E. coli, yeast was
selected as a heterologous host for artemisinin precursor production (Paddon and
Keasling 2014). In addition, this platform is more methodically and technically
strong and less sensitive to phage contamination (Kong et al. 2013; Krivoruchko
and Nielsen 2015; Perassolo et al., 2018). Attempts to raise flux by engineered net-
works are growing in yeast via combinations of engineering precursors networks
and downstream improving gene expression (Arsenault et al. 2008).
Yeast is a single-celled fungi and one of the best-characterized eukaryotic
organisms. Many cost-effective and beneficial varieties of yeast are available.
Yeasts contain considerable functions in the fermentation of many food prod-
ucts such as bread and cheese. (Tamang and Fleet 2009). S. cerevisiae is one
of the most studied eukaryotic platform organisms. Availability of a high den-
sity of elucidated information regarding its physiology, genetics, and biochemistry
offers a significant improvement in the industrial application of this microorgan-
ism. The yeast model gives numerous similar advantages as prokaryotic platforms
performed for terpene hydrocarbon biosynthesis, which also supplies the biosyn-
thetic tools necessary for the appropriate functional expression of the downstream
modifying enzymes such as cytochrome P450 hydroxylases (Kong et al. 2013). S.
cerevisiae has been extensively used in biotechnology due to its generally regarded
as safe (GRAS) is proper for large-scale performance. As a eukaryotic platform,
cell and molecular biology of S. cerevisiae have been investigated comprehensively
with many genetic engineering existing techniques. Additionally, S. cerevisiae dis-
plays high tolerance against inappropriate industrial conditions (Hong and Nielsen
2012). Thus, S. cerevisiae has been introduced as a model microorganism for syn-
thetic biology (Lian et al. 2018). Therefore, de novo production of the artemisinin
via metabolic engineering and SynBio in S. cerevisiae has relied on the com-
plete knowledge of this organism. In this regard, an effective effort to produce
semi-synthetic artemisinin was made in collaboration with the University of Cali-
fornia, Amyris, and OneWorld Health, with the primary goal of host engineering
to produce high content of artemisinin precursors that could be converted to the
artemisinin by chemical processes (Paddon and Keasling 2014). Remarkably, it
is noteworthy to imply that Amyris company is an industrial claimant for the
production of artemisinic acid in the S. cerevisiae platform (Tang et al. 2014).
The efficient application of SynBio for the semi-synthetic industrial production
of artemisinin was reported in S. cerevisiae (Paddon et al. 2013), as one of the
prominent points of synthetic artemisinin biology (Immethun et al. 2013). Here,
we review and compartmentalize the de novo reconstitution and production of
components involved in the artemisinin biosynthetic pathway in S. cerevisiae.
Yeast Synthetic Biology for Production of Artemisinin … 167
hosts (Covello 2008; Ikram et al. 2019; Ma et al. 2015; Xie et al. 2016). Many
efforts have been made to increase the production of artemisinin during various
experiments. However, there are still lots of potential routes to increase the pro-
duction of this valuable compound (Lv et al. 2017). To overcome the unsustainable
production and low-yield of artemisinin in A. annua, attempts led to positive results
in biological systems as native and heterologous expression systems (Paddon et al.
2013). One of the main approaches pursued in metabolic engineering is the sus-
tainable production of artemisinin to benefit underdeveloped and more vulnerable
communities (Ikram et al. 2017).
Insertion of Agrobacterium’s rol genes is considered an effective and efficient
method for increasing the secondary metabolites in plants (Bulgakov 2008). Dil-
shad et al. (2015) indicated that transformation via rol B and rol C genes show high
levels of gene expression and artemisinin production (9 and 4 times more than a
non-transgenic plant, respectively). Also, the expression of the ipt gene transmitted
by A. tumefaciens increases the content of artemisinin (30–70%) (Sa et al. 2001).
As mentioned earlier, ADS, Cytochrome P450 (CYP71AV1), double-bond
reductase 2 (DBR2), and aldehyde dehydrogenase 1 (ALDH1) catalyze the dif-
ferent stages of artemisinin biosynthesis pathway. One of the strategies used in
artemisinin metabolic engineering is to increase the expression of genes involved
in the biosynthetic pathway. 3-hydroxy-3-methylglutaryl-CoA reductase (HMGR)
gene is known to be one of the most important factors in MVA pathway (Farhi
170 A. Beyraghdar Kashkooli et al.
et al. 2013). Different studies have shown that high expression of the HMGR
increases artemisinin production (Aquil et al. 2009; Ma et al. 2017; Nafis et al.
2011). Aquil et al. (2009) by transferring the HMGR gene using the Agrobac-
terium-mediated gene transformation, succeeded in producing transgenic lines
with a 22.5% increase in artemisinin content compared to the non-transgenic A.
annua. Squalene synthase (SQS), as another potent enzyme in artemisinin biosyn-
thesis, converts the FPP to squalene production, thereby reduces the strength of
the biosynthetic pathway toward the artemisinin production. Studies have shown
that suppressing this enzyme can increase artemisinin biosynthesis (Paradise et al.
2008; Yang et al. 2008; Zhang et al. 2009). Zhang et al. (2009), by suppressing
the expression of SQS gene using RNA interference (RNAi) technique, indicated
that the amount of artemisinin in transgenic A. annua increased significantly to
31.4 mg on a dry weight scale. Also, with overexpression of genes involved in
the biosynthetic pathway, the amount of artemisinin increases significantly to 3.6-
fold higher than control in transgenic plants, which includes FPS, CYP71AV1, and
CPR, among the most effective genes in biosynthesis of artemisinin (Chen et al.
2013). Furthermore, in separate studies, the amount of artemisinin in transgenic
plants was 38% higher than normal plants, and it was observed that the overex-
pression of the farnesyl pyrophosphate synthase (FPS) in A. annua significantly
increased compared to the wild-type (Banyai et al. 2010; Han et al. 2006). It
has also been reported that overexpression of the chimeric farnesyl diphosphate
synthase in native plant strengthens the biosynthetic pathway of sesquiterpenes,
leading to increased artemisinin biosynthesis in some lines (up to threefold) (Chen
et al. 2000). In another study Kiani et al. (2016), with overexpression of rol ABC
genes, also in diploid plants and HMGR, FPS, ADS, and Aldh1 and upregulation of
ADS in tetraploid plants of A. annua increased artemisinin level (Lin et al. 2011).
Moreover, transcription factors (TFs) known as regulators of the location and time
of secondary metabolites production (Yang et al. 2012), are one of the modules
that can be used in the regulation of special metabolites production (Verpoorte and
Memelink 2002). Identification of TFs involved in the regulation of artemisinin
metabolism is important (Jiang et al. 2016; Shen et al. 2016). AP2/ERF, bHLH,
MYB, and WRKY are members of various TFs in artemisinin (Lv et al. 2017) that
influence the artemisinin biosynthesis in A. annua.
in the moss system (Anterola et al. 2009; Pan et al. 2015; Reski et al. 2015; Zhan
et al. 2014). Ikram et al. (2017) by engineering and inserting 5 genes involved
in the biosynthetic pathway of artemisinin into P. patens, produced 0.21 mg/g of
dry weight artemisinin after three days. Various reports indicate that the insertion
of genes involved in the production of artemisinin has been successful using a
constitutive expression in P. patens.
7 Conclusions
Given the fact that artemisinins, in addition to treating malaria, also play a key
role in fighting other diseases, meeting global demands which has been a chal-
lenge for researchers and industry. Since the artemisinin levels in Artemisia annua
plant is very low, its large-scale planting does not meet global needs. To this end,
biotechnological tools have been able to eliminate these concerns to a large extent.
In this regard, biosynthetic pathway engineering in the main host has been suc-
cessfully performed. For further production, the engineering of the artemisinin
biosynthetic pathway in two model plants (tobacco and moss) has led to signifi-
cant successes. But for much more production, the reconstitution of the artemisinin
biosynthetic pathway in microorganisms has been able to meet global demand so
that the baker’s yeast with a series of special mechanisms and potentials is more
recommended for this purpose.
References
Alcântara DDFÁ, Ribeiro HF, Cardoso PC, dos S, Araújo TMT, Burbano RR, Guimarães AC,
Khayat AS, De Oliveira Bahia M (2013) In vitro evaluation of the cytotoxic and genotoxic
effects of artemether, an antimalarial drug, in a gastric cancer cell line (PG100). J Appl Toxicol
33:151–156.https://doi.org/10.1002/jat.1734
Alejos-Gonzalez F, Qu G, Zhou LL et al (2011) Characterization of development and artemisinin
biosynthesis in self-pollinated Artemisia annua plants. Planta 234:685–697. https://doi.org/10.
1007/s00425-011-1430-z
Anterola A, Shanle E, Perroud P-F, Quatrano R (2009) Production of taxa-4(5),11(12)-diene by
transgenic Physcomitrella patens. Transgenic Res 18(4):655. https://doi.org/10.1007/s11248-
009-9252-5
Aquil S, Husaini AM, Abdin MZ, Rather GM (2009) Overexpression of the HMG-CoA reductase
gene leads to enhanced artemisinin biosynthesis in transgenic Artemisia annua plants. Planta
Med 75:1453–1458. https://doi.org/10.1055/s-0029-1185775
Arsenault P, Wobbe K, Weathers P (2008) Recent advances in Artemisinin production through het-
erologous expression. Curr Med Chem 15:2886–2896. https://doi.org/10.2174/092986708786
242813
Baadhe RR, Mekala NK, Parcha SR, Prameela Devi Y (2013) Combination of ERG9 repression
and enzyme fusion technology for improved production of amorphadiene in Saccharomyces
cerevisiae. J Anal Methods Chem.https://doi.org/10.1155/2013/140469
Banyai W, Kirdmanee C, Mii M, Supaibulwatana K (2010) Overexpression of farnesyl pyrophos-
phate synthase (FPS) gene affected Artemisinin content and growth of Artemisia annua L. Plant
Cell Tissue Organ Cult 103:255–265. https://doi.org/10.1007/s11240-010-9775-8
Yeast Synthetic Biology for Production of Artemisinin … 173
Davis EM, Croteau R (2000) Cyclization enzymes in the biosynthesis of monoterpenes, sesquiter-
penes, and diterpenes 209:53–95. https://doi.org/10.1007/3-540-48146-x_2
Dewick PM (2009) Medicinal natural products: a biosynthetic approach, 3rd edn. Doi: 10
02/9780470742761.fmatter
Dilshad E, Cusido RM, Palazon J, Estrada KR, Bonfill M, Mirza B (2015) Enhanced artemisinin
yield by expression of rol genes in Artemisia annua. Malar J 14:1–10. https://doi.org/10.1186/
s12936-015-0951-5
Disbrow GL, Baege AC, Kierpiec KA, Yuan H, Centeno JA, Thibodeaux CA, Hartmann D,
Schlegel R (2005) Dihydroartemisinin is cytotoxic to papillomavirus-expressing epithelial cells
in vitro and in vivo. Cancer Res 65:10854–10861. https://doi.org/10.1158/0008-5472.CAN-05-
1216
Donald KA, Hampton RY, Fritz IB (1997) Effects of overproduction of the catalytic domain of
3-hydroxy-3- methylglutaryl coenzyme A reductase on squalene synthesis in Saccharomyces
cerevisiae. Appl Environ Microbiol 63:3341–3344
Dwivedi A, Mazumder A, du Plessis L, du Preez JL, Haynes RK, du Plessis J (2015) In vitro anti-
cancer effects of artemisone nano-vesicular formulations on melanoma cells. Nanomedicine
Nanotechnology. Biol Med 11:2041–2050. https://doi.org/10.1016/j.nano.2015.07.010
Dudareva N, Klempien A, Muhlemann JK, Kaplan I (2013) Biosynthesis, function and metabolic
engineering of plant volatile organic compounds. New Phytol 198:16–32. https://doi.org/10.
1111/nph.12145
Efferth T (2018) Beyond malaria: the inhibition of viruses by artemisinin-type compounds.
Biotechnol Adv 36:1730–1737. https://doi.org/10.1016/j.biotechadv.2018.01.001
Farhi M, Kozin M, Duchin S, Vainstein A (2013) Metabolic engineering of plants for artemisinin
synthesis. Biotechnol Genet Eng Rev 29(2):135–148. https://doi.org/10.1080/02648725.2013.
821283
Farid F, Sideeq O, Khan F, Niaz K (2019) Saccharomyces cerevisiae, nonvitamin and nonmin-
eral nutritional supplements. Elsevier Inc https://www.sciencedirect.com/science/article/pii/
B9780128124918000667?via%3Dihub
Fei Z, Gu W, Xie R, Su H, Jiang Y (2018) Artesunate enhances radiosensitivity of esophageal can-
cer cells by inhibiting the repair of DNA damage. J Pharmacol Sci 138:131–137. https://doi.org/
10.1016/j.jphs.2018.09.011
Fuentes P, Zhou F, Erban A, Karcher D, Kopka J, Bock R (2016) A new synthetic biology approach
allows transfer of an entire metabolic pathway from a medicinal plant to a biomass crop. elife
5:1–26
Gravett AM, Liu WM, Krishna S, Chan WC, Haynes RK, Wilson NL, Dalgleish AG (2011) In vitro
study of the anti-cancer effects of artemisone alone or in combination with other chemothera-
peutic agents. Cancer Chemother Pharmacol 67:569–577. https://doi.org/10.1007/s00280-010-
1355-4
Han JL, Liu BY, Ye HC, Wang H, Li ZQ, Li GF (2006) Effects of overexpression of the endogenous
farnesyl diphosphate synthase on the artemisinin content in Artemisia annua L. J Integr Plant
Biol 48(4):482–487. https://doi.org/10.1111/j.1744-7909.2006.00208.x
Han J, Wang H, Lundgren A, Brodelius PE (2014) Effects of overexpression of AaWRKY1 on
artemisinin biosynthesis in transgenic Artemisia annua plants. Phytochemistry 102:89–96
Han JL, Li ZQ, Ye HC (2008) Molecular cloning, prokaryotic expression, and enzyme activity
assay of fps from Artemisia annua. J Agric Univ Hebei 31:71–75 (in Chinese with English
abstract)
Hao DL, Xie R, De GJ, Yi H, Zang C, Yang MY, Liu L, Ma H, Cai WY, Zhao QH, Sui F, Chen
YJ (2020) PH-responsive artesunate polymer prodrugs with enhanced ablation effect on rodent
xenograft colon cancer. Int J Nanomedicine 15:1771–1786. https://doi.org/10.2147/IJN.S24
2032
Ho WE, Peh HY, Chan TK, Wong WSF (2014) Artemisinins: pharmacological actions beyond anti-
malarial. Pharmacol Ther 142:126–139. https://doi.org/10.1016/j.pharmthera.2013.12.001
Hong KK, Nielsen J (2012) Metabolic engineering of Saccharomyces cerevisiae: a key cell factory
platform for future biorefineries. Cell Mol Life Sci 69:2671–2690
Yeast Synthetic Biology for Production of Artemisinin … 175
Ikram NBK, Zhan X, Pan X-W, King BC, Simonsen HT (2015) Stable heterologous expression
of biologically active terpenoids in green plant cells. Front Plant Sci 6:129. https://www.fronti
ersin.org/article/10.3389/fpls.2015.00129
Ikram NKBK, Simonsen HT (2017) A review of biotechnological artemisinin production in plants.
Front Plant Sci 8:1–10
Ikram NKBK, Beyraghdar Kashkooli A, Peramuna A, van der Krol AR, Bouwmeester H, Simon-
sen HT (2019) Insights into heterologous biosynthesis of arteannuin B and artemisinin in
Physcomitrella patens. Molecules 24:3822
Ikram NKBKh, Beyraghdar Kashkooli A, Peramuna AV, van der Krol AR, Bouwmeester H,
Simonsen HT (2017) Stable production of the antimalarial drug artemisinin in the moss
Physcomitrella patens. Frontiers in Bioengineering and Biotechnology 5:1–8. https://doi.org/
10.3389/fbioe.2017.00047
Immethun CM, Hoynes-O’Connor AG, Balassy A, Moon TS (2013) Microbial production of iso-
prenoids enabled by synthetic biology. Front Microbiol. https://doi.org/10.3389/fmicb.2013.
00075
Jiang J, Geng G, Yu X, Liu H, Gao J, An H, Cai C, Li N, Shen D, Wu X, Zheng L, Mi Y, Yang
S (2016) Repurposing the anti-malarial drug dihydroartemisinin suppresses metastasis of non-
small-cell lung cancer via inhibiting NF-κB/GLUT1 axis. Oncotarget 7:87271–87283. https://
doi.org/10.18632/oncotarget.13536
Jiang Z, Chai J, Chuang HHF, Li S, Wang T, Cheng Y, Chen W, Zhou D (2012) Artesunate induces
G0/G1 cell cycle arrest and iron-mediated mitochondrial apoptosis in A431 human epidermoid
carcinoma cells. Anticancer Drugs 23:606–613. https://doi.org/10.1097/CAD.0b013e328350
e8ac
Judd R, Bagley MC, Li M, Zhu Y, Lei C, Yuzuak S, Ekelöf M, Pu G, Zhao X, Muddiman DC, Xie
DY (2019) Artemisinin biosynthesis in non-glandular trichome cells of Artemisia annua. Mol
Plant 12:704–714. https://doi.org/10.1016/j.molp.2019.02.011
Kiani BH, Suberu J, Mirza B (2016) Cellular engineering of Artemisia annua and Artemisia dubia
with the rol ABC genes for enhanced production of potent anti-malarial drug artemisinin. Malar
J 15:1–17. https://doi.org/10.1186/s12936-016-1312-8
Klayman DL (1985) Qinghaosu (artemisinin): an antimalarial drug from China. Science 31:1049–
1055
Kong J, Cheng K, Wang LN (2007) Increase of copy number of HMG-CoA reductase and FPP
synthase genes improves the amorpha4,11-diene production in engineered yeast. Acta Pharm
Sin 42:1314–1319
Kong J, Yang Y, Wang W, Cheng K, Zhu P (2013) Artemisinic acid: a promising molecule poten-
tially suitable for the semi-synthesis of artemisinin. RSC Adv 3:7622–7641. https://doi.org/10.
1039/c3ra40525g
Krishna S, Augustin Y, Wang J, Xu C, Staines HM, Platteeuw H, Kamarulzaman A, Sall A, Krem-
sner P (2021) Repurposing Antimalarials to tackle the COVID-19 pandemic. Trends Parasitol
37:8–11. https://doi.org/10.1016/j.pt.2020.10.003
Krivoruchko A, Nielsen J (2015) Production of natural products through metabolic engineering
of Saccharomyces cerevisiae. Curr Opin Biotechnol 35:7–15. https://doi.org/10.1016/j.copbio.
2014.12.004
Li C, Li J, Wang G, Li X (2016) Heterologous biosynthesis of artemisinic acid in Saccharomyces
cerevisiae. J Appl Microbiol 120:1466–1478. https://doi.org/10.1111/jam.13044
Li ZQ, Liu Y, Liu BY, Wang H, Ye HC, Li GF (2006) Cloning, E. coli expression and molecular
analysis of amorpha-4,11-diene synthase from a high-yield strain of Artemisia annua L. J Integr
Plant Biol 48:1486–1492. https://doi.org/10.1111/j.1744-7909.2006.00381.x
Li J, Feng W, Lu H, Wei Y, Ma S, Wei L, Liu Q, Zhao J, Wei Q, Yao J (2019) Artemisinin
inhibits breast cancer-induced osteolysis by inhibiting osteoclast formation and breast cancer
cell proliferation. J Cell Physiol 234:12663–12675. https://doi.org/10.1002/jcp.27875
Li X, Gu S, Sun D, Dai H, Chen H, Zhang Z (2018) The selectivity of artemisinin-based drugs on
human lung normal and cancer cells. Environ Toxicol Pharmacol 57:86–94. https://doi.org/10.
1016/j.etap.2017.12.004
176 A. Beyraghdar Kashkooli et al.
Nair MS, Huang Y, Fidock DA, Polyak SJ, Wagoner J, Towler MJ, Weathers PJ (2021) Artemisia
annua L. extracts prevent in vitro replication of SARS-CoV-2. https://doi.org/10.1101/2021.01.
08.425825
Nafis T, Akmal M, Ram M, Alam P, Ahlawat S, Mohd A, Abdin MZ (2011) Enhancement
of artemisinin content by constitutive expression of the HMG-CoA reductase gene in high-
yielding strain of Artemisia annua L. Plant Biotechnol Rep 5:53–60. https://doi.org/10.1007/
s11816-010-0156-x
Nunes JJ, Pandey SK, Yadav A, Goel S, Ateeq B (2017) Targeting NF-kappa B signaling by arte-
sunate restores sensitivity of castrate-resistant prostate cancer cells to antiandrogens. Neoplasia
19:333–345. https://doi.org/10.1016/j.neo.2017.02.002
Ohgami Y, Elstad CA, Chung E, Shirachi DY, Quock RM, Lai HC (2010) Effect of hyperbaric
oxygen on the anticancer effect of artemisinin on molt-4 human leukemia cells. Anticancer Res
30:4467–4470
Olofsson L, Engström A, Lundgren A, Brodelius PE (2011) Relative expression of genes of terpene
metabolism in different tissues of Artemisia annua L. BMC Plant Biol 11:1–12
Pateraki I, Heskes AM, Hamberger B (2015) Cytochromes p450 for terpene functionalisation and
metabolic engineering. Adv Biochem Eng Biotechnol 148:107–139. https://doi.org/10.1007/
10_2014_301
Paddon CJ, Keasling JD (2014) Semi-synthetic artemisinin: A model for the use of synthetic biol-
ogy in pharmaceutical development. Nat Rev Microbiol 12:355–367. https://doi.org/10.1038/
nrmicro3240
Paddon CJ, Westfall PJ, Pitera DJ, Benjamin K, Fisher K, McPhee D, Leavell MD, Tai A, Main
A, Eng D, Polichuk DR, Teoh KH, Reed DW, Treynor T, Lenihan J, Jiang H, Fleck M, Bajad
S, Dang G, Dengrove D, Diola D, Dorin G, Ellens KW, Fickes S, Galazzo J, Gaucher SP,
Geistlinger T, Henry R, Hepp M, Horning T, Iqbal T, Kizer L, Lieu B, Melis D, Moss N,
Regentin R, Secrest S, Tsuruta H, Vazquez R, Westblade LF, Xu L, Yu M, Zhang Y, Zhao L,
Lievense J, Covello PS, Keasling JD, Reiling KK, Renninger NS, Newman JD (2013) High-
level semi-synthetic production of the potent antimalarial artemisinin. Nature 496:528–532.
https://doi.org/10.1038/nature12051
Pan X-W, Han L, Zhang Y-H, Chen D-F, Simonsen HT (2015) Sclareol production in the moss
Physcomitrella patens and observations on growth and terpenoid biosynthesis. Plant Biotechnol
Rep 9:149–159. https://doi.org/10.1007/s11816-015-0353-8
Paradise EM, Kirby J, Chan R, Keasling JD (2008) Redirection of flux through the FPP branch-
point in Saccharomyces cerevisiae by down-regulating squalene synthase. Biotechnol Bioeng
100:371–378
Peng MF, Chen M, Chen R (2011) The last gene involved in the MEP pathway of Artemisia annua:
cloning and characterization and functional identification. J Med Plants Res 5:223–230
Perassolo M, Cardillo AB, Busto VD, Giulietti AM, Talou JR (2018) Biosynthesis of sesquiterpene
lactones in plants and metabolic engineering for their biotechnological production. Chapter 4,
pp 47–91. Sesquiterpene Lactones. https://doi.org/10.1007/978-3-319-78274-4_4
Pirali M, Taheri M, Zarei S, Majidi M, Ghafouri H (2020) Artesunate, as a HSP70 ATPase activity
inhibitor, induces apoptosis in breast cancer cells. Int J Biol Macromol 164:3369–3375. https://
doi.org/10.1016/j.ijbiomac.2020.08.198
Ram M, Khan MA, Jha P (2010) HMG-CoA reductase limits artemisinin biosynthesis and accu-
mulation in Artemisia annua L. plants. Acta Physiol Plant 32:859–866
Reed J, Stephenson MJ, Miettinen K, Brouwer B, Leveau A, Brett P, Goss RJM, Goossens A,
O’Connell MA, Osbourn A (2017) A translational synthetic biology platform for rapid access
to gram-scale quantities of novel drug-like molecules. Metab Eng 42:185–193. https://doi.org/
10.1016/j.ymben.2017.06.012
Reski R, Parsons J, Decker EL (2015) Moss-made pharmaceuticals: from bench to bedside. Plant
Biotechnol J 13(8):1191–1198. Doi: 10.1111/pbi.12401
Riganti C, Doublier S, Viarisio D, Miraglia E, Pescarmona G, Ghigo D, Bosia A (2009)
Artemisinin induces doxorubicin resistance in human colon cancer cells via calcium-dependent
178 A. Beyraghdar Kashkooli et al.
Tu Y (2011) The discovery of artemisinin (qinghaosu) and gifts from Chinese medicine. Nat Med
17:1217–1220
Uckun FM, Saund S, Windlass H, Trieu V (2021) Repurposing anti-malaria phytomedicine
artemisinin as a COVID-19 drug. Front Pharmacol 12:1–5. https://doi.org/10.3389/fphar.2021.
649532
Vranová E, Coman D, Gruissem W (2013) Network analysis of the MVA and MEP pathways
for isoprenoid synthesis. Annu Rev Plant Biol 64:665–700. https://doi.org/10.1146/annurev-arp
lant-050312-120116
Vranová E, Coman D, Gruissem W (2012) Structure and dynamics of the isoprenoid pathway
network. Mol Plant 5:318–333. https://doi.org/10.1093/mp/sss015
Våtsveen TK, Myhre MR, Steen CB, Wälchli S, Lingjærde OC, Bai B, Dillard P, Theodossiou TA,
Holien T, Sundan A, Inderberg EM, Smeland EB, Myklebust JH, Oksvold MP (2018) Arte-
sunate shows potent anti-tumor activity in B-cell lymphoma. J Hematol Oncol 11:1–12. https://
doi.org/10.1186/s13045-018-0561-0
Verpoorte R, Memelink J (2002) Engineering secondary metabolite production in plants. Curr Opin
Biotechnol 13(2):181–187. https://doi.org/10.1016/S0958-1669(02)00308-7
Wang B, Beyraghdar Kashkooli A, Sallets A, Ting HM, de Ruijter NCA, Olofsson L, Brodelius
P, Pottier M, Boutry M, Bouwmeester H et al (2016) Transient production of artemisinin in
Nicotiana benthamiana is boosted by a specific lipid transfer protein from A. annua. Metab
Eng 38:159–169
Wang Y, Huang Z, Wang L, Meng S, Fan Y, Chen T, Cao J, Jiang R, Wang C (2011) The anti-
malarial artemisinin inhibits pro-inflammatory cytokines via the NF-κB canonical signaling
pathway in PMA-induced THP-1 monocytes. Int J Mol Med 27:233–241. https://doi.org/10.
3892/ijmm.2010.580
Wang JS, Wang MJ, Lu X, Zhang J, Liu QX, Zhou D, Dai JG, Zheng H (2020) Artesunate
inhibits epithelial-mesenchymal transition in non-small-cell lung cancer (NSCLC) cells by
down-regulating the expression of BTBD7. Bioengineered 11:1197–1207. https://doi.org/10.
1080/21655979.2020.1834727
Wang L, Li J, Shi X, Li S, Tang PM-K, Li Z, Li H, Wei C (2019) Antimalarial dihydroartemisinin
triggers autophagy within HeLa cells of human cervical cancer through Bcl-2 phosphorylation
at Ser70. Phytomedicine 52:147–156. https://doi.org/10.1016/j.phymed.2018.09.221
Wang W, Yang Y, Zheng XD, Huang SQ, Guo L, Kong JQ, Cheng KD (2013) The advance in
synthetic biology: towards a microbe-derived paclitaxel intermediates. Acta Pharm Sin 48:187–
192 (in Chinese with English abstract)
Willoughby JA, Sundar SN, Cheung M, Tin AS, Modiano J, Firestone GL (2009) Artemisinin
blocks prostate cancer growth and cell cycle progression by disrupting Sp1 interactions with
the cyclin-dependent kinase-4 (CDK4) promoter and inhibiting CDK4 gene expression. J Biol
Chem 284:2203–2213. https://doi.org/10.1074/jbc.M804491200
Wong YK, Xu C, Kalesh KA, He Y, Lin Q, Wong WSF, Shen HM, Wang J (2017) Artemisinin
as an anticancer drug: Recent advances in target profiling and mechanisms of action. Med Res
Rev 37:1492–1517. https://doi.org/10.1002/med.21446
Wu J, Hu D, Yang G, Zhou J, Yang C, Gao Y, Zhu Z (2011) Down-regulation of BMI-1 cooper-
ates with artemisinin on growth inhibition of nasopharyngeal carcinoma cells. J Cell Biochem
112:1938–1948. https://doi.org/10.1002/jcb.23114
Wu L, Cheng Y, Deng J, Tao W, Ye J (2019) Dihydroartemisinin inhibits proliferation and
induces apoptosis of human hepatocellular carcinoma cell by upregulating tumor necrosis fac-
tor via JNK/NF-κB pathways. Evidence-Based Complement. Altern Med.https://doi.org/10.
1155/2019/9581327
Xie DY, Ma DM, Judd R, Jones AL (2016) Artemisinin biosynthesis in Artemisia annua and
metabolic engineering: questions, challenges, and perspectives. Phytochem Rev 15:1093–1114
Yang CQ, Fang X, Wu XM, Mao YB, Wang LJ, Chen XY (2012) Transcriptional regulation of
plant secondary metabolism. J Integr Plant Biol 54:703–712. https://doi.org/10.1111/j.1744-
7909.2012.01161.x
180 A. Beyraghdar Kashkooli et al.
Yang R-Y, Feng L-L, Yang X-Q, Yin L-L, Xu X-L, Zeng Q-P (2008) Quantitative transcript
profiling reveals down-regulation of A sterol pathway relevant gene and overexpression of
artemisinin biogenetic genes in transgenic Artemisia annua plants. Planta Med 74:1510–1516.
https://doi.org/10.1055/s-2008-1081333
Yang J, He Y, Li Y, Zhang X, Wong YK, Shen S, Zhong T, Zhang J, Liu Q, Wang J (2020)
Advances in the research on the targets of anti-malaria actions of artemisinin. Pharmacol Ther
216:107697.https://doi.org/10.1016/j.pharmthera.2020.107697
Yao Z, Bhandari A, Wang Y, Pan Y, Yang F, Chen R, Xia E, Wang O (2018) Dihydroartemisinin
potentiates antitumor activity of 5-fluorouracil against a resistant colorectal cancer cell line.
Biochem Biophys Res Commun 501:636–642. https://doi.org/10.1016/j.bbrc.2018.05.026
Yin S, Yang H, Zhao X, Wei S, Tao Y, Liu M, Bo R, Li J (2020) Antimalarial agent artesunate
induces G0/G1 cell cycle arrest and apoptosis via increasing intracellular ROS levels in normal
liver cells. Hum Exp Toxicol 39:1681–1689. https://doi.org/10.1177/0960327120937331
Yin X, Liu Y, Qin J, Wu Y, Huang J, Zhao Q, Dang T, Tian Y, Yu P, Huang X (2021) Artesunate
suppresses the proliferation and development of estrogen receptor-α-positive endometrial can-
cer in HAND2-Dependent pathway. Front Cell Dev Biol 8:1–14. https://doi.org/10.3389/fcell.
2020.606969
Yu C, Sun P, Zhou Y, Shen B, Zhou M, Wu L, Kong M (2019) Inhibition of AKT enhances
the anti-cancer effects of artemisinin in clear cell renal cell carcinoma. Biomed Pharmacother
118:109383.https://doi.org/10.1016/j.biopha.2019.109383
Zhan X, Zhang YH, Chen DF, Simonsen HT (2014) Metabolic engineering of the moss
physcomitrella patens to produce the sesquiterpenoids patchoulol and α/β-santalene. Front
Plant Sci 5(NOV), 1–10. https://doi.org/10.3389/fpls.2014.00636
Zhang Y, Teoh KH, Reed DW, Maes L, Goossens A, Olson DJH, Ross ARS, Covello PS (2008)
The molecular cloning of artemisinic aldehyde 11(13) reductase and its role in glandular
trichome-dependent biosynthesis of artemisinin in Artemisia annua. J Biol Chem 283:21501–
21508
Zhang L, Jing F, Li F, Li M, Wang Y, Wang G, Sun X, Tang K (2009) Development of trans-
genic Artemisia annua (Chinese wormwood) plants with an enhanced content of artemisinin,
an effective anti-malarial drug, by hairpin-RNA-mediated gene silencing. Biotechnol Appl
Biochem 52:199. https://doi.org/10.1042/BA20080068
Zhang LX, Liu ZN, Ye J, Sha M, Qian H, Bu XH, Luan ZY, Xu XL, Huang AH, Yuan DL, Wu
YQ, Wang XX, Wang J, Huang JX, Ye LH, Zhang HT, Wang YL, Zhang J, Zhang QX (2014)
Artemisinin inhibits gastric cancer cell proliferation through upregulation of p53. Tumor Biol
38:639–646. https://doi.org/10.1002/cbin.10244
Zhang P, Luo HS, Li M, Tan SY (2015) Artesunate inhibits the growth and induces apoptosis of
human gastric cancer cells by downregulating COX-2. Onco Targets Ther 8:845–854. https://
doi.org/10.2147/OTT.S81041
Zhao F, Vakhrusheva O, Markowitsch SD, Slade KS, Tsaur I, Cinatl J, Michaelis M, Efferth T,
Haferkamp A, Juengel E (2020) Artesunate impairs growth in cisplatin-resistant bladder cancer
cells by cell cycle arrest, apoptosis and autophagy induction. Cells 9:1–19. https://doi.org/10.
3390/cells9122643
Zhou HJ, Wang WQ, Wu GD, Lee J, Li A (2007) Artesunate inhibits angiogenesis and downreg-
ulates vascular endothelial growth factor expression in chronic myeloid leukemia K562 cells.
Vascul Pharmacol 47:131–138. https://doi.org/10.1016/j.vph.2007.05.002
Zhou X, Chen Y, Wang F, Wu H, Zhang Y, Liu J, Cai Y, Huang S, He N, Hu Z, Jin X
(2020) Artesunate induces autophagy dependent apoptosis through upregulating ROS and
activating AMPK-mTOR-ULK1 axis in human bladder cancer cells. Chem Biol Interact
331:109273.https://doi.org/10.1016/j.cbi.2020.109273
Zhu S, Liu W, Ke X, Li J, Hu R, Cui H, Song G (2014) Artemisinin reduces cell proliferation
and induces apoptosis in neuroblastoma. Oncol Rep 32:1094–1100. https://doi.org/10.3892/or.
2014.3323
Yeast Synthetic Biology
for the Production of Terpenoids
Derived from Traditional Chinese
Medicinal Plants
Yongjun Wei
Abstract
Traditional Chinese medicine has been widely used to cure diverse diseases in
China. Modern medicinal chemistry and pharmaceutical research help to iden-
tify active natural products that functioned during disease treatment. Among
them, terpenoids are one of the largest natural bioactive products in traditional
Chinese medicine. Traditionally, these active terpenoids are extracted from tra-
ditional Chinese medicinal plants whose yield of the active terpenoids derived
from the plants is low and the quality is not stable. The yield cannot sat-
isfy the increasing terpenoid demand. Therefore, another sustainable supply
of active terpenoids is of great interest. The development of synthetic biol-
ogy enables yeasts to be ideal microbial cell factories for plant-derived natural
product production. In the past few years, several typical active terpenoids
derived from traditional Chinese medicinal plants have been produced in yeasts,
including artemisinic acid, rare ginsenosides, rare licorice triterpenoids, and
other value-added terpenoids. In this chapter, the terpenoid biosynthetic path-
way and current synthetic biology strategies for the production of some typical
terpenoids using yeasts are summarized. Moreover, future synthetic biology
strategies for efficient terpenoid production are discussed.
Y. Wei (B)
Key Laboratory of Advanced Drug Preparation Technologies, Ministry of Education, School of
Pharmaceutical Sciences, Zhengzhou University, Zhengzhou 450001, Henan Province, China
e-mail: [email protected]
Laboratory of Synthetic Biology, Zhengzhou University, Zhengzhou 450001, Henan Province,
China
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 181
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_7
182 Y. Wei
Fig. 1 Comparison of plant production of terpenoids and yeast synthetic biology for the pro-
duction of terpenoids. The traditional Chinese herbal plants often grow months to years, and
production of terpenoids requires plant biomass collection, extraction, and purity. The omics tech-
nologies and available database can help to identify key terpenoid biosynthetic genes, and these
genes can be used for building yeast cell factories. It takes several days to ferment. Compared with
plants, the native natural products in yeasts are few and simple, and it is easy to extract targeted
bioactive terpenoids from yeast biomass
The carbon skeleton and building block for terpenoids is isopentenyl pyrophos-
phate (IPP), and the precursor substrates for IPP formation are acetyl-CoA,
pyruvate, and glycerol-3-phosphate (Wang et al. 2019d). Two distinct pathways,
mevalonate (MVA) and 2-C-methyl-D-Erythritol-4-phosphate (MEP) pathways,
can be used for the production of IPP in nature. Normally, the MVA pathway for
IPP production is mainly reacted in the cytoplasm of the higher eukaryotes; while
in plastids of plant cells, protists, and most microorganisms, the MEP pathway is
used to synthesize IPP (Fig. 2). The intermediates of MVA and MEP pathways
have exchanges and interactions in plants (Liao et al. 2016).
For the MVA pathway, the acetyl-CoA is catalyzed by acetyl-CoA C-
acetyltransferase (AACT) to form acetoacetyl-CoA. The acetoacetyl-CoA is
further catalyzed by 3-hydroxy-3-methylglutaryl-CoA synthase (HMGS) to form
3-hydroxy-3-methylglutaryl-CoA (HMG-CoA). The HMG-CoA is used to gen-
erate MVA catalyzed by the HMG reductase (HMGR). Under multi-step cat-
alytic reactions of MVA kinase and phosphomevalonate kinase, mevalonate
diphosphate (MVAPP) is synthesized. The MVAPP is further decarboxylated
to form IPP by mevalonate diphosphate decarboxylase. The IPP is used for
the synthesis of dimethylallyl pyrophosphate (DMAPP) under the catalysis
of isopentenyl-pyrophosphate delta isomerase (IDI) (Fig. 2) (Vranová et al.
2013). For the MEP pathway, the first step is the formation of D-xylulose-5-
phosphate (DXP) catalyzed by DXP synthase, with pyruvate and glyceraldehyde-
3-phosphate (G3P) as substrates. The second step is the reduction of DXP
to MEP catalyzed by DXP reductoisomerase (Kim and Keasling 2001). With
the help of multi-step enzymatic reactions, MEP is further catalyzed to a
series of products of 4-diphosphocytidyl-2-C-methyl-D-erythritol (CDP-ME), 4-
diphosphocytidyl-2-C-methyl-D-erythritol 2-phosphate (CDP-MEP), 2-C-methyl-
D-erythritol 2,4-cyclopyrophosphate (MECPP), and 1-hydroxy-2-methyl-2-(E)-
butenyl 4-pyrophosphate (HMBPP) (Liu et al. 2014). The HMBPP is transformed
to IPP and DMAPP by 4-hydroxy-3-methylbut-2-en-1-yl diphosphate reductase
(ispH) (Fig. 2).
The terpenoids are synthesized with the common five-carbon substrate of iso-
prenoid (Fig. 2) (Vranová et al. 2012). After the production of IPP and DMAPP,
the geranyl pyrophosphate synthase (GPPS) uses IPP and DMAPP to synthesize
geranyl diphosphate (GPP), and GPP is the precursor for monoterpenoid biosyn-
thesis (Nagegowda and Gupta 2020). Farnesyl pyrophosphate synthase uses IPP
and GPP to synthesize farnesyl diphosphate (FPP), and FPP is the precursor for
sesquiterpenoid and triterpenoids biosynthesis. GPPS uses IPP and FPP to syn-
thesize geranylgeranyl pyrophosphate (GGPP), and GGPP is the precursor for
diterpenoid and tetraterpenoids biosynthesis (Nagegowda and Gupta 2020). In
order to synthesize the final terpenoids, identification of the core structure mod-
ification enzymes, such as the terpene synthases, P450 and their redox partners
of NADPH-cytochrome P450 reductases (CPRs), and glycosyltransferases, are
essential (Wang et al. 2018).
Yeast Synthetic Biology for the Production of Terpenoids … 185
186 Y. Wei
Fig. 2 Terpenoid biosynthetic pathways, MVA pathway and MEP pathway, are described.
G3P, glyceraldehyde-3-phosphate; DXP, 1-deoxy-D-xylulose 5-phosphate; MEP, 2-C-methyl-
D-erythritol 4-phosphate; CDP-ME, 4-diphosphocytidyl-2-C-methyl-D-erythritol; CDP-MEP,
4-diphosphocytidyl-2-C-methyl-D-erythritol 2-phosphate; MECPP, 2-C-methyl-D-erythritol 2,4-
cyclopyrophosphate; HMBPP, 1-hydroxy-2-methyl-2-(E)-butenyl 4-pyrophosphate; HMG-CoA,
3-hydroxy-3-methylglutaryl-CoA; MVA, mevalonic acid; MVAP, mevalonate phosphate; MVAPP,
mevalonate diphosphate; IPP, isopentenyl pyrophosphate; DMAPP, dimethylallyl pyrophosphate;
GPP, geranyl diphosphate; FPP, farnesyl diphosphate; GGPP, Geranylgeranyl pyrophosphate.
DXS, D-xylulose-5-phosphate synthase; DXR, D-xylulose-5-phosphate reductoisomerase;
ispD, 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase; ispE, 4-diphosphocytidyl-2-
C-methyl-D-erythritol kinase; ispF, 2-C-methyl-D-erythritol 2,4-cyclodiphosphate synthase; ispG,
(E)-4-hydroxy-3-methylbut-2-enyl-diphosphate synthase; ispH, 4-hydroxy-3-methylbut-2-en-1-yl
diphosphate reductase; ipk, isopentenyl phosphate kinase; AACT, acetyl-CoA C-acetyltransferase;
HMGS, 3-hydroxy-3-methylglutaryl-CoA synthase; HMGR, 3-hydroxy-3-methylglutaryl-CoA
reductase; MK, mevalonate kinase; PMK, phosphomevalonate kinase; MDC, mevalonate
diphosphate decarboxylase; IDI, isopentenyl-pyrophosphate delta isomerase; IDI, isopentenyl-
pyrophosphate delta isomerase (IDI); GPPS, Geranyl pyrophosphate synthase; FPPS, Farnesyl
pyrophosphate synthase; GGPPS, Geranylgeranyl-PP synthetase
Yeasts, including the model organisms of S. cerevisiae and Y. lipolytica, are attrac-
tive choices for microbial terpene production (Nielsen and Keasling 2016; Arnesen
et al. 2020). In order to produce terpene at a high titer, rate, and yield (TRY),
the terpenoid precursor biosynthesis in the yeasts were enhanced (Nielsen and
Keasling 2016). The common metabolic engineering and newly developed syn-
thetic biology strategies had been applied to overcome precursor limitation in yeast
terpenoid precursor biosynthesis (Farhi et al. 2011; Zu et al. 2020). The yeasts are
not the natural hosts for terpenoid production, and rewiring cellular metabolism
and directing of metabolic flux for terpenoid precursor biosynthesis is necessary
(Worland et al. 2020; Bian et al. 2017). The yeasts only have the MVA pathway
188 Y. Wei
for terpenoid biosynthesis. Based on the MVA pathway in yeast (Fig. 3), overex-
pression of some genes in the MVA pathway and optimization of the terpenoid
precursor biosynthetic pathway by introducing exogenous efficient genes are the
normal strategies (Paramasivan and Mutturi 2017). In fact, diverse genes have been
selected for terpenoid precursor production (Fig. 3).
The rate-limiting enzymes in the MVA pathway are often selected for improving
yeast terpenoid precursor synthesis. The HMGR is the first rate-limiting enzyme
in the MVA pathway. In order to enhance HMGR activity, the tHMG1 (encod-
ing the catalytic domains of yeast HMGR) is overexpressed in yeast, and the
metabolic fluxes in the MVA pathway are increased and the final production of
terpenoids is improved (Fig. 3) (Wu et al. 2018). The overexpression of tHMG1
often leads to the improvement of terpenoid production, such as the limonene pro-
duction using engineered S. cerevisiae (Amiri et al. 2016; Wu et al. 2018). HMG2
is the isoenzyme of HMG1 and is the dominant HMGR isoform under a low oxy-
gen environment (Mantzouridou and Tsimidou 2010). Overexpression of HMG2
variant (K6R) and IDI1 can improve monoterpene production levels (Ignea et al.
2011, 2012). In fact, overexpression of tHMG1, IDI1, and MAF1 increased geranyl
acetate production in S. cerevisiae (Wu et al. 2018).
The ERG20 and ERG9 are two of the most essential enzymes for terpenoid
precursor production (Fig. 3). Engineering ERG20 significantly improved the pro-
duction of sclareol and other terpenoids (Ignea et al. 2015), and overexpression of
fusion genes of ERG20 and BTS1 resulted in the increment of terpenoid production
(Dai et al. 2012). Recently, the overexpression of ERG20 and its mutant mERG20
(F96C mutation in ERG20, which might function as a geranylgeranyl diphosphate
synthase) significantly improved geranylgeraniol content of S. cerevisiae (Dong
et al. 2020). Dynamic control of ERG20 expression level might improve terpenoid
precursor production in yeast. The dynamic control of ERG20 combined with min-
imized endogenous downstream metabolism improved geraniol production (Zhao
et al. 2017); moreover, dynamic control of the expression of ERG20 and ERG9
improved diterpenoid casbene production in S. cerevisiae (Callari et al. 2018).
Downregulation of ERG9 expression level leads to the high-level production of
FPP (Asadollahi et al. 2008), and the common strategy is replacing native ERG9
promoter with HXT1 or other promoters (Scalcinati et al. 2012). Besides, the intro-
duction of exogenous rate-limiting genes of the MVA pathway in yeast could also
improve terpenoid production, such as PaGGPPS and SaGGPPS (Cao et al. 2020).
The lipid production would affect the terpenoid precursor of squalene by storage
of squalene in the lipid body, and co-overexpression of tHMG1 and DGA1 coding
for diacylglycerol acyltransferase led to over 250-fold higher squalene accumula-
tion than a control strain (Wei et al. 2018). The transcriptional activators regulated
terpenoid precursor synthesis in yeasts (Zhang et al. 2017; Davies et al. 2005). The
overexpression of the transcription factor of UPC2 led to the increase of terpenoid
production (Shianna et al. 2001), and other transcription factors, such as Rox1 and
Mot3, repressed ergosterol biosynthesis in yeasts (Montañés et al. 2011).
The extraction of Artemisia annua had been used as an antimalarial drug in tradi-
tional Chinese medicine. In the 1970s, Youyou Tu discovered the active component
artemisinin in the plant extraction of Artemisia annua. Since then, artemisinin was
used as a potent antimalarial drug (Jung et al. 2004). Recent studies suggested
190 Y. Wei
that artemisinin and its derivatives had immune-modulation effects and antitumor
activities (Yao et al. 2016; Kiani et al. 2020). Artemisinin is a sesquiterpene lac-
tone endoperoxide, and it is mainly obtained by extracting from the leaves and
other biomass of A. annua (Efferth 2017). The A. annua needs to grow several
months before harvest, and the artemisinin component in wild A. annua is low (Cai
et al. 2017). Moreover, planting A. annua occupies large amounts of fields, and the
artemisinin component is affected by the climate. Therefore, artemisinin supply is
uncertain and unstable, and the price fluctuates. The production of artemisinin
using microorganisms is of great interest, for it might produce large amounts of
artemisinin with limited space and short time.
Keasling’s group developed bacterial chassis for terpenoid production, and the
obtained E. coli strains are capable of synthesizing diverse terpenoid precursors
(Martin et al. 2003, 2001). The cytochrome P450 is not easy to express in E.
coli, and S. cerevisiae was selected for the production of high-level artemisinin.
The FPP biosynthetic pathway was engineered to increase FPP production and
decrease sterol production by overexpression of tHMGR, UPC2-1, and ERGs, and
downregulation of ERG9. The amorphadiene synthase genes from A. annua was
introduced to S. cerevisiae, in order to construct one strain with amorphadiene
producing ability. Moreover, a novel cytochrome P450 (CYP71AV1) was identified
to have the ability to perform three-step oxidation which can convert amorphadiene
to artemisinic acid. The expression of the CYP71AV1 and its redox partner from
A. annua led to the production of artemisinic acid (Ro et al. 2006). As artemisinic
acid can be easily oxidized to artemisinin using a chemical process, the production
of artemisinic acid is a good proof of concept that using yeasts as microbial cell
factories can synthesize plant natural products. The engineered yeast strains can
produce artemisinic acid in a short time (4–5 days). Compared with the fact that
planting A. annua needs months or years, these synthetic biology strategies save
time and labor for valuable natural product synthesis (Fig. 1).
Further, overexpressing every gene in the MVA pathway doubled artemisinic
acid production; however, the amporph-4,11-diene production is tenfold higher
than artemisinic acid (Westfall et al. 2012). Therefore, they developed one opti-
mized process to convert amporph-4,11-diene to dihydroartemisinic acid. The
dihydroartemisinic acid can be further transformed to artemisinin (Westfall et al.
2012). In order to produce affordable artemisinic acid using yeasts, the artemisinic
acid biosynthetic pathway was optimized by using proper promoters for genes and
introducing efficient artemisinic acid biosynthetic genes. The fermentation titers
of artemisinic acid reached as high as 25 g/L and a low-cost chemical process for
the conversion of artemisinic acid to artemisinin was established, which provided
a second semi-synthetic artemisinin source independent of botanical production
(Paddon et al. 2013). The success of artemisinic acid production using yeasts paves
the way for industrial-scale production of artemisinin derived from the traditional
Chinese medicinal plant.
Yeast Synthetic Biology for the Production of Terpenoids … 191
Fig. 4 Engineering for the production of ginsenosides in S. cerevisiae. GPP, geranyl diphosphate;
FPP, farnesyl diphosphate
catalyzing activities in Panax plants (Yang et al. 2020a). These UGT94 enzymes
were responsible for the synthesis of diverse ginsenosides (Yang et al. 2020a).
Moreover, the UGT diversity in Panax plant was investigated, and two UGTs of
UGTPg45 and UGTPg29 for ginsenosides Rh2 and Rg3 biosynthesis were identi-
fied (Wang et al. 2015). The production of Rh2 and Rg3 using synthetic biology
strategies was applied in engineered yeasts (Wang et al. 2015). A semi-rationally
designed UGT51 derived from yeast was also demonstrated to have the ability to
synthesize Rh2, showing the synthetic biology strategies had great potential for
terpenoid production in yeast (Zhuang et al. 2017). The Rh2 titer increased to
2.25 g/L via expressing multiple copies of UGTPg45 in a PPD-producing chas-
sis, which provides the possible large-scale production of ginsenosides in yeasts
(Wang et al. 2019c). Moreover, the non-conventional yeast, Y. lipolytica, had been
engineered for CK production, and the final CK titer reached to 161.8 mg/L via
fed-batch fermentation in a 5 L fermenter (Li et al. 2019a).
extracted from the roots of the licorice plants, and is one kind of triterpenoids.
Glycyrrhizic acid is mainly used as a sweetening and flavoring agent for bever-
ages, candies, chewing gum, and bitter drugs (Pastorino et al. 2018). The wild
and cultivated Glycyrrhiza plants cannot satisfy the growing market demand for
GA and glycyrrhizic acid. Moreover, the extraction of Glycyrrhizic acid from Gly-
cyrrhiza plants pollutes the environment. For GA structure is complex, chemical
synthesis yield is low and the cost is high (Wang et al. 2019a). Therefore, another
sustainable supply of GA is necessary (Guan et al. 2020).
Compared with the ginsenosides biosynthetic pathway, the GA biosynthetic
pathway is different. The precursors were synthesized with the MVA pathway,
but the 2,3-oxidosqualene is converted to β-amyrin by β-amyrin synthase. The β-
amyrin is oxidized to 11-oxo-β-amyrin with a licorice β-amyrin 11-oxidase which
is a cytochrome P450 enzyme (Seki et al. 2008). Another licorice cytochrome
P450 enzyme of CYP72A154 catalyzed three sequential oxidation steps at C-30
of 11-oxo-β-amyrin, and GA is obtained (Seki et al. 2011). In the catalysis reac-
tion, CPRs provide electrons to the cytochrome P450s (Zhu et al. 2018). The UGTs
(GuGT14 and UGT73P12) are responsible for the biosynthesis of glycyrrhizic acid
(Nomura et al. 2019; Chen et al. 2019). In order to produce high-level GA in
yeasts, the MVA pathway was strengthened and an acetyl-CoA competing path-
way was disrupted. The final β-amyrin production reached 279 mg/L (Liu et al.
2019a). By introducing efficient β-amyrin synthase and P450 enzymes and opti-
mizing metabolic flux to GA synthesis, the final 11-oxo-β-amyrin and GA titer in
shake flask are 80 mg/L and 8.78 mg/L, respectively (Wang et al. 2019a).
Table 1 Some selected terpenoids derived from traditional Chinese medicinal plants which have been synthesized by yeasts
Terpenoid name Traditional Chinese Yeast chassis Titer Terpenoid type References
medicinal plant source
Geraniol Diverse plants Saccharomyces cerevisiae 1.68 g/L Monoterpenoid (Jiang et al. 2017)
strain CEN.PK2-1C
α-Terpineol Pine and cypress plants Saccharomyces cerevisiae 21.88 mg/L Monoterpenoid (Zhang et al. 2019)
strain LCB08
Limonene Citrus fruits Saccharomyces cerevisiae 166 mg/L Monoterpenoid (Ignea et al. 2019)
strain BY4741
Miltiradiene Salvia miltiorrhiza Bge Saccharomyces cerevisiae 3.5 g/L Diterpenoid (Hu et al. 2020)
strain S288C
Triptolide Tripterygium wilfordii Saccharomyces cerevisiae 30.5 μg/g yeast Diterpenoid (Tu et al. 2020)
strain BY-HZ16 biomass
(-)-β-elemene Curcuma wenyujin Saccharomyces cerevisiae 190.7 mg/L Sesquiterpenoid (Hu et al. 2017)
strain SCIGS22a
Artemisinin Artemisia annua L Saccharomyces cerevisiae 25 g/L Sesquiterpenoid (Paddon et al. 2013)
strain Y1284
Glycyrrhetinic acid Glycyrrhize glabra L Saccharomyces cerevisiae 8.78 mg/L Triterpenoid (Wang et al. 2019a)
strain Y7
Ginsenoside compound K Panax plants Saccharomyces cerevisiae 1.7 g/L Triterpenoid (Nan et al. 2020)
(CK) strain (WLN-3)
Ginsenoside F1 Panax plants Saccharomyces cerevisiae 450. 5 mg/L Triterpenoid (Wang et al. 2019b)
strain (BY-F1)
Ginsenoside Rh2 Panax plants Saccharomyces cerevisiae 2.25 g/L Triterpenoid (Wang et al. 2019c)
strain ZWDRH2-10
(continued)
Y. Wei
Table 1 (continued)
Terpenoid name Traditional Chinese Yeast chassis Titer Terpenoid type References
medicinal plant source
Ginsenoside Rh1 Panax plants Saccharomyces cerevisiae 92.8 mg/L Triterpenoid (Wei et al. 2015a)
strain ZW-Rh1-20
Ginsenoside Rg3 Panax plants Saccharomyces cerevisiae 3.49 μmol/g dry cell Triterpenoid (Wang et al. 2015)
strain D20RG1 weight
Protopanaxadiol (PPD) Panax plants Saccharomyces cerevisiae 152.37 mg/L Triterpenoid (Gao et al. 2018)
strain GW10
Yeast Synthetic Biology for the Production of Terpenoids …
195
196 Y. Wei
The omics-based technologies can help to identify key genes easily from omics
data of the traditional Chinese medicinal plants, and these key genes can be
used for building yeast cell factories. Engineering for efficient yeast terpenoid
precursor-producing chassis, introducing heterogenous terpenoid synthetic path-
ways, increasing key gene expression level, directing metabolic flux for terpenoid
synthesis by downregulating competing pathway, enhancing cofactor supply for
MVA synthesis, engineering sub-organelle for terpenoid production, and other fre-
quently used strategies would significantly improve titer, rate, and yield of yeast
terpenoid production.
With the development of genomic-scale models and high-throughput automated
screening, engineered yeasts for efficient terpenoid production would be acceler-
ated. The optimized fermentation and bioprocess could decrease the cost of yeast
terpenoids. Besides, to produce affordable or commercial terpenoids, developing
smart biomanufacturing with a low-cost, intelligent, and continuous process will
lay the foundation for industrial-scale yeast production of valuable terpenoids. The
production of many valuable terpenoids derived from traditional Chinese medicinal
plants will be achieved in the near future.
Acknowledgements This work was supported by the National Natural Science Foundation of
China (No. 31800079).
References
Allen KD, McKernan K, Pauli C, Roe J, Torres A, Gaudino R (2019) Genomic charsacter-
ization of the complete terpene synthase gene family from Cannabis sativa. Plos One
14(9):e0222363.https://doi.org/10.1371/journal.pone.0222363
Amiri P, Shahpiri A, Asadollahi MA, Momenbeik F, Partow S (2016) Metabolic engineering of
Saccharomyces cerevisiae for linalool production. Biotechnol Lett 38(3):503–508. https://doi.
org/10.1007/s10529-015-2000-4
Arnesen JA, Kildegaard KR, Cernuda Pastor M, Jayachandran S, Kristensen M, Borodina I (2020)
Yarrowia lipolytica Strains engineered for the production of terpenoids. Front Bioeng Biotech-
nol 8(945). https://doi.org/10.3389/fbioe.2020.00945
Asadollahi MA, Maury J, Møller K, Nielsen KF, Schalk M, Clark A, Nielsen J (2008) Production
of plant sesquiterpenes in Saccharomyces cerevisiae: effect of ERG9 repression on sesquiter-
pene biosynthesis. Biotechnol Bioeng 99(3):666–677. https://doi.org/10.1002/bit.21581
Asl MN, Hosseinzadeh H (2008) Review of pharmacological effects of Glycyrrhiza sp. and its
bioactive compounds. Phytotherapy Res 22(6):709–724
Bao T, Shadrack K, Yang S, Xue X, Li S, Wang N, Wang Q, Wang L, Gao X, Cronk Q (2020)
Functional characterization of terpene synthases accounting for the volatilized-terpene hetero-
geneity in Lathyrus odoratus cultivar flowers. Plant Cell Physiol 61(10):1733–1749. https://doi.
org/10.1093/pcp/pcaa100
Bergman ME, Davis B, Phillips MA (2019) Medically useful plant terpenoids: biosynthesis, occur-
rence, and mechanism of action. Molecules (Basel, Switzerland) 24(21). https://doi.org/10.
3390/molecules24213961
Bian G, Deng Z, Liu T (2017) Strategies for terpenoid overproduction and new terpenoid discovery.
Curr Opin Biotechnol 48:234–241. https://doi.org/10.1016/j.copbio.2017.07.002
198 Y. Wei
Cai T-Y, Zhang Y-R, Ji J-B, Xing J (2017) Investigation of the component in Artemisia annua L.
leading to enhanced antiplasmodial potency of artemisinin via regulation of its metabolism. J
Ethnopharmacol 207:86–91. https://doi.org/10.1016/j.jep.2017.06.025
Callari R, Meier Y, Ravasio D, Heider H (2018) Dynamic control of ERG20 and ERG9 expres-
sion for improved casbene production in Saccharomyces cerevisiae. Front Bioeng Biotechnol
6(160). https://doi.org/10.3389/fbioe.2018.00160
Cao X, Yang S, Cao C, Zhou YJ (2020) Harnessing sub-organelle metabolism for biosynthesis
of isoprenoids in yeast. Synth Syst Biotechnol 5(3):179–186. https://doi.org/10.1016/j.synbio.
2020.06.005
Chen K, Hu Z-m, Song W, Wang Z-l, He J-b, Shi X-m, Cui Q-h, Qiao X, Ye M (2019) Diversity of
o-glycosyltransferases contributes to the biosynthesis of flavonoid and triterpenoid glycosides
in Glycyrrhiza uralensis. Acs Synth Biol 8(8):1858–1866. https://doi.org/10.1021/acssynbio.
9b00171
Cheng W, Huang C, Ma W, Tian X, Zhang X (2019) Recent development of oridonin derivatives
with diverse pharmacological activities. Mini Rev Med Chem 19(2):114–124. https://doi.org/
10.2174/1389557517666170417170609
Chu LL, Montecillo JAV, Bae H (2020) Recent advances in the metabolic engineering of yeasts for
ginsenoside biosynthesis. Front Bioeng Biotechnol 8(139). https://doi.org/10.3389/fbioe.2020.
00139
Chung SY, Seki H, Fujisawa Y, Shimoda Y, Hiraga S, Nomura Y, Saito K, Ishimoto M, Muranaka
T (2020) A cellulose synthase-derived enzyme catalyses 3-O-glucuronosylation in saponin
biosynthesis. Nat Commun 11(1):5664. https://doi.org/10.1038/s41467-020-19399-0
Dai Z, Liu Y, Huang L, Zhang X (2012) Production of miltiradiene by metabolically engineered
Saccharomyces cerevisiae. Biotechnol Bioeng 109(11):2845–2853. https://doi.org/10.1002/bit.
24547
Davies BS, Wang HS, Rine J (2005) Dual activators of the sterol biosynthetic pathway of Saccha-
romyces cerevisiae: similar activation/regulatory domains but different response mechanisms.
Mol Cell Biol 25(16):7375–7385. https://doi.org/10.1128/mcb.25.16.7375-7385.2005
Dong H, Chen S, Zhu J, Gao K, Zha W, Lin P, Zi J (2020) Enhance production of diterpenoids in
Yeast by overexpression of the fused enzyme of ERG20 and its mutant mERG20. J Biotechnol
307:29–34. https://doi.org/10.1016/j.jbiotec.2019.10.019
Efferth T (2017) From ancient herb to modern drug: Artemisia annua and artemisinin for cancer
therapy. Semin Cancer Biol 46:65–83. https://doi.org/10.1016/j.semcancer.2017.02.009
Expósito O, Bonfill M, Moyano E, Onrubia M, Mirjalili MH, Cusidó RM, Palazón J (2009)
Biotechnological production of taxol and related taxoids: current state and prospects. Anti-
cancer Agents Med Chem 9(1):109–121. https://doi.org/10.2174/187152009787047761
Farhi M, Marhevka E, Masci T, Marcos E, Eyal Y, Ovadis M, Abeliovich H, Vainstein A (2011)
Harnessing yeast subcellular compartments for the production of plant terpenoids. Metab Eng
13(5):474–481. https://doi.org/10.1016/j.ymben.2011.05.001
Gallego-Jara J, Lozano-Terol G, Sola-Martínez RA, Cánovas-Díaz M, de Diego Puente T (2020) A
compressive review about Taxol(®): History and future challenges. Molecules (Basel, Switzer-
land) 25(24). https://doi.org/10.3390/molecules25245986
Gao X, Caiyin Q, Zhao F, Wu Y, Lu W (2018) Engineering Saccharomyces cerevisiae for enhanced
production of protopanaxadiol with cofermentation of glucose and xylose. J Agr Food Chem
66(45):12009–12016. https://doi.org/10.1021/acs.jafc.8b04916
Gao Q, Wang L, Zhang M, Wei Y, Lin W (2020) Recent advances on feasible strategies for
monoterpenoid production in Saccharomyces cerevisiae. Front Bioeng Biotechnol 8:1372
Guan R, Wang M, Guan Z, Jin C-Y, Lin W, Ji X, Wei Y (2020) Metabolic engineering for gly-
cyrrhetinic acid production in Saccharomyces cerevisiae. Front Bioeng Biotechnol 8:1318
Han JY, Kwon YS, Yang DC, Jung YR, Choi YE (2006) Expression and RNA interference-
induced silencing of the dammarenediol synthase gene in Panax ginseng. Plant Cell Physiol
47(12):1653–1662. https://doi.org/10.1093/pcp/pcl032
Yeast Synthetic Biology for the Production of Terpenoids … 199
Han J-Y, Kim H-J, Kwon Y-S, Choi Y-E (2011) The Cyt P450 Enzyme CYP716A47 catalyzes the
formation of protopanaxadiol from dammarenediol-II during ginsenoside biosynthesis in Panax
ginseng. Plant Cell Physiol 52(12):2062–2073. https://doi.org/10.1093/pcp/pcr150
Han R, Rai A, Nakamura M, Suzuki H, Takahashi H, Yamazaki M, Saito K (2016) De novo deep
transcriptome analysis of medicinal plants for gene discovery in biosynthesis of plant natural
products. Methods Enzymol 576:19–45. https://doi.org/10.1016/bs.mie.2016.03.001
Han X, Li W, Duan Z, Ma X, Fan D (2020) Biocatalytic production of compound K in a deep eutec-
tic solvent based on choline chloride using a substrate fed-batch strategy. Bioresour Technol
305:123039.https://doi.org/10.1016/j.biortech.2020.123039
Hu Y, Zhou YJ, Bao J, Huang L, Nielsen J, Krivoruchko A (2017) Metabolic engineering of Sac-
charomyces cerevisiae for production of germacrene A, a precursor of beta-elemene. J Ind
Microbiol Biotechnol 44(7):1065–1072. https://doi.org/10.1007/s10295-017-1934-z
Hu T, Zhou J, Tong Y, Su P, Li X, Liu Y, Liu N, Wu X, Zhang Y, Wang J, Gao L, Tu L, Lu Y,
Jiang Z, Zhou YJ, Gao W, Huang L (2020) Engineering chimeric diterpene synthases and iso-
prenoid biosynthetic pathways enables high-level production of miltiradiene in yeast. Metab
Eng 60:87–96. https://doi.org/10.1016/j.ymben.2020.03.011
Huang M, Lu JJ, Huang MQ, Bao JL, Chen XP, Wang YT (2012) Terpenoids: natural products
for cancer therapy. Expert Opin Investig Drugs 21(12):1801–1818. https://doi.org/10.1517/135
43784.2012.727395
Hwang CR, Lee SH, Jang GY, Hwang IG, Kim HY, Woo KS, Lee J, Jeong HS (2014) Changes in
ginsenoside compositions and antioxidant activities of hydroponic-cultured ginseng roots and
leaves with heating temperature. J Ginseng Res 38(3):180–186. https://doi.org/10.1016/j.jgr.
2014.02.002
Ignea C, Cvetkovic I, Loupassaki S, Kefalas P, Johnson CB, Kampranis SC, Makris AM (2011)
Improving yeast strains using recyclable integration cassettes, for the production of plant ter-
penoids. Microb Cell Fact 10:4. https://doi.org/10.1186/1475-2859-10-4
Ignea C, Trikka FA, Kourtzelis I, Argiriou A, Kanellis AK, Kampranis SC, Makris AM (2012)
Positive genetic interactors of HMG2 identify a new set of genetic perturbations for improving
sesquiterpene production in Saccharomyces cerevisiae. Microb Cell Fact 11:162. https://doi.
org/10.1186/1475-2859-11-162
Ignea C, Pontini M, Maffei ME, Makris AM, Kampranis SC (2014) Engineering monoterpene pro-
duction in yeast using a synthetic dominant negative geranyl diphosphate synthase. Acs Synth
Biol 3(5):298–306. https://doi.org/10.1021/sb400115e
Ignea C, Trikka FA, Nikolaidis AK, Georgantea P, Ioannou E, Loupassaki S, Kefalas P, Kanellis
AK, Roussis V, Makris AM, Kampranis SC (2015) Efficient diterpene production in yeast by
engineering Erg20p into a geranylgeranyl diphosphate synthase. Metab Eng 27:65–75. https://
doi.org/10.1016/j.ymben.2014.10.008
Ignea C, Raadam MH, Motawia MS, Makris AM, Vickers CE, Kampranis SC (2019) Orthogo-
nal monoterpenoid biosynthesis in yeast constructed on an isomeric substrate. Nat Commun
10(1):3799. https://doi.org/10.1038/s41467-019-11290-x
Jaeger R, Cuny E (2016) Terpenoids with special pharmacological significance: a review. Nat Prod
Commun 11(9):1373–1390
Jansen DJ, Shenvi RA (2014) Synthesis of medicinally relevant terpenes: reducing the cost and
time of drug discovery. Future Med Chem 6(10):1127–1148. https://doi.org/10.4155/fmc.14.71
Jiang Z, Kempinski C, Chappell J (2016) Extraction and analysis of terpenes/terpenoids. Curr
Protoc Plant Biol 1:345–358. https://doi.org/10.1002/cppb.20024
Jiang G-Z, Yao M-D, Wang Y, Zhou L, Song T-Q, Liu H, Xiao W-H, Yuan Y-J (2017) Manipula-
tion of GES and ERG20 for geraniol overproduction in Saccharomyces cerevisiae. Metab Eng
41:57–66. https://doi.org/10.1016/j.ymben.2017.03.005
Jozwiak A, Sonawane PD, Panda S, Garagounis C, Papadopoulou KK, Abebie B, Massalha H,
Almekias-Siegl E, Scherf T, Aharoni A (2020) Plant terpenoid metabolism co-opts a compo-
nent of the cell wall biosynthesis machinery. Nat Chem Biol 16(7):740–748. https://doi.org/10.
1038/s41589-020-0541-x
200 Y. Wei
Jung M, Lee K, Kim H, Park M (2004) Recent advances in artemisinin and its derivatives as anti-
malarial and antitumor agents. Curr Med Chem 11(10):1265–1284. https://doi.org/10.2174/092
9867043365233
Kiani BH, Kayani WK, Khayam AU, Dilshad E, Ismail H, Mirza B (2020) Artemisinin and its
derivatives: a promising cancer therapy. Mol Biol Rep 47(8):6321–6336. https://doi.org/10.
1007/s11033-020-05669-z
Kim SW, Keasling JD (2001) Metabolic engineering of the nonmevalonate isopentenyl diphos-
phate synthesis pathway in Escherichia coli enhances lycopene production. Biotechnol Bioeng
72(4):408–415. https://doi.org/10.1002/1097-0290(20000220)72:4%3c408::aid-bit1003%3e3.
0.co;2-h
Kowalska A, Kalinowska-Lis U (2019) 18β-Glycyrrhetinic acid: its core biological properties
and dermatological applications. Int J Cosmet Sci 41(4):325–331. https://doi.org/10.1111/ics.
12548
Kumar S, Kempinski C, Zhuang X, Norris A, Mafu S, Zi J, Bell SA, Nybo SE, Kinison SE, Jiang
Z, Goklany S, Linscott KB, Chen X, Jia Q, Brown SD, Bowman JL, Babbitt PC, Peters RJ,
Chen F, Chappell J (2016) Molecular diversity of terpene synthases in the liverwort Marchantia
polymorpha. Plant Cell 28(10):2632–2650. https://doi.org/10.1105/tpc.16.00062
Leebens-Mack JH, Barker MS, Carpenter EJ, Deyholos MK, Gitzendanner MA, Graham SW,
Grosse I, Li Z, Melkonian M, Mirarab S, Porsch M, Quint M, Rensing SA, Soltis DE, Soltis
PS, Stevenson DW, Ullrich KK, Wickett NJ, DeGironimo L, Edger PP, Jordon-Thaden IE, Joya
S, Liu T, Melkonian B, Miles NW, Pokorny L, Quigley C, Thomas P, Villarreal JC, Augustin
MM, Barrett MD, Baucom RS, Beerling DJ, Benstein RM, Biffin E, Brockington SF, Burge
DO, Burris JN, Burris KP, Burtet-Sarramegna V, Caicedo AL, Cannon SB, Çebi Z, Chang Y,
Chater C, Cheeseman JM, Chen T, Clarke ND, Clayton H, Covshoff S, Crandall-Stotler BJ,
Cross H, dePamphilis CW, Der JP, Determann R, Dickson RC, Di Stilio VS, Ellis S, Fast E,
Feja N, Field KJ, Filatov DA, Finnegan PM, Floyd SK, Fogliani B, García N, Gâteblé G, God-
den GT, Goh F, Greiner S, Harkess A, Heaney JM, Helliwell KE, Heyduk K, Hibberd JM,
Hodel RGJ, Hollingsworth PM, Johnson MTJ, Jost R, Joyce B, Kapralov MV, Kazamia E, Kel-
logg EA, Koch MA, Von Konrat M, Könyves K, Kutchan TM, Lam V, Larsson A, Leitch AR,
Lentz R, Li F-W, Lowe AJ, Ludwig M, Manos PS, Mavrodiev E, McCormick MK, McKain M,
McLellan T, McNeal JR, Miller RE, Nelson MN, Peng Y, Ralph P, Real D, Riggins CW, Ruh-
sam M, Sage RF, Sakai AK, Scascitella M, Schilling EE, Schlösser E-M, Sederoff H, Servick
S, Sessa EB, Shaw AJ, Shaw SW, Sigel EM, Skema C, Smith AG, Smithson A, Stewart CN,
Stinchcombe JR, Szövényi P, Tate JA, Tiebel H, Trapnell D, Villegente M, Wang C-N, Weller
SG, Wenzel M, Weststrand S, Westwood JH, Whigham DF, Wu S, Wulff AS, Yang Y, Zhu D,
Zhuang C, Zuidof J, Chase MW, Pires JC, Rothfels CJ, Yu J, Chen C, Chen L, Cheng S, Li J,
Li R, Li X, Lu H, Ou Y, Sun X, Tan X, Tang J, Tian Z, Wang F, Wang J, Wei X, Xu X, Yan
Z, Yang F, Zhong X, Zhou F, Zhu Y, Zhang Y, Ayyampalayam S, Barkman TJ, Nguyen N-p,
Matasci N, Nelson DR, Sayyari E, Wafula EK, Walls RL, Warnow T, An H, Arrigo N, Bani-
aga AE, Galuska S, Jorgensen SA, Kidder TI, Kong H, Lu-Irving P, Marx HE, Qi X, Reardon
CR, Sutherland BL, Tiley GP, Welles SR, Yu R, Zhan S, Gramzow L, Theißen G, Wong GK-
S (2019) One thousand plant transcriptomes and the phylogenomics of green plants. Nature
574(7780):679–685. https://doi.org/10.1038/s41586-019-1693-2
Li D, Wu Y, Zhang C, Sun J, Zhou Z, Lu W (2019a) Production of triterpene ginsenoside com-
pound K in the non-conventional yeast Yarrowia lipolytica. J Agric Food Chem 67(9):2581–
2588. https://doi.org/10.1021/acs.jafc.9b00009
Li J, Mutanda I, Wang K, Yang L, Wang J, Wang Y (2019b) Chloroplastic metabolic engineering
coupled with isoprenoid pool enhancement for committed taxanes biosynthesis in Nicotiana
benthamiana. Nat Commun 10(1):4850. https://doi.org/10.1038/s41467-019-12879-y
Li JY, Cao HY, Liu P, Cheng GH, Sun MY (2014) Glycyrrhizic acid in the treatment of liver
diseases: literature review. Biomed Res Int 2014:872139. https://doi.org/10.1155/2014/872139
Liang J, Mai W, Tang J, Wei Y (2019) Highly effective treatment of petrochemical wastewater by
a super-sized industrial scale plant with expanded granular sludge bed bioreactor and aerobic
activated sludge. Chem Eng J 360:15–23. https://doi.org/10.1016/j.cej.2018.11.167
Yeast Synthetic Biology for the Production of Terpenoids … 201
Liao P, Hemmerlin A, Bach TJ, Chye M-L (2016) The potential of the mevalonate pathway for
enhanced isoprenoid production. Biotechnol Adv 34(5):697–713. https://doi.org/10.1016/j.bio
techadv.2016.03.005
Liu H, Wang Y, Tang Q, Kong W, Chung W-J, Lu T (2014) MEP pathway-mediated isopentenol
production in metabolically engineered Escherichia coli. Microb Cell Fact 13(1):135. https://
doi.org/10.1186/s12934-014-0135-y
Liu H, Fan J, Wang C, Li C, Zhou X (2019a) Enhanced β-amyrin synthesis in Saccharomyces
cerevisiae by coupling an optimal acetyl-CoA supply pathway. J Agr Food Chem 67(13):3723–
3732
Liu Z, Zhang Y, Nielsen J (2019b) Synthetic biology of yeast. Biochemistry 58(11):1511–1520.
https://doi.org/10.1021/acs.biochem.8b01236
Liu X, Zhu X, Wang H, Liu T, Cheng J, Jiang H (2020) Discovery and modification of cytochrome
P450 for plant natural products biosynthesis. Synthetic and Systems Biotechnology 5(3):187–
199. https://doi.org/10.1016/j.synbio.2020.06.008
Ma Y, Li W, Mai J, Wang J, Wei Y, Ledesma-Amaro R, Ji X-J (2020) Engineering Yarrowia lipoly-
tica for sustainable production of the chamomile sesquiterpene (−)-α-bisabolol. Green Chem.
https://doi.org/10.1039/D0GC03180A
Mai W, Hu T, Li C, Wu R, Chen J, Shao Y, Liang J, Wei Y (2020) Effective nitrogen removal of
wastewater from vitamin B2 production by a potential anammox process. J Water Process Eng
37:101515.https://doi.org/10.1016/j.jwpe.2020.101515
Malik S, Cusidó RM, Mirjalili MH, Moyano E, Palazón J, Bonfill M (2011) Production of the anti-
cancer drug taxol in Taxus baccata suspension cultures: a review. Process Biochem 46(1):23–
34. https://doi.org/10.1016/j.procbio.2010.09.004
Mantzouridou F, Tsimidou MZ (2010) Observations on squalene accumulation in Saccharomyces
cerevisiae due to the manipulation of HMG2 and ERG6. FEMS Yeast Res 10(6):699–707.
https://doi.org/10.1111/j.1567-1364.2010.00645.x
Martin VJ, Yoshikuni Y, Keasling JD (2001) The in vivo synthesis of plant sesquiterpenes by
Escherichia coli. Biotechnol Bioeng 75(5):497–503. https://doi.org/10.1002/bit.10037
Martin VJ, Pitera DJ, Withers ST, Newman JD, Keasling JD (2003) Engineering a mevalonate
pathway in Escherichia coli for production of terpenoids. Nat Biotechnol 21(7):796–802.
https://doi.org/10.1038/nbt833
Montañés FM, Pascual-Ahuir A, Proft M (2011) Repression of ergosterol biosynthesis is essential
for stress resistance and is mediated by the Hog1 MAP kinase and the Mot3 and Rox1 tran-
scription factors. Mol Microbiol 79(4):1008–1023. https://doi.org/10.1111/j.1365-2958.2010.
07502.x
Moser S, Pichler H (2019) Identifying and engineering the ideal microbial terpenoid production
host. Appl Microbiol Biotechnol 103(14):5501–5516. https://doi.org/10.1007/s00253-019-098
92-y
Nagegowda DA, Gupta P (2020) Advances in biosynthesis, regulation, and metabolic engineering
of plant specialized terpenoids. Plant Sci 294:110457.https://doi.org/10.1016/j.plantsci.2020.
110457
Namuli A, Bazira J, Casim TU, Engeu PO (2018) A review of various efforts to increase
artemisinin and other antimalarial compounds in Artemisia Annua L plant. Cogent Biol 4(1).
https://doi.org/10.1080/23312025.2018.1513312
Nan W, Zhao F, Zhang C, Ju H, Lu W (2020) Promotion of compound K production in Sac-
charomyces cerevisiae by glycerol. Microb Cell Fact 19(1):41. https://doi.org/10.1186/s12934-
020-01306-3
Navarro-Muñoz JC, Selem-Mojica N, Mullowney MW, Kautsar SA, Tryon JH, Parkinson EI, De
Los Santos ELC, Yeong M, Cruz-Morales P, Abubucker S, Roeters A, Lokhorst W, Fernandez-
Guerra A, Cappelini LTD, Goering AW, Thomson RJ, Metcalf WW, Kelleher NL, Barona-
Gomez F, Medema MH (2020) A computational framework to explore large-scale biosynthetic
diversity. Nat Chem Biol 16(1):60–68. https://doi.org/10.1038/s41589-019-0400-9
Nett RS, Lau W, Sattely ES (2020) Discovery and engineering of colchicine alkaloid biosynthesis.
Nature. https://doi.org/10.1038/s41586-020-2546-8
202 Y. Wei
Wei Y, Siewers V, Nielsen J (2017b) Cocoa butter-like lipid production ability of non-oleaginous
and oleaginous yeasts under nitrogen-limited culture conditions. Appl Microbiol Biotechnol
101(9):3577–3585. https://doi.org/10.1007/s00253-017-8126-7
Wei L-J, Kwak S, Liu J-J, Lane S, Hua Q, Kweon D-H, Jin Y-S (2018) Improved squalene pro-
duction through increasing lipid contents in Saccharomyces cerevisiae. Biotechnol Bioeng
115(7):1793–1800. https://doi.org/10.1002/bit.26595
Wei Y, Zhou H, Zhang J, Zhang L, Geng A, Liu F, Zhao G, Wang S, Zhou Z, Yan X (2015b) Insight
into dominant cellulolytic bacteria from two biogas digesters and their glycoside hydrolase
genes. Plos One 10(6):e0129921.https://doi.org/10.1371/journal.pone.0129921
Westfall PJ, Pitera DJ, Lenihan JR, Eng D, Woolard FX, Regentin R, Horning T, Tsuruta H,
Melis DJ, Owens A, Fickes S, Diola D, Benjamin KR, Keasling JD, Leavell MD, McPhee DJ,
Renninger NS, Newman JD, Paddon CJ (2012) Production of amorphadiene in yeast, and its
conversion to dihydroartemisinic acid, precursor to the antimalarial agent artemisinin. Proc Natl
Acad Sci 109(3):E111–E118. https://doi.org/10.1073/pnas.1110740109
Worland AM, Czajka JJ, Li Y, Wang Y, Tang YJ, Su WW (2020) Biosynthesis of terpene com-
pounds using the non-model yeast Yarrowia lipolytica: grand challenges and a few perspectives.
Curr Opin Biotechnol 64:134–140. https://doi.org/10.1016/j.copbio.2020.02.020
Wu T, Li S, Zhang B, Bi C, Zhang X (2018) Engineering Saccharomyces cerevisiae for the pro-
duction of the valuable monoterpene ester geranyl acetate. Microb Cell Fact 17(1):85. https://
doi.org/10.1186/s12934-018-0930-y
Yaegashi J, Kirby J, Ito M, Sun J, Dutta T, Mirsiaghi M, Sundstrom ER, Rodriguez A, Baidoo E,
Tanjore D, Pray T, Sale K, Singh S, Keasling JD, Simmons BA, Singer SW, Magnuson JK,
Arkin AP, Skerker JM, Gladden JM (2017) Rhodosporidium toruloides: a new platform organ-
ism for conversion of lignocellulose into terpene biofuels and bioproducts. Biotechnol Biofuels
10(1):241. https://doi.org/10.1186/s13068-017-0927-5
Yan X, Fan Y, Wei W, Wang P, Liu Q, Wei Y, Zhang L, Zhao G, Yue J, Zhou Z (2014) Production of
bioactive ginsenoside compound K in metabolically engineered yeast. Cell Res 24(6):770–773.
https://doi.org/10.1038/cr.2014.28
Yang L, Zou H, Gao Y, Luo J, Xie X, Meng W, Zhou H, Tan Z (2020) Insights into gastroin-
testinal microbiota-generated ginsenoside metabolites and their bioactivities. Drug Metab Rev
52(1):125–138. https://doi.org/10.1080/03602532.2020.1714645
Yang C, Li C, Wei W, Wei Y, Liu Q, Zhao G, Yue J, Yan X, Wang P, Zhou Z (2020) The
unprecedented diversity of UGT94-family UDP-glycosyltransferases in Panax plants and their
contribution to ginsenoside biosynthesis. Sci Rep-Uk 10(1):15394. https://doi.org/10.1038/s41
598-020-72278-y
Yao W, Wang F, Wang H (2016) Immunomodulation of artemisinin and its derivatives. Science
Bulletin 61(18):1399–1406. https://doi.org/10.1007/s11434-016-1105-z
Zeng F, Wang W, Guan S, Cheng C, Yang M, Avula B, Khan IA, Guo DA (2013) Simul-
taneous quantification of 18 bioactive constituents in Tripterygium wilfordii using liq-
uid chromatography-electrospray ionization-mass spectrometry. Planta Med 79(9):797–805.
https://doi.org/10.1055/s-0032-1328596
Zhang Q, Ye M (2009) Chemical analysis of the Chinese herbal medicine Gan-Cao (licorice). J
Chromatogr A 1216(11):1954–1969
Zhang C, Li M, Zhao G-R, Lu W (2019) Alpha-Terpineol production from an engineered Saccha-
romyces cerevisiae cell factory. Microb Cell Fact 18(1):160. https://doi.org/10.1186/s12934-
019-1211-0
Zhang Y, Nielsen J, Liu Z (2017) Engineering yeast metabolism for production of terpenoids for
use as perfume ingredients, pharmaceuticals and biofuels. FEMS Yeast Res 17(8). https://doi.
org/10.1093/femsyr/fox080
Zhao J, Li C, Zhang Y, Shen Y, Hou J, Bao X (2017) Dynamic control of ERG20 expression com-
bined with minimized endogenous downstream metabolism contributes to the improvement of
geraniol production in Saccharomyces cerevisiae. Microb Cell Fact 16(1):17. https://doi.org/10.
1186/s12934-017-0641-9
Yeast Synthetic Biology for the Production of Terpenoids … 205
Zhao M, Lin Y, Wang Y, Li X, Han Y, Wang K, Sun C, Wang Y, Zhang M (2019) Transcriptome
analysis identifies strong candidate genes for ginsenoside biosynthesis and reveals its underly-
ing molecular mechanism in Panax ginseng C.A. Meyer. Sci Rep-Uk 9(1):615–615. https://doi.
org/10.1038/s41598-018-36349-5
Zhou F, Pichersky E (2020) More is better: the diversity of terpene metabolism in plants. Curr Opin
Plant Biol 55:1–10. https://doi.org/10.1016/j.pbi.2020.01.005
Zhou Z-L, Yang Y-X, Ding J, Li Y-C, Miao Z-H (2012) Triptolide: structural modifications, struc-
ture–activity relationships, bioactivities, clinical development and mechanisms. Nat Prod Rep
29(4):457–475. https://doi.org/10.1039/C2NP00088A
Zhu M, Wang C, Sun W, Zhou A, Wang Y, Zhang G, Zhou X, Huo Y, Li C (2018) Boosting 11-
oxo-β-amyrin and glycyrrhetinic acid synthesis in Saccharomyces cerevisiae via pairing novel
oxidation and reduction system from legume plants. Metab Eng 45:43–50. https://doi.org/10.
1016/j.ymben.2017.11.009
Zhuang Y, Yang G-Y, Chen X, Liu Q, Zhang X, Deng Z, Feng Y (2017) Biosynthesis of plant-
derived ginsenoside Rh2 in yeast via repurposing a key promiscuous microbial enzyme. Metab
Eng 42:25–32. https://doi.org/10.1016/j.ymben.2017.04.009
Zu Y, Prather KLJ, Stephanopoulos G (2020) Metabolic engineering strategies to overcome precur-
sor limitations in isoprenoid biosynthesis. Curr Opin Biotechnol 66:171–178. https://doi.org/
10.1016/j.copbio.2020.07.005
Application of Yeast Synthetic
Biology for the Production of Citrus
Flavors
Abstract
Plants have the potential to produce an extensive variety of sesquiterpenoid
compounds being used as flavor, fragrance, and medicine. Among the sesquiter-
penoids of the citrus family, (+)-valencene and (+)-nootkatone are considered
the most valuable compounds used in various industries. Due to the low yield of
these compounds in their native hosts, commercial (+)-nootkatone is produced
from (+)-valencene by oxidation, either chemically or enzymatically. Further-
more, chemical synthesis is complex, toxic, and harmful to the environment.
Therefore, more attention should be paid to the environment-friendly and safe
methods for (+)-nootkatone synthesis. Recently, however, significant progress
has been achieved on enzymatic oxidation of (+)-valencene to (+)-nootkatone
(or to its precursor nootkatol) in the biotransformation/bioconversion process.
Nevertheless, ideally de novo (+)-nootkatone could also be produced from
(+)-valencene via biotechnological approaches in heterologous hosts using the
enzymes involved in its biosynthetic pathway. Among some of the organisms
used for the production of (+)-valencene and (+)-nootkatone, the yeast platform
has the highest potential for the production of these two valuable products. De
novo reconstitution of related-biosynthetic pathways in Saccharomyces cere-
visiae, Yarrowia lipolytica, and Pichia pastoris has led to high production of
these compounds, expressing hope to meet the needs of various industries on
large scales which will be described in the following chapter.
Keywords
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 207
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_8
208 K. Farmanpour-Kalalagh et al.
1 Introduction
Valencene is the main sesquiterpene in orange peel oil, and its titer has been
commonly applied to specify the oil’s commercial degree. To characterize the
contribution of valencene to the aroma quality of a marketable orange oil by using
multidimensional GC–O/GC–MS, thirty-seven aroma-active compounds have been
identified in orange oil with valencene concentration between 54 and 68 μg/g
(Elston et al. 2005). In addition to orange oil, valencene is reported in Chinese
bayberry (Myrica rubra) with a relative abundance of 0.27–4.25% (Fujita et al.
2014), 5.97% (Langhasova et al. 2014), and 6% (Ambrož et al. 2015), 1.24% in
Welensali (Croton flavens) (Sylvestre et al. 2006), 0.01–0.4% in mandarins culti-
vars (Merle et al. 2004), 0.1 (μL)/100 g in dehydrated lime [Citrus aurantifolia
(Christm.) Swingle] (Ramesh Yadav et al. 2004), 0.44% in Egyptian Eucalyptus
camaldulensis var. brevirostris (El-Ghorab et al. 2002).
Analyzing and comparison of essential oil components extracted from the heart-
woods of Leyland cypress, Alaska yellow cedar, and Monterey cypress via steam
distillation and solvent extraction by Liu (2009) indicated that the eighteen com-
pounds are found both in Alaska cedar and Leyland cypress although there are
only eight compounds found in Monterey cypress from steam distillation (6 and
12 h) of their heartwoods. The GC–MS analysis of volatile constituents from two
fractions of steam distillation and solvent extraction, respectively, indicates that
the main components in Alaska cedar, Monterey cypress, and Leyland cypress, are
carvacrol, nootkatene, nootkatol, nootkatone, and nootkatin, which display about
60–90% of the whole capacity of the essential oil. These ingredients are from three
diverse families which are p-Cymene, eremophilane, and tropolone (Fig. 2a and b).
The evaluation of volatile constituents from Alaska cedar lumber and fresh wood
exhibits that in parallel with the maturing and aging of the wood, the quantity of
nootkatol and epi-nootkatol decreased, whereas the content of nootkatene increases
significantly, which indirectly shows the conversion of nootkatol to nootkatene in
the lumber boosted via the increasing of acidity of the heartwood (Fig. 2c).
Many compounds are found by solvent extraction when detected with the
GC–MS. Because of the low titer of some components, only some of them are
identified. There are twelve compounds in Alaska cedar, eleven in Leyland cypress,
and only seven from Monterey cypress. Excluding two compounds, valencene-11,
12- diol and an unknown tropolone, all the others are detected in the steam distil-
lation fractions. The lacking of valencene-11, 12- diol and the unknown tropolone
are probably modified during steam distillation due to higher temperature (Fig. 2d).
Nootkatone detection in other plants such as vetiver grass oil by applying com-
prehensive two-dimensional gas chromatography techniques (GC × GC-FID/MS)
showed that the amount of this sesquiterpene is between 0.30 and 0.60 g/100 g oil
among different varieties.
Application of Yeast Synthetic Biology for the Production … 211
Fig. 2 a The components of the first 6 h distillation fraction from Alaska yellow cedar, Leyland
cypress, and Monterey cypress; b The components of the 6-12 h distillation fraction from Alaska
yellow cedar, Leyland cypress, and Monterey cypress; c The components of the 6 h distillation
fraction from Alaska yellow cedar fresh wood and lumber; d The components of the 24 h solvent
extraction from Alaska yellow cedar, Leyland cypress and, Monterey cypress analyzed by GC–MS.
(Adapted from Liu 2009)
Terpenoids, also termed isoprenoids, are known as a wide and highly diverse group
of natural products. They are regarded as one of the largest groups of secondary
metabolites with various biological functions and valuable properties for diverse
industrial applications (Kallscheuer et al. 2019). Classification of terpenoids is
based on the number of carbon atoms forming the main skeleton. The general
divisions of them are hemiterpenoids (C5), monoterpenoids (C10), sesquiter-
penoids (C15), diterpenoids (C20), sesterterpenoids (C25), triterpenoids (C30), and
tetraterpenoids (also known as carotenoids; C40) (Ashour et al. 2010; Beyraghdar
Kashkooli et al. 2018). The central hydrocarbon backbone created from isoprene
units is regularly changed via enzymes like cytochromes P450, hydrogenases,
methyltransferases, and glycosyltransferases. Totally, terpenoids content in natu-
ral sources is typically low and extraction may result in by-products (Asmund
Arnesen et al. 2020). Biosynthetically, all terpenoids are biosynthesized from
the isoprene C5 backbones isopentenyl pyrophosphate (IPP) and dimethylallyl
pyrophosphate (DMAPP), which are produced via the 2-C-methyl-D-erythritol
4-phosphate (MEP), also called 1-deoxy-D-xylulose 5-phosphate (DXP), and
214 K. Farmanpour-Kalalagh et al.
mevalonate (MVA) pathway. The MEP pathway, which commences from pyruvate
and glyceraldehyde-3-phosphate (GAP), is commonly found in plastids of bacte-
ria, cyanobacteria, green algae, and plants (Frank and Groll 2017). The cytosolic
MVA pathway mainly exists in eukaryotes (mammals, plants, and fungi), but also
archaea and a few bacteria. In the cytosol, sesquiterpenes are biosynthesized from
isoprenoid precursors by the MVA pathway. The MVA pathway starts from acetyl-
CoA and leads to the formation of isopentenyl diphosphate (IPP) and dimethyl
allyl diphosphate (DMAPP). Then, farnesyl diphosphate synthase (FPS) accumu-
lates DMAPP and IPP into farnesyl diphosphate (FPP). Terpene synthases (TPSs)
use FPP to create terpene hydrocarbons, which can be further altered via oxi-
dation into a diverse range of terpenoids (Chappell 2002). TPSs are classified
based on their substrate characteristics. Notable advances have been achieved dur-
ing the past decade in the biochemical reactions catalyzed by TPSs detail (Chappell
2004). Sesquiterpene synthases use the 15-carbon substrate farnesyl diphosphate
(FPP) (Chappell 2004). Sharon-Asa et al. (2003) have isolated and characterized a
sesquiterpene synthase (Cstps1) from Citrus that is presenting a new insight into
the biochemical process of this enzyme, and bioengineering of (sesqui)terpenoids
in native and heterologous hosts. Only a few functional (+)-valencene synthases
were reported in previous studies including VvVal from Vitis vinifera (Lücker
et al. 2004), Cstps1 from Citrus sinensis (Sharon-Asa et al. 2003), GFTpsD
from Citrus × paradisi, and CnVS from Nootka cypress (Callitropsis nootkatensis)
(Beekwilder et al. 2014).
The biosynthesis of (+)-valencene, as a precursor of (+)-nootkatone, is done
through the MVA pathway. IPP is frequently condensed via prenyl transferases to
create Farnesyl pyrophosphate (FPP). FPP is converted to (+)-valencene by the
germacrenyl carbocation. So that the formation of internal bond and deprotonating
is catalyzed by valencene synthase (Figs. 3 and 4). Not much information has been
published about the enzymatic steps of converting (+)-valencene to (+)-nootkatone.
It has been propounded that the biosynthetic pathway continues by a regioselec-
tive allylic hydroxylation of (+)-valencene to form nootkatol and then oxidized
to (+)-nootkatone. Both phases can be accelerated by the catalytic function of a
single multifunctional hydroxylase/oxidase or via a consecutive enzyme-mediated
reaction. But the reason has not been determined yet why (+)-valencene accumu-
lates in the peel of Valencia orange throughout fruit maturation while is further
oxidized to (+)-nootkatone in grapefruit (Fraatz et al. 2009a).
Fig. 3 Valencene and Nootkatone biosynthetic pathway in four major producing plants
Fig. 4 Overview of yeast synbio for production of valencene and nootkatone. ERG10, acetyl-
CoA C-acetyltransferase; ERG13, hydroxymethylglutaryl-CoA synthase; ERG12, mevalonate
kinase; tHMGR, truncated 3-hydroxy-3-methylglutaryl-CoA reductase; ERG8, phosphomeval-
onate kinase; ERG19, diphosphomevalonate decarboxylase; IDI1, isopentenyl diphosphate delta-
isomerase; ERG20, farnesyl diphosphate synthase/dimethylallyltranstransferase; VS: valencene
synthase, Cyp450: cytochrome P450, valencene oxidase
216 K. Farmanpour-Kalalagh et al.
an oxygen atom into allyl groups and are the main candidates for biotransforma-
tion of (+)-valencene into (+)-nootkatone. However, finding regioselective P450
enzymes relevant to industrial demands is still a challenge in this regard (Gavira
et al. 2013). A two-step enzymatic transformation of (+)-valencene has been sug-
gested by a regioselective allylic hydroxylation of the 2-position of (+)-valencene
to nootkatol and then by the oxidation to (+)-nootkatone. There are two hypotheses
in this case. The first is that both reactions can be enhanced by a multifunctional
cytochrome P450 enzyme. Second, the oxidative activity of cytochrome P450 leads
to the production of nootkatol, followed by an alcohol dehydrogenase activity
to produce (+)-nootkatone. Some enzymes originated from plants and microor-
ganisms catalyzing the production of either nootkatol and/or (+)-nootkatone from
(+)-valencene have been successfully identified (Cankar et al. 2014).
Cytochrome P450cam and P450BM-3 : (+)-valencene oxidation of the wild-type
and mutants of P450cam from Pseudomonas putida, and of P450BM-3 from Bacil-
lus megaterium, have been studied as a promising way towards (+)-nootkatone
production. Wild-type P450cam does not oxidize (+)-valencene but the mutants
produce (+)-trans-nootkatol and (+)-nootkatone as the major products. Ignoring
the non-selective properties, wild-type P450BM-3 and its mutant forms contain
dominant activities compared with the P450cam . Among the many compounds,
cis-(+)-valencene-1, 10-epoxide, cis- and trans-(+)-nootkatol, (+)-nootkatone, (+)-
nootkatone-13S, 14-epoxide, and trans-(+)-nootkaton-9-ol were characterized from
whole-cell reactions. The selectivity patterns express that (+)-valencene contains
a single binding direction in P450cam but several directions in P450BM-3 (Sowden
et al. 2005).
CYP from ascomycete Chaetomium globosum: Oxidizing the exogenous (+)-
valencene to nootkatone via the stereoselective production of alpha-nootkatol
can be applied in submerged cultures of the ascomycete Chaetomium globosum.
Inhibition tests propose that the insertion of the first oxygen atoms catalysis
by cytochrome P450 monooxygenase. Hence in addition to nootkatone, non-
volatile oxidation products along with flavor-active and inactive compounds can be
identified. (+)-valencene, valencene-11, 12-epoxide, alpha-nootkatol, and nootka-
tone accumulated mostly inside the C. globosum cells. On the other hand,
nootkatone-11, 12-epoxide is only in the culture medium. Therefore, active trans-
porting of compounds into the extracellular sections is done during (+)-valencene
detoxification (Kaspera et al. 2005).
CYP from algae and fungi: Studies indicate that the cytochrome P450s have
an important role in (+)-valencene biotransformation by the green algae Chlorella
species and fungi such as Mucor species, Botryosphaeria dothidea, and Botry-
odiplodia theobromae to yield nootkatone in high quantity (Furusawa et al.
2005).
CYP from Gynostemma pentaphyllum: It has been known that the suspension
cultures of Gynostemma pentaphyllum can transform valencene into nootkatone
as the major product. Also, biotransformation of valencene by Caragana cham-
lagu and Hibiscus cannabinus cultured plant cells show largely the same results
(Sakamaki et al. 2005).
218 K. Farmanpour-Kalalagh et al.
at 144 ± 10 μg/L yeast culture. CnVO is one of the CYP706 family members of
cytochrome P450 oxidases (Cankar et al. 2014).
CYP from Botryodiplodia theobromae 1368, Yarrowia lipolytica 2.2ab, and
Phanerochaete chrysosporium: It has been observed that Botryodiplodia theobro-
mae 1368, Yarrowia lipolytica 2.2ab, and Phanerochaete chrysosporium oxidize
(+)-valencene to (+)-nootkatone, with concentrations of 231.7 ± 2.1, 216.9 ± 5.8,
and 100.8 ± 2.6 mg/l (+)-nootkatone, respectively. Various biotransformation
experiments (organic, aqueous, and biphasic) have been also tested which led to
the same production levels of nootkatone (Palmerín-Carreño et al. 2015).
Peroxidase combined with a laccase from Funalia trogii: an incorporated
platform of a dye-decolorizing peroxidase (Ftr-DyP) and a laccase isolated from
the basidiomycete Funalia trogii transformed the (+)-valencene totally to the
(+)-nootkatone, with a high concentration of 1100 mg/L (Kolwek et al. 2018).
7 Conclusion
The flavor and fragrance industry has always been an attractive and rich industry
in the world. Today, meeting the needs of the people and the market is one of the
main challenges of these industries. (+)-Valencene and (+)-nootkatone are known
as natural flavors and aromas and are produced in low amounts in their native
hosts which requires a lot of raw material. On the other hand, although chem-
ical synthesis of these compounds has been successful, due to the side effects
and the fact of not being categorized as natural, this method is not very much
accepted and popular. Significant success has also been achieved in the biotrans-
formation/bioconversion method. In order to meet the market demand on a large
scale, high amounts of these compounds have been obtained via the heterolo-
gous production of (+)-valencene and (+)-nootkatone in microorganisms, including
yeasts S. cerevisiae, Y. lipolytica, and P. pastoris.
References
Ambrož M, Boušová I, Skarka A, Hanušová V, Králová V, Matoušková P, Szotáková B, Skálová
L (2015) The influence of sesquiterpenes from Myrica rubra on the antiproliferative and
pro-oxidative effects of doxorubicin and its accumulation in cancer cells. Molecules 20:15343–
15358. https://doi.org/10.3390/molecules200815343
Asadollahi MA, Maury J, Møller K, Nielsen KF, Schalk M, Clark A, Nielsen J (2008) Production
of plant sesquiterpenes in Saccharomyces cerevisiae: Effect of ERG9 repression on sesquiter-
pene biosynthesis. Biotechnol Bioeng 3:666–677. https://doi.org/10.1002/bit
Ashour M, Wink M, Gershenzon J (2010) Biochemistry of terpenoids: Monoterpenes, sesquiter-
penes and diterpenes, biochemistry of plant secondary metabolism: 2nd edition. https://doi.org/
10.1002/9781444320503.ch5
Asmund Arnesen J, Kildegaard KR, Cernuda Pastor M, Jayachandran S, Kristensen M, Borodina
I (2020) Yarrowia lipolytica strains engineered for the production of terpenoids. Front Bioeng
Biotechnol 8:1–14. https://doi.org/10.3389/fbioe.2020.00945
Beekwilder J, van Houwelingen A, Cankar K, van Dijk ADJ, de Jong RM, Stoopen G,
Bouwmeester H, Achkar J, Sonke T, Bosch D (2014) Valencene synthase from the heartwood of
Nootka cypress (Callitropsis nootkatensis) for biotechnological production of valencene. Plant
Biotechnol J 12:174–182. https://doi.org/10.1111/pbi.12124
Beyraghdar Kashkooli A (2018) Biosynthesis, transport and combinatorial metabolic engineering
of Tanacetum parthenium (feverfew) and Artemisia annua (sweet wormwood) sesquiterpene
lactones. Ph.D. Thesis, pp 1–55. https://doi.org/10.18174/455405
Beyraghdar Kashkooli A, van der Krol A, Bouwmeester H (2018) Terpenoid biosynthesis in plants.
Flavour Science, Verlag der Technischen Universität Graz, pp 3–12
Cankar K, Houwelingen AV, Bosch D, Sonke T, Bouwmeester H, Beekwilder J (2011) A chicory
cytochrome P450 mono-oxygenase CYP71AV8 for the oxidation of (+)-valencene. FEBS Lett
585:178–182. https://doi.org/10.1016/j.febslet.2010.11.040
Cankar K, Van Houwelingen A, Goedbloed M, Renirie R, De Jong RM, Bouwmeester H, Bosch D,
Sonke T, Beekwilder J (2014) Valencene oxidase CYP706M1 from Alaska cedar (Callitropsis
nootkatensis). FEBS Lett 588:1001–1007. https://doi.org/10.1016/j.febslet.2014.01.061
Cankar K, Jongedijk E, Klompmaker M, Majdic T, Mumm R, Bouwmeester H, Bosch D, Beek-
wilder J (2015) (+)-Valencene production in Nicotiana benthamiana is increased by down-
regulation of competing pathways. Biotechnol J 10:180–189. https://doi.org/10.1002/biot.201
400288
Application of Yeast Synthetic Biology for the Production … 223
Chappell J (2002) The genetics and molecular genetics of terpene and sterol origami. Curr Opin
Plant Biol 5:151–157. https://doi.org/10.1016/S1369-5266(02)00241-8
Chappell J (2004) Valencene synthase-a biochemical magician and harbinger of transgenic aromas.
Trends Plant Sci 9:265–269. https://doi.org/10.1016/j.tplants.2004.03.003
Chen H, Zhu C, Zhu M, Xiong J, Ma H, Zhuo M, Li S (2019) High production of valencene in
Saccharomyces cerevisiae through metabolic engineering. Microb Cell Fact 18:1–14. https://
doi.org/10.1186/s12934-019-1246-2
El-Ghorab AH, Fadel HM, El-Massry KF (2002) The Egyptian Eucalyptus camaldulensis var. bre-
virostris: Chemical compositions of the fruit volatile oil and antioxidant activity. Flavour Fragr
J 17:306–312. https://doi.org/10.1002/ffj.1085
Elston A, Lin J, Rouseff R (2005) Determination of the role of valencene in orange oil as a direct
contributor to aroma quality. Flavour Fragr J 20:381–386. https://doi.org/10.1002/ffj.1578
Emmerstorfer A, Wimmer-Teubenbacher M, Wriessnegger T, Leitner E, Müller M, Kaluzna I,
Schürmann M, Mink D, Zellnig G, Schwab H, Pichler H (2015) Over-expression of ICE2 stabi-
lizes cytochrome P450 reductase in Saccharomyces cerevisiae and Pichia pastoris. Biotechnol
J 10:623–635. https://doi.org/10.1002/biot.201400780
Emmerstorfer-Augustin A, Pichler H (2016) Production of aromatic plant terpenoids in recombi-
nant baker’s yeast, in: Biotechnology of plant secondary metabolism: methods and protocols,
methods in molecular biology. pp 79–89. https://doi.org/10.1007/978-1-4939-3393-8_8
Filippi JJ, Belhassen E, Baldovini N, Brevard H, Meierhenrich UJ (2013) Qualitative and quanti-
tative analysis of vetiver essential oils by comprehensive two-dimensional gas chromatography
and comprehensive two-dimensional gas chromatography/mass spectrometry. J Chromatogr
1288:127–148. https://doi.org/10.1016/j.chroma.2013.03.002
Fraatz MA, Berger RG, Zorn H (2009a) Nootkatone-a biotechnological challenge. Appl Microbiol
Biotechnol 83:35–41. https://doi.org/10.1007/s00253-009-1968-x
Fraatz MA, Riemer SJL, Stöber R, Kaspera R, Nimtz M, Berger RG, Zorn H (2009b) A novel
oxygenase from Pleurotus sapidus transforms valencene to nootkatone. J Mol Catal B Enzym
61:202–207. https://doi.org/10.1016/j.molcatb.2009.07.001
Frank A, Groll M (2017) The Methylerythritol phosphate pathway to isoprenoids. Chem Rev
117:5675–5703. https://doi.org/10.1021/acs.chemrev.6b00537
Frohwitter J, Heider SAE, Peters-Wendisch P, Beekwilder J, Wendisch VF (2014) Production of
the sesquiterpene (+)-valencene by metabolically engineered Corynebacterium glutamicum. J
Biotechnol 191:205–213. https://doi.org/10.1016/j.jbiotec.2014.05.032
Fujita S, Kajiyama K, Takabayashi M (2014) Volatile constituents of the leaf and fruit of Myrica
rubra. J Nagoya Gakuin Univ Humanit Nat Sci 50, 25–33. http://doi.org/10.15012/00000352
Furusawa M, Hashimoto T, Noma Y, Asakawa Y (2005) Highly efficient production of nootkatone,
the grapefruit aroma from valencene, by biotransformation. Chem Pharm Bull 53:1513–1514.
https://doi.org/10.1248/cpb.53.1513
Gavira C, Höfer R, Lesot A, Lambert F, Zucca J, Werck-Reichhart D (2013) Challenges and pit-
falls of P450-dependent (+)-valencene bioconversion by Saccharomyces cerevisiae. Metab Eng
18:25–35. https://doi.org/10.1016/j.ymben.2013.02.003
Girhard M, Machida K, Itoh M, Schmid RD, Arisawa A, Urlacher VB (2009) Regioselective
biooxidation of (+)-valencene by recombinant E. coli expressing CYP109B1 from Bacillus sub-
tilis in a two-liquid-phase system. Microb Cell Fact 8:1–12. https://doi.org/10.1186/1475-2859-
8-36
Guo X, Sun J, Li D, Lu W (2018) Heterologous biosynthesis of (+)-nootkatone in unconven-
tional yeast Yarrowia lipolytica. Biochem Eng J 137:125–131. https://doi.org/10.1016/j.bej.
2018.05.023
Hung LVM, Moon JY, Ryu JY, Cho SK (2019) Nootkatone, an AMPK activator derived from
grapefruit, inhibits KRAS downstream pathway and sensitizes non-small-cell lung cancer A549
cells to adriamycin. Phytomedicine 63:153000. https://doi.org/10.1016/j.phymed.2019.153000
Jaiswal Y, Liang Z, Guo P, Ho HM, Chen H, Zhao Z (2014) Tissue-specific metabolite profiling
of Cyperus rotundus L. rhizomes and (+)-nootkatone quantitation by laser microdissection,
ultra-high-performance liquid chromatography-quadrupole time-of-flight mass spectrometry,
224 K. Farmanpour-Kalalagh et al.
Scholtmeijer K, Cankar K, Beekwilder J, Wösten HAB, Lugones LG, Bosch D (2014) Production
of (+)-valencene in the mushroom-forming fungus S. commune. Appl Microbiol Biotechnol
98:5059–5068. https://doi.org/10.1007/s00253-014-5581-2
Sharon-Asa L, Shalit M, Frydman A, Bar E, Holland D, Or E, Lavi U, Lewinsohn E, Eyal Y (2003)
Citrus fruit flavor and aroma biosynthesis: Isolation, functional characterization, and develop-
mental regulation of Cstps1, a key gene in the production of the sesquiterpene aroma compound
valencene. Plant J 36:664–674. https://doi.org/10.1046/j.1365-313X.2003.01910.x
Sowden RJ, Yasmin S, Rees NH, Bell SG, Wong LL (2005) Biotransformation of the sesquiterpene
(+)-valencene by cytochrome P450cam and P450BM-3 . Org Biomol Chem 3:57–64. https://doi.
org/10.1039/b413068e
Šunjerga A (2019) Antiproliferativno djelovanje nootkatona na dvije stanične linije raka. Under-
graded Thesis, pp 1–43. https://urn.nsk.hr/urn:nbn:hr:167:215698
Sylvestre M, Pichette A, Longtin A, Nagau F, Legault J (2006) Essential oil analysis and anticancer
activity of leaf essential oil of Croton flavens L. from Guadeloupe. J Ethnopharmacol 103:99–
102. https://doi.org/10.1016/j.jep.2005.07.011
Takahashi S, Yeo YS, Zhao Y, O’Maille PE, Greenhagen BT, Noel JP, Coates RM, Chappell J
(2007) Functional characterization of premnaspirodiene oxygenase, a cytochrome P450 cat-
alyzing regio- and stereo-specific hydroxylations of diverse sesquiterpene substrates. J Biol
Chem 282:31744–31754. https://doi.org/10.1074/jbc.M703378200
Tocmo R, Pena-Fronteras J, Calumba KF, Mendoza M, Johnson JJ (2020) Valorization of pomelo
(Citrus grandis Osbeck) peel: A review of current utilization, phytochemistry, bioactivities, and
mechanisms of action. Compr Rev Food Sci Food Saf 19:1969–2012. https://doi.org/10.1111/
1541-4337.12561
Troost K, Loeschcke A, Hilgers F, Özgür AY, Weber TM, Santiago-Schübel B, Svensson V, Hage-
Hülsmann J, Habash SS, Grundler FMW, Schleker ASS, Jaeger KE, Drepper T (2019) Engi-
neered Rhodobacter capsulatus as a phototrophic platform organism for the synthesis of plant
sesquiterpenoids. Front Microbiol 10:1–14. https://doi.org/10.3389/fmicb.2019.01998
Tsoyi K, Jang HJ, Lee YS, Kim YM, Kim HJ, Seo HG, Lee JH, Kwak JH, Lee DU, Chang KC
(2011) (+)-Nootkatone and (+)-valencene from rhizomes of Cyperus rotundus increase survival
rates in septic mice due to heme oxygenase-1 induction. J Ethnopharmacol 137:1311–1317.
https://doi.org/10.1016/j.jep.2011.07.062
Waltz E (2020) Specter of eye toxicity looms over BCMA-targeted therapy. Nat Biotechnol
38:1363–1365. https://doi.org/10.1038/s41587-020-00757-8
Wang Y, Wang M, Xu M, Li T, Fan K, Yan T, Xiao F, Bi K, Jia Y (2018) Nootkatone,
a neuroprotective agent from Alpiniae oxyphyllae Fructus, improves cognitive impairment
in lipopolysaccharide-induced mouse model of Alzheimer’s disease. Int Immunopharmacol
62:77–85. https://doi.org/10.1016/j.intimp.2018.06.042
Wilson CW, Shaw PE (1978) Synthesis of nootkatone from valencene. J Agric Food Chem
26:1430–1432. https://doi.org/10.1021/jf60220a054
Wriessnegger T, Augustin P, Engleder M, Leitner E, Müller M, Kaluzna I, Schürmann M, Mink
D, Zellnig G, Schwab H, Pichler H (2014) Production of the sesquiterpenoid (+)-nootkatone
by metabolic engineering of Pichia pastoris. Metab Eng 24:18–29. https://doi.org/10.1016/j.
ymben.2014.04.001
Xie J, Sun B, Wang S, Ito Y (2009) Isolation and purification of nootkatone from the essential oil of
fruits of Alpinia oxyphylla Miquel by high-speed counter-current chromatography. Food Chem
117:375–380. https://doi.org/10.1016/j.foodchem.2009.04.011
Yamaguchi T (2019) Antibacterial properties of nootkatone against gram-positive bacteria. Nat
Prod Commun 14:1–5. https://doi.org/10.1177/1934578X19859999
Zelena K, Krings U, Berger RG (2012) Functional expression of a valencene dioxygenase from
Pleurotus sapidus in E. coli. Bioresour Technol 108:231–239. https://doi.org/10.1016/j.bio
rtech.2011.12.097
Zviely M (2009) Molecule of the month: Nootkatone. Perfum Flavorist 34:20–22. https://doi.org/
10.2174/1568026611107011301
Synthesis of Polyols and Organic
Acids by Wild-Type and Metabolically
Engineered Yarrowia lipolytica Strains
Abstract
In the yeast Yarrowia lipolytica, sugar polyols and organic acids are derived
from central metabolism, namely the citrate cycle and the pentose phosphate
pathway. Although these metabolites have numerous applications in agro-food,
chemical, and pharmaceutical industries, the main challenge is to gain produc-
tivity to obtain processes that are economically viable. Y. lipolytica is known
for its ability to use industrial wastes or raw materials as feedstock and to grow
at high cell density in large-scale bioreactor. Recent advances in metabolic
engineering and synthetic biology allowed the development of efficient Y.
lipolytica-based cell factories to bioconvert these feedstocks into added-value
metabolites. This book chapter will focus on current knowledge on the synthesis
of the most important polyols and organic acids in Y. lipolytica.
C. Li · W. Lin
Shenzhen Branch, Guangdong Laboratory for Lingnan Modern Agriculture, Shenzhen Key
Laboratory of Agricultural Synthetic Biology, Genome Analysis Laboratory of the Ministry of
Agriculture and Rural Affairs, Agricultural Genomics Institute at Shenzhen, Chinese Academy of
Agricultural Sciences, Shenzhen, China
W. Lin
School of Life Sciences, Henan University, Kaifeng 47500, China
K. L. Ong · J. Mou · C. S. K. Lin
School of Energy and Environment, City University of Hong Kong, Kowloon Tong, Hong Kong
P. Fickers (B)
Microbial Process and Interaction, TERRA Teaching and Research Centre, University of
Liege – Gembloux Agro-Bio Tech, Gembloux, Belgium
e-mail: [email protected]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 227
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_9
228 C. Li et al.
1 Introduction
Several polyols and organic acids have been included in the top 12 platform
chemical list of the US Department of Energy (Werpy and Petersen 2004). This
highlights their importance in nowadays industries. Historically, most of these
compounds have been produced industrially by chemical synthesis in a fossil
fuel-dependent fashion. For instance, succinic acid (SA) could be produced by
the catalytic hydrogenation of maleic anhydride (Ong et al. 2020) while erythri-
tol, four-carbon sugar alcohol, could be obtained from dialdehyde starch in the
presence of nickel as a catalyst at high temperatures (Carly and Fickers 2018).
In microorganisms, these compounds are also intermediates of biochemical
pathways, such as SA that is involved in the Krebs cycle, or cell metabolites
such as erythritol that is produced by certain osmotolerant yeasts in response to
osmotic stress (Carly and Fickers 2018). Actually, there is an increasing demand
from global drivers for greener and more sustainable processes to be developed in
the context of circular economy. With the advent of molecular technologies such
as metabolic engineering and synthetic biology, industrial microorganisms could
now be engineered to bioconvert organic and industrial wastes into value-added
chemicals using the so-called biorefinery concept. In this book chapter, we will
focus on the non-conventional yeast Yarrowia lipolytica as a cell factory and detail
recent advances made in strain and process engineering to produce polyols and
organic acid of biotechnological interest.
2 Yarrowia lipolytica
Sassi et al. 2016; Trassaert et al. 2017; Park et al. 2019), DNA assembly meth-
ods (Wong et al. 2017; Vandermies et al. 2017; Celińska et al. 2017; Larroude
et al. 2019), gene disruption systems (Fickers et al. 2003), and genome editing
tools including CRISPR/Cas9 (Larroude et al. 2020). All these technologies have
been recently reviewed (Abdel-Mawgoud et al. 2018; Larroude et al. 2018; Bilal
et al. 2020). Efficient strategies have been also developed to cultivate Y. lipolytica
in bioreactors with the aim to maximize recombinant protein or metabolite pro-
duction titers (Do et al. 2019; Vandermies and Fickers 2019). Also, review papers
related to Y. lipolytica physiology have been published (Barth and Gaillardin 1996;
Fickers et al. 2005; Nicaud 2012).
3 Organic Acid
Citric acid (CA) is a tricarboxylic organic acid that plays a central role in the
metabolism of all aerobic organisms. It is an intermediate of the Krebs cycle
(Fig. 1). Based on its non-toxic, pH buffering, and chelating properties, CA has
become an important industrial product. Applications for CA include the food,
cosmetic, and pharmaceutical industries, where it is used as a flavoring agent,
antioxidant, preservative, or pH buffering system. About 70% of the global total
CA production is used by the food industry, 20% by detergent, and the remaining
10% by the chemical and pharmaceutical industries (Carsanba et al. 2019). The
global CA production exceeded 2 million tons in 2018. The market is projected
to reach a volume of nearly 3 million tons by 2024, with an expected Compound
Annual Growth Rate (CAGR) of 4% between 2019 and 2024 (Fickers et al. 2020;
Ciriminna et al. 2017). Historically, CA was isolated from citrus fruits, but this
method was displaced by fermentation of sugars using the filamentous fungus
Aspergillus niger. In bench-top bioreactors, a production titer of 130 kg/m3 could
be obtained after five to eight days of batch cultivation (Ciriminna et al. 2017).
Because of its high productivity, A. niger dominated the market for CA production.
However, the processes developed for CA production from molasses are multi-
stage, are not environmentally friendly, and are being limited by the raw material
sources used (Morgunov et al. 2013).
230 C. Li et al.
Table 1 (continued)
Y. lipolytica strain Genetic Medium Organic acid, References
modification production
(g/L)
PGC202 Ylshd5, ach; Glycerol 110.7, SA Cui et al. (2017)
PCKEO,
SCS2EO
Mix food waste 31.7, SA Li et al. (2016)
PO1f-CAD CADEO Glucose 0.033, IA Blazeck et al.
(2015)
PCM-ACO-MFS MTT EO , Glucose 22, IA Zhao et al.
CADEO (2019)
H355 WT N-alkanes 195, KG Otten et al.
(2015)
H355A IDH1EO , Raw glycerol 186, KG Chin et al.
PYC1EO (2015)
VKM Y -2412 WT Ethanol 172, KG Chernyavskayaá
et al. (2000)
Rapeseed oil 102.5, KG Yin et al. (2012)
RoPYC2 RoPYC2EO Glycerol 62.5, KG Kamzolova et al.
(2012)
Abbreviations: OE: gene overexpression, WT: wild-type, MT: mutant strain (obtained by chem-
ical and/or UV mutagenesis), ME: evolved strain, CA: citric acid, ICA: iso-citric acid, SA: suc-
cinic acid, IA: itaconic acid, KG: ketoglutarate. GUT1: glycerokinase, GUT2: glyceraldehyde-3-
phosphate dehydrogenase, ACL1: ATP citrate lyase, ICL: isocitrate lyase, YHM2: mitochondrial
citrate carrier, YIAMPD: adenosine monophosphate deaminase, ACO1: aconitase, Ylshd5: succi-
nate dehydrogenase subunit 5, ACH: acetyl-coenzyme A hydrolase, PCK: phosphoenolpyruvate
carboxykinase, SCS2: succinyl CoA synthetase, CAD:cis-aconitate decarboxylase, MTT: mito-
chondrial cis-aconitate transporter, IDH1: isocitrate dehydrogenase, PYC1: pyruvate carboxylase,
RoPYC2: pyruvate carboxylase
Fig. 1 Overview of the principal metabolic pathways for organic acid synthesis in Y. lipolytica.
GK: glycerokinase; GDH: Glyceraldehyde-3-phosphate dehydrogenase; PCK: Phosphoenolpyru-
vate carboxykinase; PYC: pyruvate carboxylase; PDC: pyruvate decarboxylase; ACS: acetyl-
coenzyme A synthetase; ACH: acetyl-coenzyme A hydrolase; CIT: citrate synthetase; ACO: aconi-
tase; IDH: isocitrate dehydrogenase; KGDH: αketoglutarate dehydrogenase; SCS: succinyl CoA
synthetase; SDH: succinate dehydrogenase; FUM: fumarase; ACL: ATP citrate lyase; ICL: isoci-
trate lyase; MLS: malate synthase; PEP: phosphoenol pyruvate; CAD: cis-aconitate decarboxylase
GUT2 could be crucial for CA production (Mironczuk et al. 2016). With strain
Y. lipolytica AJD pADUTGut1/2, CA titer of 63.9 g/L was obtained at pH 3.0
which corresponds to a 14.5-fold increase as compared to the wild-type strain
A101 under the same experimental conditions. On glycerol-based medium, a CA
titer of 93 g/L was obtained at pH 6.0 (Mironczuk et al. 2016). The purification
of glycerol is an expensive process, leading to CA production process unprof-
itable. As an economical alternative, crude glycerol, which is a direct byproduct
of biodiesel production or saponification could be used. However, the utilization
of such substrates could be limited by their high content of impurities such as
salts, sodium hydroxide, methanol, or other organic compounds, which could be
toxic for yeasts or hinder the production process. Despite these facts, Y. lipolytica
can easily grow on such substrates and was found to produce CA with titer of
76 g/L at pH 3 (Rzechonek et al. 2018a). Overexpression of gene ICL1 encod-
ing iso-citrate lyase led to a decrease in ICA production from 10–12 to 3–6%
(Forster et al. 2007b). However, the overexpression of ICL1 did not influence the
total production of CA and ICA. When ICL1 overexpression was coupled with the
disruption of ACL1 gene (YALI0E34793g) encoding ATP citrate lyase and over-
expression of the K. marxianus INU1 gene encoding inulinase, CA and ICA titers
of 84.0 g/L and 1.8 g/L were obtained, respectively, from inulin (100 g/L) after
214 h of culture in a 2L scale bioreactor (Liu et al. 2013). Besides this, overex-
pression of genes encoding mitochondrial citrate carrier (YHM2, YALI0B10736g)
and adenosine monophosphate deaminase (AMDP, YALI0E11495g) in Y. lipolyt-
ica wild-type strain W29, a CA titer with productivity and yield from glucose of
97.1 g/L, 0.93 g/(L h) and 0.5 g/g were obtained, respectively, during culture in
bioreactor (Yuzbasheva et al. 2019).
The CA production by different Y. lipolytica wild-type strains has been also
reported. With an acetate-negative mutant strain Y. lipolytica A101-1.22 grown
on glycerol-containing waste from the biodiesel industry, a CA titer of 124.2 g/L
was obtained in repeated batch culture (Rymowicz et al. 2010). From a screening
study, strain K57 was found able to produce CA with titer of 72.1 g/L and yield
of 0.77 g/g on glucose-based medium during batch culture (Carsanba et al. 2019).
Industrial wastes have been also considered as possible carbon substrates for
CA production (Wojtatowicz et al. 1991; Imandi et al. 2008; Papanikolaou et al.
2008; Morgunov et al. 2013; Liu et al. 2014; Morgunov and Kamzolova 2015).
Using mutant strain Y. lipolytica NG40/UV7 and glycerol-containing wastes, CA
production was found to increase by 40.6% (122.2 g/L) as compared to that
obtained with pure glycerol (Morgunov et al. 2013; Morgunov and Kamzolova
2015). Pineapple wastes were also used successfully with a production titer of
202 g/kg of waste in optimized conditions (namely, 0.34% yeast extract, 70.71%
moisture content of the substrate, 0.64% KH2 PO4, and 0.69% Na2 HPO4 ) (Imandi
et al. 2008). Olive-mill wastewater (OMW) blended with glucose was used also
for CA production in nitrogen-limited conditions. The titer obtained was 28.9 g/L
(Papanikolaou et al. 2008). With Y. lipolytica SWJ-1b grown in the presence of
waste cooking oil (80 g/L), a CA titer of 31.7 g/L was obtained within 336 h
of culture in a 10 L bioreactor (Liu et al. 2014). The highest CA concentration,
234 C. Li et al.
100 g/L, was obtained by the mutant strain A-101–1.14 following the shortest
growth (24 h) and production (80 h) phases from potato starch (Wojtatowicz et al.
1991).
Iso-citric acid (ICA) exists in the form of four stereoisomers, of which only the
threo-Ds-form is an intermediate of Krebs cycle (Kamzolova et al. 2016, 2018).
A promising application of ICA is sports medicine. ICA has a marked energetic
and anti-hypoxic effect and can be used as a physiological stimulant of sports-
men undergoing intensive long-term physical training (Kamzolova et al. 2016).
Recently, ICA has been tested as a natural prophylactic and therapeutic agent. It
has been shown efficient for the treatment of iron-deficiency anemia and in the
resorption of blood clots (Morgunov et al. 2019).
At present, ICA is produced via chemical synthesis. However, this results in a
racemic mixture of stereoisomers that cannot be separated by chemical methods.
ICA can be found also in the leaves and stems of some plants from the Cras-
sulaceae family or in fruits such as blackberry and blackcurrant. However, the
purification of ICA from plant extracts or fruit juices that contain a wide range of
organic acids and other components is a complicated and very expensive techno-
logical process. Therefore, the development of biotechnological methods for ICA
production is extremely important. At present, the most promising method for ICA
production is considered to be microbiological synthesis (Kamzolova et al. 2016;
Laptev et al. 2017).
It is known that Y. lipolytica accumulates ICA and CA in the culture medium
when cell growth is nutrients limited. Moreover, the ICA/CA ratio greatly depends
on the carbon source used for cell growth. For example, wild-type strains of Y.
lipolytica secrete mainly CA and about 8–16% ICA on carbohydrates or glycerol
while approximately 50–65% CA and 35–50% ICA on substrates like alkanes,
triglycerides, ethanol, or acetate. When using ethanol as a carbon source, an ICA
proportion of 35–67% was found depending on the cultivation conditions (Kam-
zolova et al. 2016; Holz et al. 2009). Some authors have conducted several studies
to increase the ICA production titer. The recombinant Y. lipolytica H222-S4 strain
deleted for ICL1 gene, when grown on glycerol or glucose, showed only a smaller
enhancement (by 2–5%) of ICA /CA ratio. With Y. lipolytica H222-S4 T5 that
overexpress ICL1 gene, the relative ICA content in the medium was as low as
5–7% of the total acids (CA and ICA) (Forster et al. 2007a, b).
Y. lipolytica VKM Y-2723 and its mutant derivative 704-UV4-A/NG50 were
selected from 60 yeast strains for their ability for ICA production. Under optimal
culture conditions (i.e., iron concentration of 1.2 mg/L, temperature of 29 °C, pH
6.0, pO2 50–55% of saturation, and 30 mM itaconic acid), Y. lipolytica VKM
Y-2373 produced 70.6 g/L ICA and 22.4 g/L CA with an ICA/CA ratio of
1:0.32 when grown on rapeseed oil. In similar conditions, a titer of 86 g/L ICA
Synthesis of Polyols and Organic Acids by Wild-Type … 235
and 20 g/L CA with an ICA/CA ratio of 1:0.23 was obtained with Y. lipolyt-
ica 704-UV4-A/NG50 (Kamzolova et al. 2013). Under the above optimal culture
conditions, Y. lipolytica VKM Y-2723 produced 90.5 g/L ICA with a yield of
0.77 g/g (Kamzolova et al. 2018). Also, catalytic activities of enzymes such as
alcohol dehydrogenase, citrate synthase, aconitate hydratase, iso-citrate dehydro-
genase, and iso-citrate lyase but not aldehyde dehydrogenase were found higher
in a repeated-batch cultivation of 748 h than in batch cultivation. Therefore, under
optimal repeated-batch culture using ethanol as substrate, Y. lipolytica strain VKM
Y-2723 produced 109.6 g/L ICA with a production rate of 1.35 g/L/h (Morgunov
et al. 2019). Besides this, overexpression of ACO1 (YALI0D09361g) encoding
aconitase allowed to increase ICA titer (Laptev et al. 2017; Holz et al. 2009). For
the wild-type strain H222 and H222-S4 grown on sunflower oil, the ICA propor-
tion ranged between 35 and 49%, whereas it increased up to 71% with the ACO1
multi-copy transformant T1 without any modification in the total organic acid titer
(both CA and ICA). However, ICA production was only moderately increased from
8–12% up to 13–17% with carbon sources such as glucose, sucrose, and glycerol
(Holz et al. 2009). When ACO1 was expressed in multicopy, ICA and CA titers
were, respectively, of 72.6 g/L and 29.0 g/L with an ICA/CA ratio of 2.5:1 during
culture in 10L bioreactor in a rapeseed oil-based medium (Laptev et al. 2017).
(Li et al. 2017). However, it was reported that the accumulation of acetate was
a major factor that impeded the fermentation at low pH, resulting in an obvious
increase in downstream process cost (Cui et al. 2017). In order to solve this prob-
lem, the gene encoding the acetyl-coenzyme A hydrolase (ACH1) was deleted,
and PCK from S. cerevisiae encodes phosphoenolpyruvate carboxykinase together
with SCS2 encoding endogenous succinyl CoA synthetase were overexpressed in
PGC01003 to achieve the strain PGC202. As result, SA titer at 110.7 g/L with
a yield of 0.53 g/g and productivity of 0.8 g/L/h was obtained by this strain via
fed-batch fermentation with pH without control (final pH at 3.4) (Cui et al. 2017).
Nevertheless, all the metabolic evolution of Y. lipolytica for SA production has
led to a partial or total loss of its ability to grow in glucose-based medium, which
limits its industrial application (Yang et al. 2017). In this case, a strategy termed
metabolic engineering was applied by Yang et al. to obtain a glucose-consuming
Y. lipolytica (PSA02004) after a 21-day repeated fermentation of PGC01003 (Yang
et al. 2017). As a result, the evolved strain PSA02004 could produce SA at
65.7 g/L and 87.9 g/L from the YPD medium and food waste hydrolysate at pH
6.0, respectively. Interestingly, with the help of isFBB, the pH for the cultivation
of PSA02004 could be decreased to a level lower than 3.0 gradually by metabolic
evolution, and the evolved strain named Y. lipolytica PSA3.0 that could produce
SA with a titer of 19.3 g/L, productivity of 0.52 g/L/h, and yield of 0.29 g/g at
pH 3.0 from YPD was achieved (Li et al. 2018a). The enzyme activity analysis
demonstrated that the pathway from pyruvate to acetate was partially blocked in
Y. lipolytica PSA3.0 after the evolution, which is beneficial to cell growth and SA
production at low pH.
However, as for PGC01003, the accumulation of acetate is a limiting factor
for further improvement of SA production as SA recovery request a pH adjust-
ment leading to an increase of downstream process cost (Cui et al. 2017). To
solve this issue, a new recombinant strain, named PGC202, was deleted for gene
ACH1 (YALI0E30965g) encoding the acetyl-coenzyme A hydrolase and overex-
pressing PCK gene from Saccharomyces cerevisiae encoding phosphoenolpyruvate
carboxykinase together with gene SCS2 encoding endogenous succinyl CoA syn-
thetase was constructed. It allowed improving the SA production process through
the elimination of acetic acid overflow and by-products formation. Indeed, SA titer
of 110.7 g/L with a maximum yield from glycerol of 0.53 g/g and a productivity of
0.8 g/L/h were obtained during fed-batch fermentation of strain PGC202 (Cui et al.
2017). Furthermore, the strain produces 31.7 g/L SA with a yield of 0.52 g/g and
productivity of 0.60 g/L/h in isFBB fermentation when using glucose-containing
MFWs hydrolysate as the carbon source supplemented with 3% of tryptone (Li
et al. 2019).
routes are a lack of selectivity, a low yield, high risk from manipulation of harsh
chemicals, and the generation of environmental hazards. These drawbacks sharply
increased the downstream process cost and restricted the utilization of α-KG in
food, medicine, and cosmetics applications.
Y. lipolytica is unable to synthesize the pyrimidine structure of the thiamine
molecule. Thiamine is the cofactor of α-ketoglutarate dehydrogenase (KGDH)
which is a key enzyme in the metabolism of α-KG; thus, the limitation of thi-
amine availability can reduce the activity of α-ketoglutarate dehydrogenase of the
Krebs cycle. Under conditions of thiamine deficiency, the conversion of α-KG in
the Krebs cycle is inhibited, which leads to its accumulation in the fermentation
broth (Chernyavskayaá et al. 2000; Kamzolova et al. 2012; Yin et al. 2012).
Currently, different strains producing α-KG from different carbon sources have
been reported. Y. lipolytica H355 allows α-KG titer of 195 g/L with a productivity
of 1.3 g/L/h in the medium containing a mixture of n-alkanes (Guo et al. 2016b).
Under optimal conditions, Y. lipolytica VKM Y-2412 produced 172 g/L of α-KG
with a yield of 0.70 g/g and 102.5 g/L α-KG with the yield of 0.95 g/g from
ethanol and rapeseed oil, respectively (Kamzolova et al. 2012; Kamzolova and
Morgunov 2013). However, in order to overcome the disadvantages of low yield
and accumulation of byproducts, several approaches including metabolic engi-
neering strategies and different fermentation configurations have been investigated
(Guo et al. 2016a). Y. lipolytica H355A that overexpress the iso-citrate dehydroge-
nase (IDH1) and pyruvate carboxylase (PYC1) genes can produce 186 g/L α-KG
from raw glycerol. This represents a 19% increased production as compared to
the control strain H355. Y. lipolytica-RoPYC2 which overexpress RoPYC2 gene
encoding pyruvate carboxylase produced α-KG with a titer of 62.5 g/L with only
13.5 g/L pyruvate (PA). As compared to the parental strain, α-KG production in
strain RoPYC2 increased by 35.3% while PA production is reduced by 69.8%
(Yovkova et al. 2013; Yin et al. 2012).
4 Polyols
In some biological systems, polyols are obtained by the reduction of their keto or
aldo groups into hydroxyl groups (Rice et al. 2019). As polyols, Y. lipolytica syn-
thesize mainly mannitol (MAN), which is suggested to provide cofactor NADPH
for fatty acid synthesis and erythritol (EOL). The latter being produced in response
to osmotic stress. Derivatives of erythritol, namely erythrulose (EOSE) and threitol
(TOL), are also of industrial interest and will be also discussed below (Table 2).
4.1 Mannitol
also reported (Dulermo et al. 2015). MAN production has been investigated for
several Y. lipolytica strains and different experimental conditions. For cells grown
in shake-flasks in the presence of crude glycerol, mannitol was co-produced with
EOL with titers ranging from 2.6 to 14.9 g/L. During glycerol fed-batch bioreactor
culture, strains A-15 and A UV’1 synthetized MAN with a titer of, respectively,
41.4 and 38.1 g/L, corresponding to the productivity of 0.29 and 0.28 g/(L h)
(Tomaszewska et al. 2012). With a derivative of strain AIB overexpressing GUT1
gene encoding glycerol kinase grown in a batch bioreactor on a mixture of molasse
and crude glycerol, maximal MAN titer of 11 g/L was obtained (Rakicka et al.
2017a). Similarly, using a mixture of crude glycerol and olive oil mill wastewater,
the MAN titer reached 13 g/L within 140 h with a yield from glycerol of 0.21 g/g
(Sarris et al. 2019).
Synthesis of Polyols and Organic Acids by Wild-Type … 241
4.2 Erythritol
Erythritol (EOL) is a four-carbon polyol. Its industrial interest relies on its sweet-
ening property as it has the texture and taste of table sugar. EOL has also
other interesting properties regarding its application. It is not metabolized by the
human body, and thus, it is calories-free and does not modify glycemia (Carly
and Fickers 2018). It could be synthesized by several osmotolerant microorgan-
isms, including Y. liploytica and Candida magnoliae, as an osmoprotectant. In
Y. lipolytica, EOL derives from erythrose-4P, an intermediate of the pentose phos-
phate pathway (PPP) (Fig. 2). The latter is dephosphorylated by an erythrose-4P
phosphatase (E4PP) before being reduced by an erythrose reductase (ER) into
erythritol. In this yeast, genes coding for erythrose reductase have been character-
ized (namely, YALI0F18590g, YALI0D07634g, and YALI0C13508g) (Janek et al.
2017; Cheng et al. 2018). The complete EOL catabolic pathway has been also
Fig. 2 Overview of the principal metabolic pathways for polyols and derivatives synthesis in Y.
lipolytica. DHAP: dihydroxyacetone-P; E4PP: Erythrose 4P phosphatase; ER: erythrose reduc-
tase; EDH: erythritol dehydrogenase; EK: erythritol kinase; E1PI: erythrulose-1P isomerase;
E4PI: erythrulose-4P isomerase; FBA: fructose-bisphosphate aldolase; F6P-P: Fructose-6P phos-
phatase; GK: glycerol kinase; G3PDH: glycerol-3P dehydrogenase; GPI: glucose-6P isomerase;
G6PDH: glucose-6P dehydrogenase; HK: hexokinase; MDH: mannitol dehydrogenase; PFK:
phosphofructokinase; PGDH: phopshogluconate dehydrogenase; RPE: Ribulose-P3 epimerase;
R5PI: ribose-5P isomerase; TK: transketolase; TIM: triose isomerase; TAL: transaldolase; 6PGL:
6-phosphonogluconolactonase, Ss-XDH: xylitol dehydrogenase from Scheffersomyces stipites.
Pathway in italics is heterologous
242 C. Li et al.
reported recently (Niang et al. 2020). It is first converted into erythrulose by an ery-
thritol dehydrogenase (EYD1, YALI0F01650g) and subsequently phosphorylated
into L-erythrulose-1P by an erythrulose kinase (EYK1, YALI0F01606g). Then, L-
erythrulose-1P is isomerized into D-erythrulose-4P by an erythrulose-1P isomerase
(EYL1, YALI0F01584g). Finally, erythrose-4P is generated by the activity of an
erythrulose-4P isomerase (EYL2, YALI0F01628g).
Several review papers focusing on EOL have been published, and only the
most striking information will be reported below (Carly and Fickers 2018; Reg-
nat et al. 2018; Rzechonek et al. 2018b). An efficient strategy for EOL synthesis
consisted of constitutively expressing genes encoding erythrose reductase (ER).
Overexpression of ER gene YALI0F18590g in Y. lipolytica strain AMM led to
a 20% increase in EOL titer as compared to the parental strain (44.4 g/L)
(Janek et al. 2017). This corresponded to the productivity of 0.77 g/(L h) and
a yield from glycerol of 0.44 g/g. More recently, two additional ERs have been
characterized (YALI0D07634g and YALI0C13508g) (Cheng et al. 2018). Upon
overexpression of these three ERs under the control of the strong constitutive
promoter hp4d, an EOL titer of 178 g/L was obtained from an initial glucose
concentration of 300 g/L within 84 h. This corresponded to productivity and
yield of 2.1 g/(L h) and 0.59 g/g, respectively. As ERs are NADPH-dependent
enzymes, the authors engineered the co-factor metabolism by overexpressing genes
YALI0B15598g and YALI0E22694g encoding 6-phosphogluconate dehydrogenase
(GND1) and glucose-6P dehydrogenase (ZWF1), respectively. The correspond-
ing enzymes are known to generate NADPH from NADP+ . The resulting strain,
HCY108, produced EOL with a titer of 190 g/L within 80 h of culture with produc-
tivity and yield from the glucose of 2.4 g/(L h) and 0.63 g/g, respectively (Cheng
et al. 2018). Other genes from the PPP pathway have been also overexpressed
to increase EOL titer. Overexpression of gene YALI0E06479g encoding transketo-
lase (TKL1) in strains Po1d and MK1 yielded a 19% (43 g/L) and 51% (51 g/L)
increased EOL titer, respectively. This corresponded to a productivity of 0.04
and 0.05 g/(gDCW h) (Carly et al. 2017b; Mirończuk et al. 2017). By coexpres-
sion of genes YALI0E06479g (TKL1) and YALI0F15587g encoding transaldolase
(TAL1), an increase in EOL titer and productivity of 45% and 46% was obtained,
respectively (46.7 g/L and 0.5 g/(L h) (Mirończuk et al. 2017). Gene involved
in carbon source catabolism were also overexpressed to feed the PPP pathway
with precursors (i.e., fructose-6P and glyceraldehyde-3P). Overexpression of genes
YALI0F00384g (GUT1) and YALI0B02948g (GUT2) encoding, respectively, glyc-
erol kinase (GK) and glycerol-3-P dehydrogenase (G3P-DH) allowed an EOL titer
of 78 g/L from an initial glycerol concentration of 100 g/L within 72 h in a 5-L
bioreactor (Mirończuk et al. 2016). Besides this, overexpression of genes GUT1
and TKL1 in a strain disrupted for gene EYK1 allowed an erythritol titer, produc-
tivity, and yield from glycerol of 80 g/L, 1.03 g/(L h), and 0.53 g/g, respectively
(Carly et al. 2017b). By overexpressing invertase encoding gene SUC2 from Sac-
charomyces cerevisiae and native GUT1 gene, an EOL titer of 100 g/L with a
productivity and yield of 1.1 g/(L h) and 0.67 g/g, respectively, were obtained
from a mixture of raw industrial molasses (60 g/L) and crude glycerol (100 g/L)
Synthesis of Polyols and Organic Acids by Wild-Type … 243
4.3 Erythrulose
4.4 Threitol
Due to its potential for organic acids and polyols production and good tolerance to
low pH, Y. lipolytica has gradually become a cell factory for their production. How-
ever, several key steps allowing to increase strain production titer and productivity
remain to be identified. Process operations are still not optimal, especially on a
large scale. Further investigations must be made to obtain cost-effective processes
for most of the metabolites described herein.
Acknowledgements The authors would like to thank Andrew Zicler for his contribution in editing
Figs. 1 and 2.
References
Abdel-Mawgoud AM, Markham KA, Palmer CM, Liu N, Stephanopoulos G, Alper HS (2018)
Metabolic engineering in the host Yarrowia lipolytica. Metab Eng 50:192–208. https://doi.org/
10.1016/j.ymben.2018.07.016
Ahn JH, Jang YS, Lee SY (2016) Production of succinic acid by metabolically engineered microor-
ganisms. Curr Opinion Biotechnol 42:54–66. https://doi.org/10.1016/j.copbio.2016.02.034
Barth G, Gaillardin C (1996) Yarrowia lipolytica. In: Wolf K (ed) Nonconventional yeasts in
biotechnology: a handbook. Springer, Heidelberg, pp 313–388
Beauprez JJ, De Mey M, Soetaert WK (2010) Microbial succinic acid production: natural ver-
sus metabolic engineered producers. Proc Biochem 45:1103–1114. https://doi.org/10.1016/j.
procbio.2010.03.035
Bellasio M, Mattanovich D, Sauer M, Marx H (2015) Organic acids from lignocellulose: Candida
lignohabitans as a new microbial cell factory. J Ind Microbiol Biotechnol 42:681–691. https://
doi.org/10.1007/s10295-015-1590-0
Beopoulos A, Cescut J, Haddouche R, Uribelarrea JL, Molina-Jouve C, Nicaud JM (2009)
Yarrowia lipolytica as a model for bio-oil production. Prog Lipid Res 48:375–387. https://doi.
org/10.1016/j.plipres.2009.08.005
Beopoulos A, Chardot T, Nicaud JM (2009) Yarrowia lipolytica: a model and a tool to understand
the mechanisms implicated in lipid accumulation. Biochimie 91:692–696. https://doi.org/10.
1016/j.biochi.2009.02.004
Bilal M, Xu S, Iqbal HMN, Cheng H (2020) Yarrowia lipolytica as an emerging biotechnologi-
cal chassis for functional sugars biosynthesis. Crit Rev Food Sci Nutr 1–18. https://doi.org/10.
1080/10408398.2020.1739000
Bio-Succinic Acid Market Size is Expected to Grow at a CAGR of 15.7% - Valuates Reports
(2020a). https://www.prnewswire.com/news-releases/bio-succinic-acid-market-size-is-exp
ected-to-grow-at-a-cagr-of-15-7---valuates-reports-301116323.html
Blazeck J, Hill A, Jamoussi M, Pan A, Miller J, Alper HS (2015) Metabolic engineering of
Yarrowia lipolytica for itaconic acid production. Metab Eng 32:66–73. https://doi.org/10.1016/
j.ymben.2015.09.005
Bondarenko PY, Fedorov AS, Sineoky SP (2016) Optimization of repeated-batch fermentation of
a recombinant strain of the yeast Yarrowia lipolytica for succinic acid production at low pH.
Appl Biochem Microbiol 53:882–887. https://doi.org/10.1134/s0003683817090022
Carly F, Fickers P (2018) Erythritol production by yeasts: a snapshot of current knowledge. Yeast
35:455–463. https://doi.org/10.1002/yea.3306
Carly F, Gamboa-Melendez H, Vandermies M, Damblon C, Nicaud JM, Fickers P (2017) Identifi-
cation and characterization of EYK1, a key gene for erythritol catabolism in Yarrowia lipolytica.
Appl Microbiol Biotechnol 101:6587–6596. https://doi.org/10.1007/s00253-017-8361-y
Carly F, Steels S, Telek S, Vandermies M, Nicaud J-M, Fickers P (2018) Identification and char-
acterization of EYD1, encoding an erythritol dehydrogenase in Yarrowia lipolytica and its
Synthesis of Polyols and Organic Acids by Wild-Type … 245
Fickers P, Le Dall MT, Gaillardin C, Ph, Thonart, Nicaud JM (2003) New disruption cassettes for
rapid gene disruption and marker rescue in the yeast Yarrowia lipolytica. J Microbiol Methods
55:727–737
Forster A, Aurich A, Mauersberger S, Barth G (2007) Citric acid production from sucrose using a
recombinant strain of the yeast Yarrowia lipolytica. Appl Microbiol Biotechnol 75:1409–1417.
https://doi.org/10.1007/s00253-007-0958-0
Forster A, Jacobs K, Juretzek T, Mauersberger S, Barth G (2007) Overexpression of the ICL1
gene changes the product ratio of citric acid production by Yarrowia lipolytica. Appl Microbiol
Biotechnol 77:861–869. https://doi.org/10.1007/s00253-007-1205-4
Gao C, Yang X, Wang H, Rivero CP, Li C, Cui Z, Qi Q, Lin CSK (2016) Robust succinic acid
production from crude glycerol using engineered Yarrowia lipolytica. Biotechnol Biofuels
9:179–190. https://doi.org/10.1186/s13068-016-0597-8
Global Citric Acid Markets Report, 2011–2018 & 2019–2024 (2019) https://www.prnewswire.
com/news-releases/global-citric-acid-markets-report-2011-2018--2019-2024-300814817.html
Global Itaconic Acid (IA) Industry (2020) https://www.reportlinker.com/p03646033/Global-Ita
conic-Acid-IA-Industry.html?utm_source=PRN
Groenewald M, Boekhout T, Neuvéglise C, Gaillardin C, van Dijck P, Wyss M (2014) Yarrowia
lipolytica: Safety assessment of an oleaginous yeast with a great industrial potential. Crit Rev
Microbiol 40:187–206. https://doi.org/10.3109/1040841X.2013.770386
Guo H, Madzak C, Du G, Zhou J (2016) Mutagenesis of conserved active site residues of
dihydrolipoamide succinyltransferase enhances the accumulation of alpha-ketoglutarate in
Yarrowia lipolytica. Appl Microbiol Biotechnol 100:649–659. https://doi.org/10.1007/s00253-
015-6995-1
Guo H, Su S, Madzak C, Zhou J, Chen H, Chen G (2016) Applying pathway engineering to
enhance production of alpha-ketoglutarate in Yarrowia lipolytica. Appl Microbiol Biotechnol
100:9875–9884. https://doi.org/10.1007/s00253-016-7913-x
Holz M, Forster A, Mauersberger S, Barth G (2009) Aconitase overexpression changes the product
ratio of citric acid production by Yarrowia lipolytica. Appl Microbiol Biotechnol 81:1087–
1096. https://doi.org/10.1007/s00253-008-1725-6
Imandi SB, Bandaru VV, Somalanka SR, Bandaru SR, Garapati HR (2008) Application of statis-
tical experimental designs for the optimization of medium constituents for the production of
citric acid from pineapple waste. Bioresour Technol 99:4445–4450. https://doi.org/10.1016/j.
biortech.2007.08.071
Janek T, Dobrowolski A, Biegalska A, Mirończuk AM (2017) Characterization of erythrose reduc-
tase from Yarrowia lipolytica and its influence on erythritol synthesis. Microb Cell Fact 16:118.
https://doi.org/10.1186/s12934-017-0733-6
Jost B, Holz M, Aurich A, Barth G, Bley T, Muller RA (2014) The influence of oxygen limita-
tion for the production of succinic acid with recombinant strains of Yarrowia lipolytica. Appl
Microbiol Biotechnol 99:1675–1686. https://doi.org/10.1007/s00253-014-6252-z
Kamzolova SV, Allayarov RK, Lunina JN, Morgunov IG (2016) The effect of oxalic and itaconic
acids on threo-Ds-isocitric acid production from rapeseed oil by Yarrowia lipolytica. Bioresour
Technol 206:128–133. https://doi.org/10.1016/j.biortech.2016.01.092
Kamzolova SV, Chiglintseva MN, Lunina JN, Morgunov IG (2012) alpha-Ketoglutaric acid pro-
duction by Yarrowia lipolytica and its regulation. Appl Microbiol Biotechnol 96:783–791.
https://doi.org/10.1007/s00253-012-4222-x
Kamzolova SV, Dedyukhina EG, Samoilenko VA, Lunina JN, Puntus IF, Allayarov RL, Chiglint-
seva MN, Mironov AA, Morgunov IG (2013) Isocitric acid production from rapeseed oil by
Yarrowia lipolytica yeast. Appl Microbiol Biotechnol 97:9133–9144. https://doi.org/10.1007/
s00253-013-5182-5
Kamzolova SV, Morgunov IG (2017) Metabolic peculiarities of the citric acid overproduction from
glucose in yeasts Yarrowia lipolytica. Bioresour Technol 243:433–440. https://doi.org/10.1016/
j.biortech.2017.06.146
Synthesis of Polyols and Organic Acids by Wild-Type … 247
Kamzolova SV, Morgunov IG (2013) alpha-Ketoglutaric acid production from rapeseed oil by
Yarrowia lipolytica yeast. Appl Microbiol Biotechnol 97:5517–5525. https://doi.org/10.1007/
s00253-013-4772-6
Kamzolova SV, Shamin RV, Stepanova NN, Morgunov GI, Lunina JN, Allayarov RK, Samoilenko
VA, Morgunov IG (2018) Fermentation conditions and media optimization for isocitric acid
production from ethanol by Yarrowia lipolytica. BioMed Res Int 2018:2543210. https://doi.org/
10.1155/2018/2543210
Kuenz A, Krull S (2018) Biotechnological production of itaconic acid-things you have to know.
Appl Microbiol Biotechnol 102:3901–3914. https://doi.org/10.1007/s00253-018-8895-7
Laptev IA, Filimonova NA, Allayarov RK, Kamzolova SV, Samoilenko VA, Sineoky SP,
Morgunov IG (2017) New recombinant strains of the yeast Yarrowia lipolytica with overex-
pression of the aconitate hydratase gene for the obtainment of isocitric acid from rapeseed oil.
Appl Biochem Microbiol 52:699–704. https://doi.org/10.1134/s000368381607005x
Larroude M, Park Y-K, Soudier P, Kubiak M, Nicaud J-M, Rossignol T (2019) A modular golden
gate toolkit for Yarrowia lipolytica synthetic biology. Microb Biotechnol 12:1249–1259. https://
doi.org/10.1111/1751-7915.13427
Larroude M, Rossignol T, Nicaud J-M, Ledesma-Amaro R (2018) Synthetic biology tools for
engineering Yarrowia lipolytica. Biotechnol Adv 36:2150–2164. https://doi.org/10.1016/j.bio
techadv.2018.10.004
Larroude M, Trabelsi H, Nicaud J-M, Rossignol T (2020) A set of Yarrowia lipolytica
CRISPR/Cas9 vectors for exploiting wild-type strain diversity. Biotechnol Lett 42:773–785.
https://doi.org/10.1007/s10529-020-02805-4
Li C, Gao S, Li X, Yang X, Lin CSK (2018) Efficient metabolic evolution of engineered Yarrowia
lipolytica for succinic acid production using a glucose-based medium in an in situ fibrous biore-
actor under low-pH condition. Biotechnol Biofuels 11:236–248. https://doi.org/10.1186/s13
068-018-1233-6
Li C, Gao S, Yang X, Lin CSK (2017) Green and sustainable succinic acid production from crude
glycerol by engineered Yarrowia lipolytica via agricultural residue based in situ fibrous bed
bioreactor. Bioresour Technol 249:612–619. https://doi.org/10.1016/j.biortech.2017.10.011
Li C, Ong KL, Yang X, Lin CSK (2019) Bio-refinery of waste streams for green and efficient
succinic acid production by engineered Yarrowia lipolytica without pH control. Chem Eng J
371:804–812. https://doi.org/10.1016/j.cej.2019.04.092
Li C, Yang X, Gao S, Chuh AH, Lin CSK (2018) Hydrolysis of fruit and vegetable waste for effi-
cient succinic acid production with engineered Yarrowia lipolytica. J Clean Prod 179:151–159.
https://doi.org/10.1016/j.jclepro.2018.01.081
Li C, Yang X, Gao S, Wang H, Lin CSK (2016) High efficiency succinic acid production from
glycerol via in situ fibrous bed bioreactor with an engineered Yarrowia lipolytica. Bioresour
Technol 225:9–16. https://doi.org/10.1016/j.biortech.2016.11.016
Liu X, Lv J, Xu J, Zhang T, Deng Y, He J (2014) Citric acid production in Yarrowia lipolytica
SWJ-1b yeast when grown on waste cooking oil. Appl Biochem Biotechnol 175:2347–2356.
https://doi.org/10.1007/s12010-014-1430-0
Liu XY, Chi Z, Liu GL, Madzak C, Chi ZM (2013) Both decrease in ACL1 gene expression and
increase in ICL1 gene expression in marine-derived yeast Yarrowia lipolytica expressing INU1
gene enhance citric acid production from inulin. Mar Biotechnol (NY) 15:26–36. https://doi.
org/10.1007/s10126-012-9452-5
McKinlay JB, Vieille C, Zeikus JG (2007) Prospects for a bio-based succinate industry. Appl
Microbiol Biotechnol 76:727–740. https://doi.org/10.1007/s00253-007-1057-y
Mirończuk AM, Biegalska A, Dobrowolski A (2017) Functional overexpression of genes involved
in erythritol synthesis in the yeast Yarrowia lipolytica. Biotechnol Biofuels 10:77. https://doi.
org/10.1186/s13068-017-0772-6
Mironczuk AM, Rzechonek DA, Biegalska A, Rakicka M, Dobrowolski A (2016) A novel strain of
Yarrowia lipolytica as a platform for value-added product synthesis from glycerol. Biotechnol
Biofuels 9:180–192. https://doi.org/10.1186/s13068-016-0593-z
248 C. Li et al.
Mirończuk AM, Rzechonek DA, Biegalska A, Rakicka M, Dobrowolski A (2016) A novel strain of
Yarrowia lipolytica as a platform for value-added product synthesis from glycerol. Biotechnol
Biofuels 9. https://doi.org/10.1186/s13068-016-0593-z
Morgunov IG, Kamzolova SV (2015) Physiologo-biochemical characteristics of citrate-producing
yeast Yarrowia lipolytica grown on glycerol-containing waste of biodiesel industry. Appl
Microbiol Biotechnol 99:6443–6450. https://doi.org/10.1007/s00253-015-6558-5
Morgunov IG, Kamzolova SV, Karpukhina OV, Bokieva SB, Inozemtsev AN (2019) Biosynthe-
sis of isocitric acid in repeated-batch culture and testing of its stress-protective activity. Appl
Microbiol Biotechnol 103:3549–3558. https://doi.org/10.1007/s00253-019-09729-8
Morgunov IG, Kamzolova SV, Lunina JN (2013) The citric acid production from raw glycerol by
Yarrowia lipolytica yeast and its regulation. Appl Microbiol Biotechnol 97:7387–7397. https://
doi.org/10.1007/s00253-013-5054-z
Niang PM, Arguelles-Arias A, Steels S, Denies O, Nicaud J-M, Fickers P (2020) In Yarrowia
lipolytica erythritol catabolism ends with erythrose phosphate. Cell Biol Int 44:651–660.
https://doi.org/10.1002/cbin.11265
Nicaud J-M (2012) Yarrowia lipolytica. Yeast 29:409–418. https://doi.org/10.1002/yea.2921
Okamoto S, Chin T, Hiratsuka K, Aso Y, Tanaka Y, Takahashi T, Ohara H (2014) Production of
itaconic acid using metabolically engineered Escherichia coli. J Gen Appl Microbiol 60:191–
197. https://doi.org/10.2323/jgam.60.191
Ong KL, Fickers P, Pang KP, Raffel Dharma P, Luk HS, Uisan K, Lin CSK (2020) Fermentation
of fruit and vegetable wastes for biobased products. In: Food industry waste: assessment and
recuperation of commodities, Kisseva M.R. and Webb C. Academic Press, pp 255–274
Ong KL, Li C, Li X, Zhang Y, Xu J, Lin CSK (2019) Co-fermentation of glucose and xylose from
sugarcane bagasse into succinic acid by Yarrowia lipolytica. Bioch Eng J 148:108–115. https://
doi.org/10.1016/j.bej.2019.05.004
Otten A, Brocker M, Bott M (2015) Metabolic engineering of Corynebacterium glutamicum for the
production of itaconate. Metab Eng 30:156–165. https://doi.org/10.1016/j.ymben.2015.06.003
Otto C, Yovkova V, Barth G (2011) Overproduction and secretion of alpha-ketoglutaric acid by
microorganisms. Appl Microbiol Biotechnol 92:689–695. https://doi.org/10.1007/s00253-011-
3597-4
Papanikolaou S, Galiotou-Panayotou M, Fakas S, Komaitis M, Aggelis G (2008) Citric acid pro-
duction by Yarrowia lipolytica cultivated on olive-mill wastewater-based media. Bioresour
Technol 99:2419–2428. https://doi.org/10.1016/j.biortech.2007.05.005
Park Y-K, Vandermies M, Soudier P, Telek S, Thomas S, Nicaud J-M, Fickers P (2019) Effi-
cient expression vectors and host strain for the production of recombinant proteins by Yarrowia
lipolytica in process conditions. Microb Cell Fact 18:167. https://doi.org/10.1186/s12934-019-
1218-6
Rakicka M, Biegalska A, Rymowicz W, Dobrowolski A, Mirończuk AM (2017) Polyol production
from waste materials by genetically modified Yarrowia lipolytica. Bioresour Technol 243:393–
399. https://doi.org/10.1016/j.biortech.2017.06.137
Rakicka M, Mirończuk AM, Tomaszewska-Hetman L, Rywińska A, Rymowicz W (2017b) An
effective method of continuous production of erythritol from glycerol by Yarrowia lipolytica
MK1. Food Technol Biotechnol 55:125–130. https://doi.org/10.17113/ftb.55.01.17.4812
Rakicka-Pustułka M, Mirończuk AM, Celińska E, Białas W, Rymowicz W (2020) Scale-up of the
erythritol production technology—process simulation and techno-economic analysis. J Clean
Prod 257:120533. https://doi.org/10.1016/j.jclepro.2020.120533
Regnat K, Mach RL, Mach-Aigner AR (2018) Erythritol as sweetener—wherefrom and whereto?
Appl Microbiol Biotechnol 102:587–595. https://doi.org/10.1007/s00253-017-8654-1
Rice T, Zannini E, Arendt EK, Coffey A (2019) A review of polyols—biotechnological production,
food applications, regulation, labeling and health effects. Crit Rev Food Sci Nutr 1–18. https://
doi.org/10.1080/10408398.2019.1625859
Rymowicz W, Fatykhova AR, Kamzolova SV, Rywinska A, Morgunov IG (2010) Citric acid pro-
duction from glycerol-containing waste of biodiesel industry by Yarrowia lipolytica in batch,
Synthesis of Polyols and Organic Acids by Wild-Type … 249
repeated batch, and cell recycle regimes. Appl Microbiol Biotechnol 87:971–979. https://doi.
org/10.1007/s00253-010-2561-z
Rzechonek DA, Dobrowolski A, Rymowicz W, Mironczuk AM (2018) Aseptic production of citric
and isocitric acid from crude glycerol by genetically modified Yarrowia lipolytica. Bioresour
Technol 271:340–344. https://doi.org/10.1016/j.biortech.2018.09.118
Rzechonek DA, Dobrowolski A, Rymowicz W, Mirończuk AM (2018) Recent advances in bio-
logical production of erythritol. Crit Rev Biotechnol 38:620–633. https://doi.org/10.1080/073
88551.2017.1380598
Sarris D, Rapti A, Papafotis N, Koutinas AA, Papanikolaou S (2019) Production of added-value
chemical compounds through bioconversions of olive-mill wastewaters blended with crude
glycerol by a Yarrowia lipolytica strain. Molecules 24:222. https://doi.org/10.3390/molecules
24020222
Sassi H, Delvigne F, Kar T, Nicaud J-M, Coq A-MC-L, Steels S, Fickers P (2016) Deciphering how
LIP2 and POX2 promoters can optimally regulate recombinant protein production in the yeast
Yarrowia lipolytica. Microb Cell Fact 15:159. https://doi.org/10.1186/s12934-016-0558-8
Sekova VY, Dergacheva DI, Isakova EP, Gessler NN, Tereshina VM, Deryabina YI (2019) Soluble
sugar and lipid readjustments in the Yarrowia lipolytica yeast at various temperatures and pH.
Metabolites 9:307. https://doi.org/10.3390/metabo9120307
Shabbir Hussain M, Gambill L, Smith S, Blenner MA (2016) Engineering promoter architecture
in oleaginous yeast Yarrowia lipolytica. ACS Synth Biol 5:213–223. https://doi.org/10.1021/
acssynbio.5b00100
Tan JP, Md. Jahim J, Wu TY, Harun S, Kim BH, Mohammad AW (2014) Insight into biomass
as a renewable carbon source for the production of succinic acid and the factors affecting the
metabolic flux toward higher succinate yield. Ind Eng Chem Res 53:16123–16134. https://doi.
org/10.1021/ie502178j
Tan Z, Zhu X, Chen J, Li Q, Zhang X (2013) Activating phosphoenolpyruvate carboxylase and
phosphoenolpyruvate carboxykinase in combination for improvement of succinate production.
App Env Microbiol 79:4838–4844. https://doi.org/10.1128/AEM.00826-13
Tomaszewska L, Rakicka M, Rymowicz W, Rywińska A (2014) A comparative study on glyc-
erol metabolism to erythritol and citric acid in Yarrowia lipolytica yeast cells. FEMS Yeast Res
14:966–976. https://doi.org/10.1111/1567-1364.12184
Tomaszewska L, Rywińska A, Gładkowski W (2012) Production of erythritol and mannitol by
Yarrowia lipolytica yeast in media containing glycerol. J Ind Microbiol Biotechnol 39:1333–
1343. https://doi.org/10.1007/s10295-012-1145-6
Trassaert M, Vandermies M, Carly F, Denies O, Thomas S, Fickers P, Nicaud J-M (2017) New
inducible promoter for gene expression and synthetic biology in Yarrowia lipolytica. Microb
Cell Fact 16:141. https://doi.org/10.1186/s12934-017-0755-0
Vandermies M, Denies O, Nicaud JM, celin P (2017) EYK1 encoding erythrulose kinase as a
catabolic selectable marker for genome editing in the non-conventional yeast Yarrowia lipoly-
tica. J Microbiol Methods 139
Vandermies M, Fickers P (2019) Bioreactor-scale strategies for the production of recombinant pro-
tein in the yeast Yarrowia lipolytica. Microorganisms 7:40. https://doi.org/10.3390/microorga
nisms7020040
Wang N, Chi P, Zou Y, Xu Y, Xu S, Bilal M, Fickers P, Cheng H (2020) Metabolic engineering
of Yarrowia lipolytica for thermoresistance and enhanced erythritol productivity. Biotechnol
Biofuels 13:176. https://doi.org/10.1186/s13068-020-01815-8
Werpy T, Petersen G (2004) Top value-added chemicals from biomass: volume I—results of
screening for potential candidates from sugars and synthesis gas United States: N. p., 2004.
Web. https://doi.org/10.2172/15008859.
Wojtatowicz M, Rymowicz W, Kautola H (1991) Comparison of different strains of the yeast
Yarrowia lipolytica for citric acid production from glucose hydrolysate. Appl Biochem Biotech-
nol 31:165–174
250 C. Li et al.
Wong L, Engel J, Jin E, Holdridge B, Xu P (2017) YaliBricks, a versatile genetic toolkit for stream-
lined and rapid pathway engineering in Yarrowia lipolytica. Metab Eng Commun 5:68–77.
https://doi.org/10.1016/j.meteno.2017.09.001
Yang X, Wang H, Li C, Lin CSK (2017) Restoring of glucose metabolism of engineered Yarrowia
lipolytica for succinic acid production via a simple and efficient adaptive evolution strategy.
Agri Food Chem 65:4133–4139
Yin X, Madzak C, Du G, Zhou J, Chen J (2012) Enhanced alpha-ketoglutaric acid production
in Yarrowia lipolytica WSH-Z06 by regulation of the pyruvate carboxylation pathway. Appl
Microbiol Biotechnol 96:1527–1537. https://doi.org/10.1007/s00253-012-4192-z
Yovkova V, Otto C, Aurich A, Mauersberger S, Barth G (2013) Engineering the alpha-ketoglutarate
overproduction from raw glycerol by overexpression of the genes encoding NADP+-dependent
isocitrate dehydrogenase and pyruvate carboxylase in Yarrowia lipolytica. Appl Microbiol
Biotechnol 98:2003–2013. https://doi.org/10.1007/s00253-013-5369-9
Yuzbashev TV, Yuzbasheva EY, Sobolevskaya TI, Laptev IA, Vybornaya TV, Larina AS, Matsui
K, Fukui K, Sineoky SP (2010) Production of succinic acid at low pH by a recombinant strain
of the aerobic yeast Yarrowia lipolytica. Biotechnol Bioeng 107:673–682. https://doi.org/10.
1002/bit.22859
Yuzbasheva EY, Agrimi G, Yuzbashev TV, Scarcia P, Vinogradova EB, Palmieri L, Shutov AV,
Kosikhina IM, Palmieri F, Sineoky SP (2019) The mitochondrial citrate carrier in Yarrowia
lipolytica: Its identification, characterization and functional significance for the production of
citric acid. Metab Eng 54:264–274. https://doi.org/10.1016/j.ymben.2019.05.002
Zhang M, Gu L, Cheng C, Ma J, Xin F, Liu J, Wu H, Jiang M (2018) Recent advances in microbial
production of mannitol: utilization of low-cost substrates, strain development and regulation
strategies. World J Microbiol Biotechnol 34:41. https://doi.org/10.1007/s11274-018-2425-8
Zhao C, Cui Z, Zhao X, Zhang J, Zhang L, Tian Y, Qi Q, Liu J (2019) Enhanced itaconic acid pro-
duction in Yarrowia lipolytica via heterologous expression of a mitochondrial transporter MTT.
Appl Microbiol Biotechnol 103:2181–2192. https://doi.org/10.1007/s00253-019-09627-z
Recent Advances in Synthetic Biology
Applications of Pichia Species
Wan Sun, Yimeng Zuo, Zhanyi Yao, Jucan Gao, Zengyi Shao,
and Jiazhang Lian
Abstract
The rapid development of metabolic engineering and synthetic biology has
recently attracted great attention to exploring many non-model microbial
species. Pichia pastoris and Pichia kudriavzevii are the two representative
Pichia species, which possess uniquely desired biochemical, metabolic, and
physiological features favoring large-scale industrial production. This review
begins with illustrating synthetic biology parts and tools that have been devel-
oped for manipulating these two species, followed by summarizing their
applications in biomanufacturing as hosts to produce recombinant proteins, bulk
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 251
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_10
252 W. Sun et al.
chemicals, and natural products, and concludes on the challenges and potential
strategies for advancing them for broader biotechnological applications.
1 Introduction
With an increasing demand for fuels and petrochemicals, in addition to the growing
food shortage and global warming, researchers have been working on engineer-
ing microbes to generate desirable products (Choi et al. 2015; Curran and Alper
2012). Model organisms such as Escherichia coli and Saccharomyces cerevisiae
have been studied extensively and engineered as cell factories for the production
of value-added chemicals and proteins (Pontrelli et al. 2018; Lian et al. 2018).
With the rapid development of next-generation sequencing and synthetic biology
tools, non-conventional organisms with unique traits (e.g., thermotolerance (Varela
et al. 2017), stress tolerance (Abdel-Mawgoud et al. 2018), special feedstock uti-
lization (Agbogbo and Coward-Kelly 2008; Gong et al. 2015; Yaegashi et al. 2017;
Yaguchi et al. 2017), and high protein secretion capacity (Li et al. 2007)) have been
explored as well-suited hosts for industrial processes.
The most renowned Pichia species is Pichia pastoris, a methylotrophic yeast
that can utilize methanol as the sole carbon source. It can grow to a very high
cell density (>100 g/L dry cell weight) (Wang et al. 2012). The medium for
growing P. pastoris is simple and inexpensive (containing only methanol or glyc-
erol, biotin, salts, and trace elements), making it ideal for large-scale industrial
production (Cereghino et al. 2002). In addition, P. pastoris can express heterol-
ogous proteins intracellularly or extracellularly at high levels whilst minimizing
secretion of endogenous proteins. It possesses the machinery required for proper
post-translational modifications (e.g., glycosylation, disulfide bond formation, and
proteolytic processing) (Cereghino and Cregg 2000). As a “generally regarded as
safe” (GRAS) yeast, it has been engineered to produce industrial enzymes and
chemicals (Zhu 2019), therapeutic agents such as vaccines (Wang et al. 2016) and
drugs (Yu et al. 2007), and protein-based polymers (Werten et al. 2019). In recent
years, it has also been engineered to produce a leghemoglobin protein (LegH) from
soy, to create a meat-like flavor in plant-based meat products (Fraser et al. 2018).
The well-annotated genome sequence of P. pastoris (Love et al. 2016; Sturm-
berger et al. 2016) and a few genome-scale metabolic models are also available
for engineering purposes (Ye et al. 2017; Saitua et al. 2017; Torres 2019).
This review change it to summarizes the current synthetic biology parts and
tools available for P. pastoris, including promoters, terminators, plasmids, genome-
editing tools, and signal peptides, followed by a discussion on the engineering
efforts carried out to improve protein secretion and various applications of this
yeast. In addition, we have also highlighted Pichia kudriavzevii, another Pichia
species whose acid tolerance is attractive in the production of value-added organic
acids such as succinic acid (Xiao et al. 2014), D-lactic acid (Park et al. 2018),
and itaconic acid (Sun 2020). This review concludes with future perspectives in
establishing Pichia species for biotechnological applications.
Recent Advances in Synthetic Biology Applications … 253
The strength and tunability of promoters are crucial for efficient production (Porro
et al. 2005). Strong promoters are usually preferred because of the usually low
expression levels of heterologous genes, whereas tunable promoters are desirable
when multiple heterologous genes are involved in complex pathways. Commonly
used promoters include inducible and constitutive promoters. The former allows
genes of interest to be switched on or off at different stages through induction
or repression via transcription factors. In practice, applying an inducible promoter
can increase the cell density of cultures first and then initiate heterologous protein
production. Such a separation of biomass accumulation and protein production
provides a great advantage if the accumulated intermediates or products are toxic
to the cells (Cereghino and Cregg 2000; Ahmad et al. 2014; Macauley-Patrick
et al. 2005). However, using an inducible promoter requires an extra step during
cultivation (e.g., carbon source swapping or compound supplementation), which
incurs an additional cost for large-scale industrial production. In contrast, strong
and steady expression of genes of interest mediated by constitutive promoters con-
tributes to decreasing the operational cost while enhancing the yield, which is a
commonly adopted strategy if the constitutive expression does not negatively affect
cell growth (Vogl and Glieder 2013).
Inducible promoters are usually identified from unique biochemical pathways,
and constitutive promoters generally originate from housekeeping genes. The two
most commonly used promoters in P. pastoris are the inducible PAOX1 and the
constitutive PGAP (Rajamanickam et al. 2017). PAOX1 is the promoter of the alcohol
oxidase gene AOX1, which is induced by the inexpensive carbon source, methanol.
When methanol is used as a carbon source, the methanol-induced alcohol oxidase
expression can reach 30% of the total soluble protein content (Cregg et al. 1993). It
is one of the most effective promoters that have been found for protein expression
in P. pastoris (Cregg et al. 1989; Jahic et al. 2006; Koutz et al. 1989). Heterologous
protein synthesis at levels of approximately 20 g/L was achieved two decades ago
(Hasslacher et al. 1997; Werten et al. 1999). Extensive efforts have been made to
elucidate the regulatory mechanisms of PAOX1 . The determination of its cis-acting
regulatory sequence elements has allowed researchers to engineer the promoter
by means of deletion and duplication of putative transcription factor-binding sites,
yielding a library of variants with strengths ranging from 6 to 160% of the wild-
type PAOX1 (Hartner 2008). A dozen transcription factors have been identified to
be involved in the induction of PAOX1 . This work provided various PAOX1 variants
with tunable activities by combining cis-acting elements with the basal promoter
identified based on deletion analysis.
PGAP is the promoter of the glycolytic glyceraldehyde 3-phosphate dehydro-
genase gene GAP. It remains constitutively expressed, although its activity varies
when different carbon sources are used (Waterham et al. 1997). Similar to the
engineering strategy applied to PGAP , a library of PGAP variants was constructed
254 W. Sun et al.
using mutagenesis, in which the activity varied from 0.6% to 1960% of the wild-
type PGAP (Qin et al. 2011). Several transcription factors have been examined and
suggested to play roles in the regulation of PGAP . Most of the available P. pastoris
promoters are summarized in two review articles, with the strengths benchmarked
to those of PAOX1 and PGAP (Vogl and Glieder 2013; Turkanoglu Ozcelik et al.
2019). These promoters have been much less extensively studied but are highly
desired for controlling multi-gene pathways. Most of their cis-acting regulatory
sequences and mechanisms remain unclear. Beyond the ones in the two lists, there
are also a strong native promoter (PCAT1 ) from the catalase gene in P. pastoris and
a strong heterogeneous one (PMOX ) originating from the methanol oxidase gene in
Hansenula polymorpha. The former is induced by methanol with the PCAT1 variant
P4 even stronger than PAOX1 (Nong et al. 2020); the latter was completely inac-
tivated in the presence of xylose and sorbitol but showed strong activities in the
glucose, glycerol, and methanol feeds (Mombeni 2020).
In conjunction to a promoter, a terminator (tt) also plays a critical role in reg-
ulating the expression level, mainly by influencing the mRNA stability and the
subsequent translation process (Shalgi et al. 2005). However, compared with pro-
moter engineering, much less attention has been focused on terminators. In S.
cerevisiae, hundreds of terminators have been examined and characterized, which
have been demonstrated to regulate protein expression levels over a broad range,
whereas in P. pastoris, only a few recent studies have investigated the impact of
terminators.
In P. pastoris, both endogenous and heterogeneous terminators have been used
for heterologous expression. The commonly used endogenous terminators are
mostly from either the methanol utilization pathway (i.e., AOX1tt) or housekeep-
ing genes (i.e., GAPtt), whereas heterogeneous terminators are from other yeasts
such as S. cerevisiae (e.g., CYC1tt from cytochrome C isoform 1, PRM9tt from
pheromone-regulated membrane protein 9, and VPS13tt from vacuolar protein
sorting-associated protein 13), H. polymorpha (e.g., MOXtt from methanol oxi-
dase), and Kluyveromyces lactis (e.g., LAC4tt from beta-galactosidase). To date,
the impact of different terminators on heterologous expression has been primarily
examined by evaluating the expression level under the control of PAOX1 and PGAP .
Vogl et al. examined 20 terminators, of which 15 originated from the endoge-
nous methanol assimilation pathway and the remaining five from S. cerevisiae.
These terminators provided comparable expression levels of green fluorescent pro-
tein (GFP) under the control of PAOX1 , with the lowest active terminator still
reaching 57% of the highest active terminator (Vogl et al. 2016). Prielhofer et al.
focused more on the terminators from highly expressed endogenous genes, includ-
ing many ribosomal terminators. Using CYC1tt from S. cerevisiae as a reference,
all ten terminators yielded similar GFP levels under the control of PGAP (Priel-
hofer et al. 2017). Interestingly, in both studies, the heterogeneous terminators of
genes of interest offered comparable and sometimes even higher activities than the
endogenous ones, indicating that the terminators isolated from other yeasts could
be effectively recognized in P. pastoris. Recently, Ito et al. created a catalog of 72
terminators, including 28 endogenous terminators, 41 heterogeneous terminators
Recent Advances in Synthetic Biology Applications … 255
from S. cerevisiae, and three strong synthetic terminators developed originally for
S. cerevisiae. Under the control of PGAP , these terminators resulted in a 17-fold
degree of tunability in P. pastoris (Ito et al. 2020). In these studies, AOX1tt seemed
to yield the highest activity, regardless of the promoter being used (Karbalaei et al.
2020; Weninger et al. 2016).
A recent study using Candida antarctica lipase B (CALB) as a reporter pro-
tein showed that the activities of terminators are closely associated with those
of promoters (Ramakrishnan et al. 2020). Ten terminators from the endogenous
methanol utilization pathway, glycolysis, tricarboxylic acid (TCA) cycle, and other
housekeeping genes and five terminators from S. cerevisiae were compared. Their
activities were estimated by evaluating the corresponding CALB activity under the
control of PAOX1 and PGAP . Compared to AOX1tt, three terminators led to lower
lipase activities when paired with PAOX1 but higher activities when paired with
PGAP , which suggested that the performance of terminators is not insulated from
promoter influences and may also be subjected to regulatory mechanisms as seen
in promoter studies. However, the mechanism by which individual terminators
enhance expression along with different promoters remains unclear in P. pastoris.
In addition, the terminator of dihydroxyacetone synthase (DHAStt) provided a
slightly higher CALB expression level than AOX1tt under the control of PAOX1 , but
nearly threefold higher activity under the control of PGAP . Therefore, DHAStt can
potentially serve as a strong terminator when seeking high heterologous expression
in P. pastoris. In general, terminators play a critical role in protein expression, but
more studies are needed to elucidate terminator-mediated regulatory mechanisms.
Most of the protein expression and metabolic engineering tasks in P. pastoris were
achieved through genome integration, many of which targeted the AOX1 gene via
homologous recombination (HR) (Cereghino and Cregg 2000). However, unlike S.
cerevisiae, non-homologous end joining (NHEJ) is the dominant mechanism for
repairing chromosomal double-stranded breaks (DSBs) in P. pastoris. Transfor-
mants with targeted integration have to be identified using a laborious screening
process because of the uncontrollable random integration events, including large-
scale relocation of an integration locus, off-target integration potentially affecting
cell growth, and even co-integration of the DNA elements originating from the
F-plasmid and the genome of the E. coli host used to prepare the shuttle plasmid
(Schwarzhans et al. 2016).
An alternative strategy is to use replicative plasmids, which yield higher trans-
formation efficiency and are easier to screen, although stability is sometimes an
issue in this case (Lee et al. 2005). An autonomously replicating sequence (ARS)
is a key element in replicative plasmids. The first P. pastoris-specific ARS, desig-
nated PARS1, was identified over 35 years ago and enabled the use of replicative
plasmids with high transformation efficiency (Cregg et al. 1985). Over the past
decade, other elements have been discovered that can serve as alternative ARSs
256 W. Sun et al.
single crossover may occur again on the genome between two identical target
sequences, especially when the selection pressure is removed, which results in a
loss of integrated expression cassette. To achieve a double crossover, a selection
marker-containing expression cassette flanked by two homologous arms to the tar-
get site is usually transformed, resulting in the direct replacement of the genomic
sequence located between the two homologous sequences by the expression cas-
sette. Flanking the desired genes of interest and a selectable marker with the 5’
and 3’ AOX1 sequences leads to the disruption of AOX1, thereby changing the phe-
notypic substrate utilization of P. pastoris. Lastly, the selection markers commonly
used for screening include auxotrophic genes such as HIS4, URA3, and ADE1, as
well as the genes encoding resistance to zeocin and G418 sulfate (Daly and Hearn
2005; Papakonstantinou et al. 2009).
NHEJ was repressed to a great extent, thereby heightening the role of HR in repair-
ing DSB repair and raising the target integration efficiency to approximately 100%
(Weninger et al. 2016, 2018).
In addition to repressing NHEJ, future studies should focus on enhancing HR
performance to increase editing efficiency when using CRISPR/Cas9 in P. pastoris.
For example, the overexpression of RAD family recombinases can be considered.
In S. cerevisiae, overexpression of Rad51 has been shown to elevate the integra-
tion correctness and overexpression of an engineered Rad51 variant, which has a
higher affinity to recombinase Rad54, thus significantly increasing the targeting
efficiency in S. cerevisiae (Liu et al. 2004). Furthermore, expressing Rad52 from
S. cerevisiae in Yarrowia lipolytica has been shown to increase the targeting effi-
ciency from 15 to 95% (Ji 2020). When KU70 was disrupted, certain chemicals,
such as hydroxyurea, were applied to synchronize Y. lipolytica cells to the S-phase
of the cell cycle, which is when HR has the highest activity (Jang 2018). Similar
strategies can be used to enhance the HR performance in P. pastoris.
Theoretically, any locus on a genome can serve as a potential target site for
integration, except for those related to essential genes. However, several additional
features directly affect the integration efficiency. First, a higher accessibility ren-
ders a higher chance for the Cas9/gRNA complex, and later, the donor DNA to
form a complex with the genome during the cutting and repair process, which is
especially important for the integration of large pathways. AOX1, DAS1, DAS2,
and GUT1 have been widely targeted owing to their easy accessibility (Pena et al.
2018). In addition, the nano-environment surrounding the integration locus is key
to determining the expression level and dynamics of the integrated gene. Although
a high expression level does not necessarily lead to high production, it is usu-
ally preferred by the rate-limiting step for the synthesis of a product through a
multi-step pathway.
Increasing the copy number of a gene of interest has been found to be an
efficient way to increase heterologous protein production. Accordingly, ribosomal
RNA (rRNA)-encoding loci, namely ribosomal DNA (rDNA), have been popu-
lated because of their highly repeated sequences, where several loci can be targeted
simultaneously using a single plasmid containing one gRNA design (Marx et al.
2009). In P. pastoris, each repeat consists of 25S, 5.8S, and 18S rRNA genes
arranged identically in a head-to-tail tandem array and a non-transcribed spacer
(NTS) is located between two rDNA repeats. The sequences and locations of
these rDNA repeats on the chromosomes in many yeasts have been specified for
decades. For example, Wang et al. successfully integrated ten copies of the resver-
atrol biosynthetic pathway consisting of genes from Herpetosiphon aurantiacus,
Arabidopsis thaliana, and Vitis vinifera, onto the NTS regions of O. polymor-
pha simultaneously via CRISPR/Cas9, without using a selection marker (Wang
et al. 2018). In Y. lipolytica, Luu et al. obtained an integrant with eight copies
of the genes encoding capsid proteins originating from red-spotted grouper ner-
vous necrosis virus at the 26S rRNA loci through HR (Luu et al. 2017). In P.
pastoris, high copy-number integration of human serum albumin and human super-
oxide dismutase was achieved by targeting the NTS region of the rDNA locus
Recent Advances in Synthetic Biology Applications … 259
Strains Y-11430 (CBS-7435), GS115, and X-33 are the most commonly used P.
pastoris strains. Y-11430 is a wild-type strain that was originally isolated from
California Black oak and deposited in the United States Department of Agri-
culture Culture Collection (USDA-NRRL). It is renowned because of its robust
growth rate and high activity in the methanol utilization pathway. GS115 is a his-
tidine auxotroph obtained by mutagenizing Y-11430 with nitrosoguanidine. It has
become popular for its ease of integration and screening while using HIS4 as a
selectable marker (Cregg et al. 1985; Schutter et al. 2009). X-33 is the revertant
of GS115 created by the complementation of HIS4, and zeocin or blasticidin can
still be used to select X-33 transformants carrying the antibiotic resistance gene
(Higgins et al. 1998). X-33 shares a few mutations with GS115 in genes encod-
ing cell wall biosynthesis, which enhance the secretion of membrane-associated
proteins and result in a higher transformation efficiency than other species with
thick cell walls, making X-33 a popular commercial strain for heterologous protein
expression (Brady et al. 2020).
Fig. 1 General organization of the α-mating factor (α-MF) secretion signal originating from S.
cerevisiae and applied in mediating protein secretion in P. pastoris. The pre-sequence consists of
three parts and is cleaved by signal peptides in ER. The pro-sequence is cleaved at KR site by
endoprotease Kex2 and the two EA repeats are removed by dipeptidyl aminopeptidase Ste13 in
the Golgi
Table 1 (continued)
Signal peptide Origin Target protein References
Exg P. pastoris Exg1p EGFP, CALB Liang et al.
(2013)
MF4I Codon-modified Sc α-mating Phytase (6.1 g L−1 ) Xiong et al.
factor (2005)
HFBI Trichoderma reesei class II β-galactosidase Cao et al.
hydrophobin HFBI (2017b)
Apre S. cerevisiae, preregion of β-galactosidase Cao et al.
α-factor (2017b)
SP FAEC Talaromyces stipitatus feruloyl FAEC (297 mg L−1 ) Crepin et al.
esterase (FAEC) (2003)
SP Fae-1 Neurospora crassa Fae-1 FAEC (260 mg L−1 ) Crepin et al.
feruloyl esterase (2003)
SP Bovine Bovine β-casein Xylanase He et al.
β-casein (2,755.126 IU mL−1 ) (2012)
Pptox Viral K28 preprotoxin GFP Eiden-Plach
et al. (2004)
SP pGKL pGKL killer toxin Mouse α-amylase Kato et al.
(~240 mg L−1 ) (2001)
SP DDDK PMT1-gene-inactivated P. Porcine carboxypeptidase Govindappa
pastoris DDDK protein B (CPB) and Erythrina et al. (2014)
trypsin inhibitor (ETI)
Modified signal Mouse salivary α-amylase Glucoamylase (GA, Liu et al.
peptide (MSP) (S8L) + pro-region of S. 12.619 IU mL−1 ) (2005)
cerevisiae α-MF
phytase that contains an extra non-consecutive disulfide bond showed that co-
expression of Pdi increased the phytase ApV1 thermostability, and consequently,
the production by ~12-fold compared to the expression of ApV1 alone (Navone
et al. 2021). Immunoglobulin binding protein (BiP) is another abundant chaper-
one protein that resides in the ER. Belonging to the heat shock protein Hsp70
family, it facilitates protein folding and plays an important role in the ERAD path-
way. Co-expression of BiP with an A33 single-chain antibody fragment (A33scFv)
in P. pastoris increased the ER folding capacity and resulted in an approximately
threefold increase in A33scFv secretion (Damasceno et al. 2007). Co-expression of
the chaperon gene KAR2 with different copies of the gene encoding hydrophobin
(HFBI) also resulted in increased HFBI secretion. The highest HFBI secretion with
3-copy HFBI was 22 ± 1.6-fold higher than that of the strain overexpressing only
single-copy HFBI (Sallada et al. 2019).
Apart from molecular chaperones, chaperonins have also been engineered
to facilitate protein production in P. pastoris. D-phenylglycine aminotransferase
(D-PhgAT) from Pseudomonas stutzeri ST-201 is an intracellular protein that is dif-
ficult to express in the soluble active form. Jariyachawalid et al. overexpressed this
enzyme in P. pastoris and found that most of the D-PhgAT protein was insoluble.
By co-expressing E. coli chaperonins GroEL-GroES intracellularly with D-PhgAT,
a considerable amount of soluble D-PhgAT was produced, and the activity also
increased significantly. Compared to the D-PhgAT gene expressed alone, a 14,400-
fold higher volumetric activity was achieved when ten copies of chaperonins were
co-expressed (Jariyachawalid et al. 2012). In another study, GroEL-GroES resid-
ing in the ER was co-expressed with extracellular bacterial phytase or intracellular
D-PhgAT in P. pastoris. The volumetric activity of extracellular phytase was 1.5–
2.3-fold higher than that of phytase expression alone. However, the majority of
the D-PhgAT protein was inactive and found in the insoluble protein fraction
(Summpunn et al. 2018). These results suggested that the GroEL-GroES chaperone
could potentially enhance the production of functional proteins in P. pastoris when
they are present within the same compartment. Some of the major chaperones
overexpressed in P. pastoris are summarized in Table 2.
et al. 2007; Su et al. 2010a; Dong 2013), Sed1 (Su et al. 2010b; Li et al. 2015a),
Tip1 (Jo et al. 2011), Aga2 (Jacobs et al. 2008), and Flo1 (Jiang et al. 2007), all
of which are from S. cerevisiae, as well as, Pir1 (Khasa et al. 2011; Yang et al.
2017) and Pir2 (Khasa et al. 2011) from P. pastoris. In another study, 13 endoge-
nous glycosylphosphatidylinositol-modified cell wall proteins were identified upon
screening the genome of P. pastoris GS115 (Zhang et al. 2013), three of which
were chosen as anchor proteins for displaying CALB (Wang et al. 2017). These
three anchors (i.e., GCW21, GCW51, and GCW61) have also been applied to dis-
play bacterial PETase on the surface of P. pastoris, to degrade highly crystallized
polyethylene terephthalate (PET). The turnover rate of the whole-cell biocata-
lyst displaying PETase was approximately 36-fold higher than that of the purified
PETase (Chen 2020). Another anchor protein identified in P. pastoris is Flo9. The
displayed lipase B with Flo9 showed higher thermostability at 45 °C and stability
in organic solvents (Moura 2015).
P. pastoris X-33 has also been engineered to assemble protein complexes
such as minicellulosomes on the cell surface (Ou and Cao 2014). The truncated
CipA, which contains a cellulose-binding module and two cohesin modules from
Clostridium acetobutylicum, was fused to the C-terminus of the anchor flocculation
266 W. Sun et al.
After more than 20 years of development, P. pastoris has become one of the most
popular protein expression systems that is widely used in protein preparation,
structural analysis, and functional characterization. As a GRAS microorganism
approved by the United States Food and Drug Administration, thousands of pro-
teins, including medicinal proteins (i.e., insulin, human serum albumin, hepatitis B
surface antigen, and epidermal growth factor (Weinacker et al. 2013)) and indus-
trial enzymes (i.e., mannanase, phytase, xylanase, and lipase (Rabert et al. 2013))
have been successfully expressed in P. pastoris. In addition, due to the develop-
ment of pathway assembly and genome editing tools, a growing interest has been
seen in establishing P. pastoris as a microbial cell factory to produce chemicals
and natural products.
Recent Advances in Synthetic Biology Applications … 267
Recombinant proteins can be produced using bacterial, yeast, mold, insect, plant,
and mammalian expression systems. P. pastoris is particularly attractive for the
large-scale production of recombinant proteins. It naturally contains several highly
expressed genes that encode methanol assimilation and dissimilation pathway
enzymes, enabling growth with methanol as the sole carbon and energy source
(Wegner and Harder 1987). Thus, high-level expression of target proteins can
be readily achieved using these methanol-inducible promoters. When compared
with the plant and mammalian expression systems, the P. pastoris expression sys-
tem offers the advantages of low cost, fast growth, high cell density fermentation
(HCDF), and consequently high expression levels. In contrast to prokaryotes, the
biggest advantage of P. pastoris is the capability of post-translational modifica-
tions (e.g., O- and N-glycosylation and disulfide bond formation. When compared
to S. cerevisiae, P. pastoris poses little concerns regarding over-glycosylation, and
the secretory expression level of recombinant proteins is much higher (Karbalaei
et al. 2020). Therefore, P. pastoris has been widely used to produce therapeutic
glycoproteins. Moreover, P. pastoris can efficiently secrete the target proteins into
the fermentation broth, making the downstream separation and purification pro-
cess simpler, which is a paramount variable in designing viable industrial-scale
processes.
an L1/L2-HPV-16 chimeric fragment, which was cloned into the pPICZA plasmid
for heterologous expression. The chimeric protein could be positively detected by
both L1-HPV-16 and L2-HPV-16 antibodies (Sanchooli et al. 2018). Meanwhile,
when Bredell et al. expressed the HPV-16L1/L2 chimeric protein (VLP) in P. pas-
toris KM71 (MutS ) or GS115 (Mut+ ) under a constant dissolved oxygen level
Recent Advances in Synthetic Biology Applications … 269
(DO stat) fed-batch culture supplemented with methanol, they achieved a titer of
23.61 mg/L for the chimeric protein (Bredell et al. 2018).
Although a panel of medicinal proteins has been expressed in P. pastoris, cor-
rect folding of target proteins is a major concern. Due to the lack of sufficient
chaperone factors, a considerable number of recombinant proteins cannot fold into
their correct configurations. To overcome this limitation and improve the expres-
sion level, rational design and reverse engineering strategies can be adopted to
improve the protein folding microenvironment. As mentioned above, genes related
to protein folding in the ER, such as BiP (an Hsp70 chaperone), can help secretory
proteins fold correctly. The secretion of A33scFv fragment was increased approx-
imately threefold upon overexpression of the chaperone protein BiP in P. pastoris
(Damasceno et al. 2007). Overexpression of Pdi (responsible for the formation of
disulfide bonds) in P. pastoris can increase the expression of the antibody protein
2F5Fab by 2-fold (Gasser et al. 2006). Zhang et al. tested three factors related
to protein transport from S. cerevisiae (Sec63p, Ydj1p, and Ssa1p), whose overex-
pression increased the expression of GCSF by 2.8-, 3.6-, and 6.8-fold, respectively,
in P. pastoris. Therefore, finding suitable protein mates remains a major challenge
in establishing an efficient recombinant protein expression system (Zhang et al.
2006). Gasser et al. identified new protein chaperone genes, including CUP5,
SSA4, BMH2, KIN2, SSE1, and BFR2 at the transcriptional level. The overexpres-
sion of these genes significantly enhanced the secretion of 2F5Fab antibody in P.
pastoris, with the final titer of 2F5Fab reaching up to 47.27 mg/L (Gasser et al.
2007). Stadlmayr et al. established a cDNA overexpression library in P. pastoris
and identified three new protein chaperones as the secretion-enhancing factors,
which increased the expression level of the model protein by up to 45% (Stadl-
mayr et al. 2010). Huang et al. identified six significantly upregulated genes related
to recombinant protein production using comparative proteomic analysis. In par-
ticular, the co-expression of TPX, FBA, and PGAM increased the expression level
of the reporter gene by 2.46-, 1.58-, and 1.33-fold, respectively (Huangfu et al.
2015). Noteworthy, owing to the different properties of foreign proteins, there is a
dearth of generally applicable engineering approaches. In other words, the optimal
protein chaperone or secretory factors can be different case by case and should be
evaluated individually. For example, overexpression of Pdi in P. pastoris failed to
increase the production of A33scFv antibody protein and overexpression of BiP
even reduced the yield of glucose oxidase by 10-fold (Heide et al. 2002).
the Aspergillus oryzae lipase gene (AOL) to yield the plasmid pPICZαA-AOL,
which was subsequently integrated into the genome of P. pastoris X-33. Using the
methanol feeding strategy, AOL with a specific activity of 432 U/mg was obtained
in a 5 L bioreactor (Zheng et al. 2019). To increase the yield of lipase, Zhang et al.
employed a fusion expression strategy by fusing small ubiquitin modifying pro-
tein (SUMO) with Aspergillus niger lipase (ANL) to obtain SANL. The resultant
chimeric gene was cloned into pPIC9K for heterologous expression in P. pastoris
GS115. The highest activity of SANL was ~960 U/mL in a 3 L fermenter, which
was 1.85-fold higher than that of its parent ANL (Zhang et al. 2019a).
Although the expression of recombinant proteins is mainly induced by
methanol, a co-substrate culture strategy has been found to increase the yield of
industrial enzymes. Berrios et al. cloned the Rhizopus oryzae lipase gene (ROL)
and constructed the plasmid pPICZαA-ROL for heterologous expression in P. pas-
toris X-33. The engineered strain was continuously cultured with methanol and
glycerol as co-substrates in a 1.5 L BioStatAplus bioreactor. The results showed
that using glycerol as a co-substrate at 22 and 30 °C could increase the volumet-
ric productivity of recombinant lipase and reduce the consumption of methanol
(Berrios et al. 2017). In addition, the co-substrate culture could also be applied
for the industrial production of phytase. As an animal feed additive, phytase can
decompose phytic acid and greatly reduce the input of animal feed. Li et al. engi-
neered phytase production in P. pastoris by modifying PAOX1 and the α factor
signal peptide and increasing gene copy numbers. Phytase activity as high as
2,119 U/mL with a corresponding titer of 0.75 g/L was achieved, which was 4.12-
fold higher than that of the parent strain. In a 10 L fermenter, using glycerol and
methanol as co-substrates for fed-batch fermentation, the titer and enzyme activity
of phytase could be further improved to as high as 9.58 g/L and 35,032 U/mL,
respectively (Li et al. 2015c).
Besides the traditional strategies in engineering secretion signals and modify-
ing PAOX1 promoter, genome-scale metabolic models can be employed to regulate
the metabolic fluxes from a systems perspective, to improve the expression level
of recombinant proteins. Saitua et al. employed the dynamic flux balance analy-
sis (dFBA) framework to establish a dynamic genome-scale metabolic model, to
simulate recombinant protein expression process in P. pastoris (Saitua et al. 2017).
Starting with seven state variables including glucose, biomass, and fermentation
quantity, they analyzed the kinetics of substrate assumption and distribution of
metabolic flux. On this basis, Nocon et al. optimized the dFBA algorithm and pre-
dicted gene targets (including both gain- and loss-of-function targets), to enhance
the production of recombinant proteins. Overexpression targets were identified to
reside in the pentose phosphate pathway and the TCA cycle, whereas knockout tar-
gets were found to belong to several branch points of glycolysis (Fig. 2). Five out
of the nine predicted targets were found to increase the expression level of cytoso-
lic human superoxide dismutase (hSOD). More importantly, most of the same
genetic modifications led to enhanced expression of bacterial β-glucuronidase,
indicating the general applicability of the identified metabolic engineering targets
(Nocon et al. 2014).
272 W. Sun et al.
With the rise of synthetic biology, yeast has become an important cell factory to
produce fine chemicals. Currently, S. cerevisiae is the preferred host to produce
a wide range of chemicals, including, but not limited to: glycerol, L-propanediol,
lactic acid, succinic acid, and isoprene. Comparative metabolomics indicated that
the intermediate metabolites in P. pastoris could cover more than 90% of those
in S. cerevisiae, indicating great potential for chemical production in P. pastoris
(Carnicer et al. 2012). Currently, the production of S-adenosyl-L-methionine (Chu
et al. 2013), xylitol (Louie et al. 2021), hyaluronic acid (Oliveira et al. 2016),
gluconic acid (Liu et al. 2016), and lactic acid (Lima et al. 2016) has been
Recent Advances in Synthetic Biology Applications … 273
and 14.4 mg/L, respectively. After bioprocess optimization, including the employ-
ment of a co-culture strategy and fed-batch fermentation with glycerol as the
co-substrate, the yield of Monachine J and lovastatin reached 594 mg/L and
251 mg/L, respectively. In terms of methanol conversion, Yamada et al. integrated
D-LDH into the rDNA loci of P. pastoris. Multicopy integration of the D-LDH
expression cassette was achieved following post-transformational gene amplifica-
tion and selection on gradually increasing zeocine concentrations. The optimally
engineered P. pastoris strain produced D-lactic acid with a titer of 3.48 g/L by
means of test-tube fermentation for 96 h, with methanol as the sole carbon source.
This is the first report on the establishment of P. pastoris as a microbial cell factory
for the conversion of methanol to lactic acid (Yamada et al. 2019). This study pro-
vides a basis for the application of gene integration strategy at the rDNA loci in P.
pastoris for the construction of microbial cell factories for the production of value-
added chemicals. Although research on the synthesis of chemicals from methanol
is still in its infancy and the titer is still not high enough for industrial produc-
tion, the challenges in methanol conversion should be able to be addressed using
metabolic engineering and synthetic biology approaches in the future. P. pastoris is
believed to play an increasingly important role in biorefinery and biomanufacturing
in the near future.
6-MSA GS115 NpgA, ATX PAOX1 2.2 g/L (fed-batch) Gao et al. (2013)
Citrinin GS115 pksCT, MPL1, MPL2 PAOX1 0.6 ± 0.1 mg/L Xue et al. (2017)
MPL4, MPL6
MPR7
Lovastatin GS115 LovA, LovB PAOX1 24.6 mg/L Liu et al. (2018)
LovC, LovD 250.8 mg/L (fed-batch)
LovF, LovG
(continued)
275
276
Table 5 (continued)
Nature product Host Genes Promoter Yield References
NpgA
17-Hydroxyprogesterone GS115 Human CYP17 cDNA PAOX1 Not reported Kolar et al. (2007)
Note CrtE: geranylgeranyl diphosphate synthase; CrtB: phytoene synthase; CrtI: phytoene desaturase; CrtL: lycopene β-cyclase; CrtW: xanthophylls; CrtZ:
β-carotene hydroxylase; CPR: A. thaliana cytochrome P450 reductase; HPO: premnaspirodiene oxygenase of H. muticus; ADH: alcohol dehydrogenase; trun-
cated HMG1: truncated hydroxy-methylglutaryl-CoA reductase; ERG1: 2,3-oxidosqualene synthase; ERG7: responsible for conversion of 2,3-oxidosqualene to
lanosterol; PgDDS: dammarenediol-II synthase from Panax ginseng; SaGGPPS: geranylgeranyl diphosphate synthase from the archaebacterium Sulfolobus aci-
docaldarius; HsUBIAD1: a human homologue of E. coli prenyltransferase menA and mammalian mitochondrial prenyltransferase COQ2 ; IDI: IPP isomerase;
NpgA: A. nidulans phosphopantetheinyl transferase; ATX: Aspergillus terrus 6-methylsalicylic acid synthase; pksCT: M. purpureus citrinin polyketide synthase;
MPL1: serine hydrolase; MPL2: oxygenase. MPL4: dehydrogenase. MPL6: short-chain dehydrogenase. MPL7: oxidoreductase. MPR1: a transporter. LovA: a
cytochrome P450 monooxygenase; LovB: lovastatin nonaketide synthase; LovC: enoyl-reductase; LovD: acyl-transferase; LovF: lovastatin diketide synthase;
LovG: thioesterase; Human CYP17 cDNA: catalyzing the conversion of progesterone to 17-hydroxyprogesterone as well as 16-hydroxyprogesterone
W. Sun et al.
Recent Advances in Synthetic Biology Applications … 277
Fig. 4 Reconstitution of the citrinin biosynthetic pathway in P. pastoris. CitS (pksCT ): polyketide
synthase; CitA (MPL1): serine hydrolase; CitB (MPL2): iron II oxidase; CitD (MPL4): aldehyde
dehydrogenase; CitE (MPR1): short-chain dehydrogenase
After 24 h of cultivation with methanol as the sole carbon source, citrinin was
produced up to a concentration of 0.6 mg/L (Xue et al. 2017).
able to produce D-lactic acid at a titer of 135 g/L (pH 3.6) and 154 g/L (pH 4.7)
(Park et al. 2018). Our group also engineered P. kudriavzevii YB4010 to produce
1.23 g/L itaconic acid at pH 3.9 in fed batch fermentation by overexpressing a
cis-aconitic acid decarboxylase gene from Aspergillus terreus and a native mito-
chondrial tricarboxylate transporter in the strain with the isocitrate dehydrogenase
gene deleted (Sun et al. 2020). P. kudriavzevii can also produce ethanol at a high
salt concentration (50 g/L Na2 SO4 ) at pH 2.0 or a temperature as high as 43 °C
(Isono et al. 2012). Other applications have been demonstrated in wine fermenta-
tion (Mónaco et al. 2014, 2016), production of potential probiotics (Greppi et al.
2017), biological control (Bajaj et al. 2013), and bioremediation for heavy metal
removal (Li et al. 2016b; Zhang et al. 2019b). To date, the registered P. kudriavzeii
strains are mainly diploid (Xiao et al. 2014; Xi et al. 2021), although triploid and
aneuploid strains have also been reported (Douglass 2018).
Prior to the creation of episomal plasmids, engineering works in P. kudriavzeii
were usually performed by directly transforming a linear fragment carrying the
target gene(s) and a selection marker flanked by homologous arms of the target
integration site (Park et al. 2018). URA3 is often used as a selection marker because
of the relatively easy protocol for marker recycling. Considering that the strain is
a diploid, a single-round integration will likely create a heterozygous strain, and
a second-round integration to the wild-type allele is recommended to improve
genetic stability. Recently, Tran and Cao et al. created an episomal plasmid con-
sisting of an ARS and LEU2 originating from S. cerevisiae, P. kudriavzeii URA3,
and a GFP reporter gene. The percentage of GFP+ cells in the culture grown from
a single colony was approximately 60% after 24 h (Cao et al. 2020). They iso-
lated a centromere-like (CEN-L) sequence from the P. kudriavzeii genome with the
assistance of an in silico GC3 analysis to identify the “GC3 valley” on each chro-
mosome, followed by sequence alignment to identify the conserved regions (Cao
et al. 2020). As a CEN is responsible for faithful chromosome segregation and
plays a critical role in stabilizing plasmids, the newly constructed plasmid includ-
ing CEN-L led to an increased percentage of GFP+ cells, to 81% after cultivation
for 24 h and to 67% after cultivation for 120 h.
RNA sequencing is usually implemented to identify strong constitutive promot-
ers and terminators. Cao et al. grew P. kudriavzeii under four growth conditions
(YNB medium with or without lignocellulosic biomass inhibitors under aerobic
or anaerobic conditions) and analyzed the transcriptome. Thirty-five promoters
of the most highly expressed genes identified based on RNA-sequencing data
were selected and cloned with the GFP reporter gene and TEF1 terminator on
an episomal plasmid containing ARS. Strong, medium, and weak constitutive pro-
moters were categorized based on flow cytometry. For terminator identification,
14 terminators of the strong promoters identified above were selected for further
comparison. Double reporter genes (i.e., GFP and mCherry) were placed between
the TDH3 promoter and the PGK1 terminator. Each of the candidate terminator
sequences was cloned between the GFP and the mCherry open reading frames
(ORFs) with a random sequence or no sequence inserted between the two ORFs
as controls. Quantitative PCR was used to calculate the transcriptional ratios of
280 W. Sun et al.
mCherry and GFP. Thirteen of the 14 candidate terminators had ratios below 0.03
and were categorized as strong terminators. Moreover, similar to S. cerevisiae, P.
kudriavzeii has a relatively high HR efficiency. An HR-mediated DNA assembly
method was developed to facilitate rapid plasmid construction in a single step. Co-
transforming five linear DNA fragments, with 70–80 bp overlaps designed between
the adjacent fragments, directly into I. orientalis SD108 led to the successful con-
struction of a 14.5 kb plasmid containing the xylose utilization pathway with an
assembly efficiency of 100% (Cao et al. 2020).
Genome editing tools have been developed for P. kudriavzeii. The promoter
used to transcribe the gene encoding sgRNA needs to be an RNA polymerase
(RNAP) III promoter because a typical RNAP II promoter used to transcribe
proteins will make the gene undergo post-transcriptional modifications such as
5 -end capping and 3 -end polyadenylation, which may inactivate the Cas9/gRNA
complex (Gao and Zhao 2014). Tran and Cao et al. chose a series of RNAP III
promoters including tRNALeu , tRNASer , 5S rRNA, RPR1 (the RNA component of
RNase P, the 250 bp upstream sequence of RPR1), and fusions of 5S rRNA or
RPR1’ (the 250 bp upstream sequence of RPR1 and the first 120 bp of RPR1)
with tRNALeu . An iCas9 (containing D147Y and P411T) that possesses a higher
activity than Cas9 from Streptococcus pyogenes was tagged with a nuclear local-
ization sequence and expressed by an episomal plasmid, together with each of
the sgRNA cassettes. Targeting ADE2, LEU2, HIS3, and TRP1 in I. orientalis
SD108 showed that RPR1’-tRNALeu led to the highest single-gene disruption effi-
ciency of 97–100% and was therefore used to create double and triple knockouts.
Double-gene knockouts of ADE2/TRP1 and ADE2/HIS3 were attained with effi-
ciencies of 72.8% and 89.9%, respectively, whereas triple-gene knockout efficiency
for ADE2/HIS3/SDH2 was approximately 47% (Tran et al. 2019). In parallel, our
group also carried out a similar study by evaluating ADE2 disruption efficiencies
with five versions of Cas9 (including S. pyogenes Cas9, iCas9, a codon-optimized
Cas9 version for Homo sapiens, Candida albicans, and Scheffersomyces stipitis)
and RPR1 (the 311 bp upstream sequence of RPR1) as the sgRNA promoter in
P. kudriavzevii YB4010. The highest efficiency (42%) was achieved using iCas9
(Sun et al. 2020).
Another interesting area for exploring P. kudriavzevii as a production host for
organic acids is the identification of their transporters. Previously, our group iden-
tified a mitochondrial tricarboxylate transporter, Pk_MttA, which can potentially
transport citrate and cis-aconitate from the mitochondria to the cytosol, thereby
increasing itaconic acid production in P. kudriavzevii YB4010 (Sun et al. 2020).
A recent genome sequencing and transcriptome analysis of P. kudriavzevii CY902
led to the identification of two JEN family carboxylate transporters (PkJEN2-1
and PkJEN2-2), which can import succinate into cell (Xi et al. 2021). Substrate
specificity analysis showed that both PkJEN2-1 and PkJEN2-2 are dicarboxylate
importers for succinate, L-malate, and fumarate. In addition, PkJEN2-1 can import
α-ketoglutarate, whereas PKJEN2-2 can also uptake citrate but not α-ketoglutarate.
The structural basis of PkJEN2-2 specificity toward tricarboxylate substrates was
studied using model-based structure analysis and rational design. By inactivating
Recent Advances in Synthetic Biology Applications … 281
Acknowledgements YZ, JG, and JL were supported by the National Key Research and Develop-
ment Program of China (2018YFA0901800). WS and ZS were supported by the U.S. Department
of Energy, Office of Science, Biological and Environmental Research through Ames Laboratory.
Ames Laboratory is operated for the U.S. Department of Energy by Iowa State University under
Contract No. DE-AC02-07CH11358.
References
Abdel-Mawgoud AM et al (2018) Metabolic engineering in the host Yarrowia lipolytica. Metab
Eng 50:192–208
Abdelmoula-Souissi S et al (2013) Secreted recombinant P53 protein from Pichia pastoris is a
useful antigen for detection of serum p53: autoantibody in patients with advanced colorectal
adenocarcinoma. Mol Biol Rep 40(5):3865–3872
Agbogbo FK, Coward-Kelly G (2008) Cellulosic ethanol production using the naturally occurring
xylose-fermenting yeast Pichia Stipitis. Biotechnol Lett 30(9):1515–1524
Aggarwal S, Mishra S (2020) Differential role of segments of α-mating factor secretion signal in
Pichia pastoris towards granulocyte colony-stimulating factor emerging from a wild type or
codon optimized copy of the gene. Microb Cell Fact 19(1):199
Ahmad M et al (2014) Protein expression in Pichia pastoris: recent achievements and perspectives
for heterologous protein production. Appl Microbiol Biotechnol 98(12):5301–5317
Ahn J et al (2016) Codon optimization of Saccharomyces cerevisiae mating factor alpha
prepro-leader to improve recombinant protein production in Pichia pastoris. Biotechnol Lett
38(12):2137–2143
Andreu C, Del Olmo ML (2018) Yeast arming systems: pros and cons of different protein anchors
and other elements required for display. Appl Microbiol Biotechnol 102(6):2543–2561
Recent Advances in Synthetic Biology Applications … 283
Araya-Garay JM et al (2012a) Construction of new Pichia pastoris X-33 strains for production of
lycopene and beta-carotene. Appl Microbiol Biotechnol 93(6):2483–2492
Araya-Garay JM et al (2012b) Construction of a novel Pichia pastoris strain for production of
xanthophylls. AMB Express 2(1):24
Aw R et al (2017) Expressing anti-HIV VRC01 antibody using the murine IgG1 secretion signal
in Pichia pastoris. AMB Express 7(1):70
Aza P et al (2021) Design of an improved universal signal peptide based on the α-factor mating
secretion signal for enzyme production in yeast. Cell Mol Life Sci 78(7):3691–3707
Baeshen MN et al (2016) Expression and purification of C-peptide containing insulin using Pichia
pastoris expression system. Biomed Res Int 2016:3423685
Bajaj BK, Raina S, Singh S (2013) Killer toxin from a novel killer yeast Pichia kudriavzevii RY55
with idiosyncratic antibacterial activity. J Basic Microbiol 53(8):645–656
Barrero JJ et al (2018) An improved secretion signal enhances the secretion of model proteins from
Pichia pastoris. Microb Cell Fact 17(1):161
Ben Azoun S et al (2016) Molecular optimization of rabies virus glycoprotein expression in Pichia
pastoris. Microb Biotechnol 9(3):355–368
Berrios J et al (2017) A comparative study of glycerol and sorbitol as co-substrates in methanol-
induced cultures of Pichia pastoris: temperature effect and scale-up simulation. J Ind Microbiol
Biotechnol 44(3):407–411
Brady JR et al (2020) Comparative genome-scale analysis of Pichia pastoris variants informs
selection of an optimal base strain. Biotechnol Bioeng 117(2):543–555
Brake AJ et al (1984) Alpha-factor-directed synthesis and secretion of mature foreign proteins in
Saccharomyces cerevisiae. Proc Natl Acad Sci USA 81(15):4642–4646
Bredell H et al (2018) Expression of unique chimeric human papilloma virus type 16 (HPV-16)
L1–L2 proteins in Pichia pastoris and Hansenula polymorpha. Yeast 35(9):519–529
Cankorur-Cetinkaya A et al (2018) Process development for the continuous production of heterol-
ogous proteins by the industrial yeast Komagataella Phaffii. Biotechnol Bioeng 115(12):2962–
2973
Cao M et al (2017a) Centromeric DNA facilitates nonconventional yeast genetic engineering. ACS
Synth Biol 6(8):1545–1553
Cao L et al (2017b) Improving the secretion yield of the β-galactosidase Bgal1-3 in Pichia pastoris
for use as a potential catalyst in the production of prebiotic-enriched milk. J Agric Food Chem
65(49):10757–10766
Cao M et al (2020) A genetic toolbox for metabolic engineering of Issatchenkia orientalis. Metab
Eng 59:87–97
Carnicer M et al (2012) Development of quantitative metabolomics for Pichia pastoris.
Metabolomics 8(2):284–298
Caspeta L et al (2012) Genome-scale metabolic reconstructions of Pichia stipitis and Pichia
pastoris and in silico evaluation of their potentials. BMC Syst Biol 6:24
Cereghino JL, Cregg JM (2000) Heterologous protein expression in the methylotrophic yeast
Pichia pastoris. FEMS Microbiol Rev 24(1):45–66
Cereghino GP et al (2002) Production of recombinant proteins in fermenter cultures of the yeast
Pichia pastoris. Curr Opin Biotechnol 13(4):329–332
Chahal S et al (2017) Structural characterization of the α-mating factor prepro-peptide for secretion
of recombinant proteins in Pichia pastoris. Gene 598:50–62
Chamnipa N et al (2018) The potential of the newly isolated thermotolerant yeast Pichia kudri-
avzevii RZ8-1 for high-temperature ethanol production. Braz J Microbiol 49(2):378–391
Chan MK et al (2018) Expression of stable and active human DNA topoisomerase I in Pichia
pastoris. Protein Expr Purif 141:52–62
Chen Z et al (2011) Recombinant antimicrobial peptide hPAB-beta expressed in Pichia pastoris,
a potential agent active against methicillin-resistant Staphylococcus aureus. Appl Microbiol
Biotechnol 89(2):281–291
Chen Z et al (2020) Efficient biodegradation of highly crystallized polyethylene terephthalate
through cell surface display of bacterial PETase. Sci Total Environ 709:136138
284 W. Sun et al.
Cheng H et al (2014) Genetically engineered Pichia pastoris yeast for conversion of glucose to
xylitol by a single-fermentation process. Appl Microbiol Biotechnol 98(8):3539–3552
Choi S et al (2015) Biorefineries for the production of top building block chemicals and their
derivatives. Metab Eng 28:223–239
Choi DH, Park EH, Kim MD (2017) Isolation of thermotolerant yeast Pichia kudriavzevii from
nuruk. Food Sci Biotechnol 26(5):1357–1362
Chu J et al (2013) Progress in the research of S-adenosyl-L-methionine production. Appl Microbiol
Biotechnol 97(1):41–49
Chung BK et al (2010) Genome-scale metabolic reconstruction and in silico analysis of methy-
lotrophic yeast Pichia pastoris for strain improvement. Microb Cell Fact 9:50
Coughlan AY et al (2016) Centromeres of the yeast Komagataella phaffii (Pichia pastoris) have a
simple inverted-repeat structure. Genome Biol Evol 8(8):2482–2492
Cregg JM et al (1985) Pichia pastoris as a host system for transformations. Mol Cell Biol
5(12):3376–3385
Cregg JM et al (1989) Functional characterization of the two alcohol oxidase genes from the yeast
Pichia pastoris. Mol Cell Biol 9(3):1316–1323
Cregg JM, Vedvick TS, Raschke WC (1993) Recent advances in the expression of foreign genes
in Pichia pastoris. Biotechnology (n y) 11(8):905–910
Crepin VF, Faulds CB, Connerton IF (2003) Production and characterization of the Talaromyces
stipitatus feruloyl esterase FAEC in Pichia pastoris: identification of the nucleophilic serine.
Protein Expr Purif 29(2):176–184
Curran KA, Alper HS (2012) Expanding the chemical palate of cells by combining systems biology
and metabolic engineering. Metab Eng 14(4):289–297
Dagar VK, Khasa YP (2018) Combined effect of gene dosage and process optimization strategies
on high-level production of recombinant human interleukin-3 (hIL-3) in Pichia pastoris fed-
batch culture. Int J Biol Macromol 108:999–1009
Daly R, Hearn MT (2005) Expression of heterologous proteins in Pichia pastoris: a useful exper-
imental tool in protein engineering and production. J Mol Recognit 18(2):119–138
Damasceno LM et al (2007) Cooverexpression of chaperones for enhanced secretion of a single-
chain antibody fragment in Pichia pastoris. Appl Microbiol Biotechnol 74(2):381–389
de Lima PB et al (2016) Novel homologous lactate transporter improves L-lactic acid production
from glycerol in recombinant strains of Pichia pastoris. Microb Cell Fact 15(1):158
de Oliveira JD et al (2016) Genetic basis for hyper production of hyaluronic acid in natural and
engineered microorganisms. Microb Cell Fact 15(1):119
De Schutter K et al (2009) Genome sequence of the recombinant protein production host Pichia
pastoris. Nat Biotechnol 27(6):561–566
De Vuyst L et al (2016) Yeast diversity of sourdoughs and associated metabolic properties and
functionalities. Int J Food Microbiol 239:26–34
Del Mónaco SM et al (2014) Selection and characterization of a Patagonian Pichia kudriavzevii
for wine deacidification. J Appl Microbiol 117(2):451–464
Del Mónaco SM, Rodríguez ME, Lopes CA (2016) Pichia kudriavzevii as a representative yeast
of North Patagonian winemaking terroir. Int J Food Microbiol 230:31–39
Delgado-Ospina J et al (2020) Functional biodiversity of yeasts isolated from Colombian fer-
mented and dry cocoa beans. Microorganisms 8(7)
Deng J et al (2020) Co-expressing GroEL-GroES, Ssa1-Sis1 and Bip-PDI chaperones for enhanced
intracellular production and partial-wall breaking improved stability of porcine growth hor-
mone. Microb Cell Fact 19(1):35
Dong JX et al (2013) Surface display and bioactivity of Bombyx mori acetylcholinesterase on
Pichia pastoris. PLoS One 8(8):e70451
Dong C et al (2020) Engineering Pichia pastoris with surface-display minicellulosomes for car-
boxymethyl cellulose hydrolysis and ethanol production. Biotechnol Biofuels 13:108
Douglass AP et al (2018) Population genomics shows no distinction between pathogenic Can-
dida krusei and environmental Pichia kudriavzevii: one species, four names. PLoS Pathog
14(7):e1007138
Recent Advances in Synthetic Biology Applications … 285
Huangfu J et al (2015) Novel helper factors influencing recombinant protein production in Pichia
pastoris based on proteomic analysis under simulated microgravity. Appl Microbiol Biotechnol
99(2):653–665
Idiris A et al (2010) Engineering of protein secretion in yeast: strategies and impact on protein
production. Appl Microbiol Biotechnol 86(2):403–417
Iglesias-Figueroa B et al (2016) High-Level expression of recombinant bovine lactoferrin in Pichia
pastoris with antimicrobial activity. Int J Mol Sci 17(6)
Isono N et al (2012) A comparative study of ethanol production by Issatchenkia orientalis strains
under stress conditions. J Biosci Bioeng 113(1):76–78
Ito Y et al (2020) Exchange of endogenous and heterogeneous yeast terminators in Pichia pastoris
to tune mRNA stability and gene expression. Nucleic Acids Res 48(22):13000–13012
Jacobs PP et al (2008) Pichia surface display: display of proteins on the surface of glycoengineered
Pichia pastoris strains. Biotechnol Lett 30(12):2173–2181
Jahic M et al (2003) Temperature limited fed-batch technique for control of proteolysis in Pichia
pastoris bioreactor cultures. Microb Cell Fact 2(1):6
Jahic M et al (2006) Process technology for production and recovery of heterologous proteins with
Pichia pastoris. Biotechnol Prog 22(6):1465–1473
Jang IS et al (2018) Improving the efficiency of homologous recombination by chemical and
biological approaches in Yarrowia lipolytica. PLoS One 13(3):e0194954
Jariyachawalid K et al (2012) Effective enhancement of Pseudomonas stutzeri D-phenylglycine
aminotransferase functional expression in Pichia pastoris by co-expressing Escherichia coli
GroEL-GroES. Microb Cell Fact 11:47
Ji Q et al (2020) Improving the homologous recombination efficiency of Yarrowia lipolytica
by grafting heterologous component from Saccharomyces cerevisiae. Metab Eng Commun
11:e00152
Jiang ZB et al (2007) Cell surface display of functionally active lipases from Yarrowia lipolytica
in Pichia pastoris. Protein Expr Purif 56(1):35–39
Jiao L et al (2018) Efficient heterologous production of Rhizopus oryzae lipase via optimization of
multiple expression-related helper proteins. Int J Mol Sci 19(11)
Jin P et al (2014) High-yield novel leech hyaluronidase to expedite the preparation of specific
hyaluronan oligomers. Sci Rep 4:4471
Jo JH et al (2011) Surface display of human lactoferrin using a glycosylphosphatidylinositol-
anchored protein of Saccharomyces cerevisiae in Pichia pastoris. Biotechnol Lett 33(6):1113–
1120
Julius D et al (1984) Isolation of the putative structural gene for the lysine-arginine-cleaving
endopeptidase required for processing of yeast prepro-alpha-factor. Cell 37(3):1075–1089
Kang Z, Zhang N, Zhang Y (2016) Enhanced production of leech hyaluronidase by optimizing
secretion and cultivation in Pichia pastoris. Appl Microbiol Biotechnol 100(2):707–717
Karbalaei M, Rezaee SA, Farsiani H (2020) Pichia pastoris: a highly successful expression system
for optimal synthesis of heterologous proteins. J Cell Physiol 235(9):5867–5881
Kato S et al (2001) Efficient expression, purification and characterization of mouse salivary alpha-
amylase secreted from methylotrophic yeast Pichia Pastoris. Yeast 18(7):643–655
Khasa YP et al (2011) Isolation of Pichia pastoris PIR genes and their utilization for cell surface
display and recombinant protein secretion. Yeast 28(3):213–226
Kittl R et al (2012) Constitutive expression of Botrytis aclada laccase in Pichia pastoris. Bioengi-
neered 3(4):232–235
Kolar NW et al (2007) Functional expression and characterisation of human cytochrome
P45017alpha in Pichia pastoris. J Biotechnol 129(4):635–644
Kong S et al (2020) De novo biosynthesis of 2-phenylethanol in engineered Pichia pastoris.
Enzyme Microb Technol 133:109459
Koutz P et al (1989) Structural comparison of the Pichia pastoris alcohol oxidase genes. Yeast
5(3):167–177
Kuroda K, Ueda M (2011) Cell surface engineering of yeast for applications in white biotechnol-
ogy. Biotechnol Lett 33(1):1–9
Recent Advances in Synthetic Biology Applications … 287
Luley-Goedl C et al (2016) Combining expression and process engineering for high-quality pro-
duction of human sialyltransferase in Pichia pastoris. J Biotechnol 235:54–60
Luu VT et al (2017) Development of recombinant Yarrowia lipolytica producing virus-like parti-
cles of a fish nervous necrosis virus. J Microbiol 55(8):655–664
Macauley-Patrick S et al (2005) Heterologous protein production using the Pichia pastoris expres-
sion system. Yeast 22(4):249–270
Majeke BM et al (2020) Synergistic codon optimization and bioreactor cultivation toward
enhanced secretion of fungal lignin peroxidase in Pichia pastoris: enzymatic valorization of
technical (industrial) lignins. Enzyme Microb Technol 139:109593
Marx H et al (2009) Directed gene copy number amplification in Pichia pastoris by vector inte-
gration into the ribosomal DNA locus. FEMS Yeast Res 9(8):1260–1270
Massahi A, Çalık P (2015) In-silico determination of Pichia pastoris signal peptides for extracel-
lular recombinant protein production. J Theor Biol 364:179–188
Massahi A, Çalık P (2016) Endogenous signal peptides in recombinant protein production by
Pichia pastoris: from in-silico analysis to fermentation. J Theor Biol 408:22–33
Meng Y et al (2014) Production and characterization of recombinant glucose oxidase from
Aspergillus niger expressed in Pichia pastoris. Lett Appl Microbiol 58(4):393–400
Mombeni M et al (2020) pMOX: a new powerful promoter for recombinant protein production in
yeast Pichia pastoris. Enzyme Microb Technol 139:109582
Moura MV et al (2015) Displaying lipase B from Candida Antarctica in Pichia pastoris using the
yeast surface display approach: prospection of a new anchor and characterization of the whole
cell Biocatalyst. PLoS One 10(10):e0141454
Nakamura Y et al (2018) A stable, autonomously replicating plasmid vector containing Pichia
pastoris centromeric DNA. Appl Environ Microbiol 84(15)
Navone L et al (2021) Disulfide bond engineering of AppA phytase for increased thermostability
requires co-expression of protein disulfide isomerase in Pichia pastoris. Biotechnol Biofuels
14(1):80
Nocon J et al (2014) Model based engineering of Pichia pastoris central metabolism enhances
recombinant protein production. Metab Eng 24:129–138
Nong L et al (2020) Engineering the regulatory site of the catalase promoter for improved heterol-
ogous protein production in Pichia pastoris. Biotechnol Lett 42(12):2703–2709
Obst U, Lu TK, Sieber V (2017) A modular toolkit for generating Pichia pastoris secretion
libraries. ACS Synth Biol 6(6):1016–1025
Ou J, Cao Y (2014) Incorporation of Nasutitermes takasagoensis endoglucanase into cell surface-
displayed minicellulosomes in Pichia pastoris X33. J Microbiol Biotechnol 24(9):1178–1188
Owji H et al (2018) A comprehensive review of signal peptides: structure, roles, and applications.
Eur J Cell Biol 97(6):422–441
Pajot HF et al (2011) Unraveling the decolourizing ability of yeast isolates from dye-polluted
and virgin environments: an ecological and taxonomical overview. Antonie Van Leeuwenhoek
99(3):443–456
Papakonstantinou T, Harris S, Hearn MT (2009) Expression of GFP using Pichia pastoris vectors
with zeocin or G-418 sulphate as the primary selectable marker. Yeast 26(6):311–321
Park HJ et al (2018) Low-pH production of d-lactic acid using newly isolated acid tolerant yeast
Pichia kudriavzevii NG7. Biotechnol Bioeng 115(9):2232–2242
Park HJ et al (2018) Draft genome sequence of a multistress-tolerant yeast, Pichia kudriavzevii
NG7. Genome Announ 6(3)
Pena DA et al (2018) Metabolic engineering of Pichia pastoris. Metab Eng 50:2–15
Peng XB et al (2019) High-level secretive expression of a novel achieved Talaromyces cellulolyti-
cus endo-polygalacturonase in Pichia pastoris by improving gene dosage for hydrolysis of
natural pectin. World J Microbiol Biotechnol 35(6):84
Piva LC et al (2020) Construction and characterization of centromeric plasmids for Komagataella
phaffii using a color-based plasmid stability assay. PLoS One 15(7):e0235532
Pontrelli S et al (2018) Escherichia coli as a host for metabolic engineering. Metab Eng 50:16–46
Porro D et al (2005) Recombinant protein production in yeasts. Mol Biotechnol 31(3):245–259
Recent Advances in Synthetic Biology Applications … 289
Prielhofer R et al (2017) GoldenPiCS: a Golden Gate-derived modular cloning system for applied
synthetic biology in the yeast Pichia pastoris. BMC Syst Biol 11(1):123
Qin X et al (2011) GAP promoter library for fine-tuning of gene expression in Pichia pastoris.
Appl Environ Microbiol 77(11):3600–3608
Qin H et al (2016) Microbial diversity and biochemical analysis of suanzhou: a traditional Chinese
fermented cereal gruel. Front Microbiol 7:1311
Rabert C et al (2013) Recombinants proteins for industrial uses: utilization of Pichia pastoris
expression system. Braz J Microbiol 44(2):351–356
Raemaekers RJ et al (1999) Functional phytohemagglutinin (PHA) and Galanthus nivalis agglu-
tinin (GNA) expressed in Pichia pastoris correct N-terminal processing and secretion of het-
erologous proteins expressed using the PHA-E signal peptide. Eur J Biochem 265(1):394–403
Rajamanickam V et al (2017) A novel bi-directional promoter system allows tunable recombinant
protein production in Pichia pastoris. Microb Cell Fact 16(1):152
Rakestraw JA et al (2009) Directed evolution of a secretory leader for the improved expression
of heterologous proteins and full-length antibodies in Saccharomyces cerevisiae. Biotechnol
Bioeng 103(6):1192–1201
Ramakrishnan K et al (2020) Transcriptional control of gene expression in Pichia pastoris by
manipulation of terminators. Appl Microbiol Biotechnol 104(18):7841–7851
Römisch K (2005) Endoplasmic reticulum-associated degradation. Annu Rev Cell Dev Biol
21:435–456
Saitua F et al (2017) Dynamic genome-scale metabolic modeling of the yeast Pichia pastoris. BMC
Syst Biol 11(1):27
Sakai K et al (2008) Construction of a citrinin gene cluster expression system in heterologous
Aspergillus oryzae. J Biosci Bioeng 106(5):466–472
Sallada ND, Harkins LE, Berger BW (2019) Effect of gene copy number and chaperone coexpres-
sion on recombinant hydrophobin HFBI biosurfactant production in Pichia pastoris. Biotechnol
Bioeng 116(8):2029–2040
Sams L et al (2017) Constitutive expression of human gastric lipase in Pichia pastoris and site-
directed mutagenesis of key lid-stabilizing residues. Biochim Biophys Acta Mol Cell Biol
Lipids 1862(10 Pt A):1025–1034
Sanchooli A et al (2018) VLP production from recombinant L1/L2 HPV-16 protein expressed in
Pichia Pastoris. Protein Pept Lett 25(8):783–790
Schwarzhans JP et al (2016) Non-canonical integration events in Pichia pastoris encountered
during standard transformation analysed with genome sequencing. Sci Rep 6:38952
Schwarzhans JP et al (2017) A mitochondrial autonomously replicating sequence from Pichia
pastoris for uniform high level recombinant protein production. Front Microbiol 8:780
Sha C et al (2013) Enhancement of lipase r27RCL production in Pichia pastoris by regulating gene
dosage and co-expression with chaperone protein disulfide isomerase. Enzyme Microb Technol
53(6–7):438–443
Shalgi R et al (2005) A catalog of stability-associated sequence elements in 3’ UTRs of yeast
mRNAs. Genome Biol 6(10):R86
Shen Q et al (2012) The effect of gene copy number and co-expression of chaperone on production
of albumin fusion proteins in Pichia pastoris. Appl Microbiol Biotechnol 96(3):763–772
Shimizu T, Kinoshita H, Nihira T (2007) Identification and in vivo functional analysis by gene
disruption of ctnA, an activator gene involved in citrinin biosynthesis in Monascus purpureus.
Appl Environ Microbiol 73(16):5097–5103
Silva AJD et al (2021) Pichia pastoris displaying ZIKV protein epitopes from the envelope and
NS1 induce in vitro immune activation. Vaccine 39(18):2545–2554
Sohn SB et al (2010) Genome-scale metabolic model of methylotrophic yeast Pichia pastoris and
its use for in silico analysis of heterologous protein production. Biotechnol J 5(7):705–715
Song W et al (2020) Multiple strategies to improve the yield of chitinase a from Bacillus licheni-
formis in Pichia pastoris to obtain plant growth enhancer and GlcNAc. Microb Cell Fact
19(1):181
290 W. Sun et al.
Abstract
In Kluyveromyces marxianus, a thermotolerant yeast that has already been uti-
lized for producing various useful materials, recent advances including complete
genome sequencing and transcriptome analysis have provided valuable infor-
mation on its physiological and metabolic characteristics including its capacity
for assimilation of various sugars and tolerance to high temperatures, which
are distinct from Saccharomyces cerevisiae. These prominent properties enable
the development of high-temperature fermentation (HTF) technology for indus-
trial applications. In addition, the yeast is able to integrate DNA fragments
into its genome, which makes it easy to introduce foreign genes or gene sets
N. Lertwattanasakul
Department of Microbiology, Faculty of Science, Kasetsart University, Bangkok 10900, Thailand
M. Nurcholis
Department of Food Science and Technology, Faculty of Agricultural Technology, Brawijaya
University, Malang 65145, Indonesia
N. Rodrussamee
Department of Biology, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
T. Kosaka · M. Yamada (B)
Department of Biological Chemistry, Faculty of Agriculture, Yamaguchi University,
Yamaguchi 753-8515, Japan
e-mail: [email protected]
T. Kosaka · M. Murata · M. Yamada
Graduate School of Science and Technology for Innovation, Yamaguchi University,
Yamaguchi 753-8515, Japan
T. Kosaka · M. Yamada
Research Center for Thermotolerant Microbial Resources, Yamaguchi University,
Yamaguchi 753-8515, Japan
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 293
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_11
294 N. Lertwattanasakul et al.
into the genome. Moreover, DNA editing technology and simulation models
are in place. Therefore, K. marxianus is a new platform in synthetic biology for
producing useful materials.
1 Introduction
Microbial production of fuels and chemicals from biomass and other renewable
carbon sources is an attractive alternative to petroleum-derived production. Saccha-
romyces cerevisiae is the organism of choice because of its high rate of production
and tolerance to ethanol titers upwards of 120 g L−1 (Qiu and Jiang 2017; Nielsen
et al. 2013). These outstanding phenotypes, among others, may have led to the
widespread study of S. cerevisiae and its development as a model eukaryotic host
for chemical biosynthesis.
Kluyveromyces marxianus is a Crabtree-negative yeast and favors respiration
over fermentation. This yeast is also industrially relevant because of its wide
substrate spectrum including pentose sugars, fast-growth characteristics, and ther-
motolerance to ~50 °C (Varela et al. 2017; Löbs et al. 2016), making it a promising
host for industrial biotechnology to produce renewable chemicals from agricultural
biomass feedstocks. Native strains of K. marxianus are also known to synthesize
ethyl acetate at rates above 2 g L−1 h−1 in aerated bioreactors (Löser et al. 2013,
2015). In recent years, increasing interest has been shown in several new applica-
tions including production of biomolecules (Hughes et al. 2017; Lin et al. 2017),
biocatalysts (Wang et al. 2017; Simoness et al. 2017), and heterologous protein
expression (Lee et al. 2017; Gombert et al. 2016). However, major genetic engi-
neering limitations have kept this yeast from replacing the commonly used yeast
S. cerevisiae in industrial applications.
One of the features of K. marxianus that are suitable for industrial applications
is its ability to utilize a broad range of substrates. The major common feature of
sugar utilization by K. marxianus is the ability to assimilate lactose and inulin
as a carbon source, a feature that is absent in S. cerevisiae (Lane and Morrissey
2010). The ability has evolved by acquiring LAC genes, LAC12 and LAC4, that
are accountable for the uptake of lactose and subsequent cleavage into galactose
and glucose, respectively, and the inulinase gene INU1, which liberates fructose
molecules from oligo- or poly-sugars with β-(2,1)-linked fructose units at the
terminal (Rouwenhorst et al. 1988).
Consistent with the capacity of this yeast to utilize a wide variety of substrates,
there are a number of sugar transporters, including 27 putative sugar transporters,
encoded by the genome of K. marxianus DMKU 3-1042 (Lertwattanasakul et al.
2015). The ability for utilization of a wide variety of carbon sources by various
strains of K. marxianus compared to S. cerevisiae has been shown (Nonklang et al.
Kluyveromyces marxianus as a Platform in Synthetic Biology … 295
2008) (Fig. 1). These carbon sources include monosaccharides of six-carbon sugars
(galactose, fructose, and mannose), monosaccharides of five-carbon sugars (xylose,
xylitol, and arabinose), disaccharides (sucrose, lactose, and cellobiose), trisaccha-
ride (raffinose), polysaccharide (inulin), and a non-fermentable carbon source of
glycerol. Besides, K. marxianus DMB1 can utilize sorbitol, which is not generally
utilized by K. marxianus strains, as a non-fermentable carbon source (Goshima
et al. 2013b). K. marxianus can utilize these carbon sources, while S. cerevisiae
cannot utilize cellobiose, xylose, xylitol, arabinose, glycerol, sorbitol, and lactose.
Ethanol is one of the classic bulk chemical products of yeasts. Much interest has
been shown in K. marxianus for the possibility of it being a platform for ethanol
production because of not only its capability of utilizing a broad range of substrates
but also its fermentation ability at high temperatures. These properties lead us to
develop a stable HTF technology with several benefits as described below and
to utilize not only first-generation substrates but also second- or third-generation
substrates.
Related to conversion from lactose to ethanol, which has been received much
attention, expression of genes for β-galactosidases in K. marxianus is induced
by its natural inducers, galactose, and lactose. However, the production of β-
galactosidases seems to be dependent on the substrate concentration. In K.
marxianus CBS6556, the maximum activity of β-galactosidase is obtained at low
concentrations of the inducer carbohydrates, in the range between 0.5 and 15 mM.
When shifted to a high concentration of D-galactose or lactose, a repressive mech-
anism is superimposed to the inducing effect of the substrate (Martins et al. 2002).
K. marxianus exhibits glucose repression in galactose utilization, and galactose is
utilized after diminishing glucose in a mixed-sugar medium (Rodrussamee et al.
2011). Moreover, mutation of MIG1, which plays key roles as a regulator complex
in glucose repression, increases the activity of lactose hydrolysis (Zoppellari and
Bardi 2013).
In dairy industries, dairy effluents as high strength wastewater contains a
high concentration of lactose, which is responsible for the high chemical oxy-
gen demand (COD), causing an environmental problem (Karim et al. 2020). Whey
and permeate are effluents from cheese processing, while scotta is effluent from
ricotta processing. Whey is a waste produced by dairy industries in large amounts
(approximately, 10 L per kg of cheese), and it contains 4–6% lactose in addi-
tion to proteins and other nutrients (Guimarães et al. 2010). Many studies on the
conversion of whey to ethanol have been performed because whey is a low-cost
and abundant material with high carbohydrate concentrations (Zafar and Owais
2006; Ozmihci and Kargi 2007; Guimarães et al. 2010; Zoppellari and Bardi 2013;
Roohina et al. 2016). It has been reported that the yield of ethanol from cheese
whey powder produced by K. marxianus DSMZ-7239 is equal to the theoretical
yield of 0.54 g EtOH per g lactose (Ozmihci and Kargi 2007). Moreover, whey
permeate and scotta have the potential to become key inexpensive substrates for
bioethanol production by K. marxianus (Jedrzejewska and Kozak 2011; Sansonetti
et al. 2009; Zoppellari and Bardi 2013). Besides dairy byproducts, third-generation
substrates such as marine biomass have been used for producing ethanol. For
example, the red seaweed Gracilaria verrucosa, which contains glucose and galac-
tose, has been used for producing ethanol, with S. cerevisiae, Candida lusitaniae,
and K. marxianus adapted to high concentrations of galactose (Park et al. 2020).
Among them, K. marxianus has shown the highest ethanol yield coefficient of
0.47 g g−1 .
Conversion of inulin as a D-fructose polymer to ethanol is a notable conversion
process that maximally utilizes the characteristics of K. marxianus, and extensive
studies on inulin have been performed using Jerusalem artichoke (JA), a crop that
contains nearly 20% of carbohydrates, 70–90% of which is inulin (Hu et al. 2012;
Yuan et al. 2012; Kim et al. 2013; Kim and Kim 2014; Charoensopharat et al.
2015). Inulinase encoded by INU1 is very effective in hydrolyzing inulin to D-
fructose monomers, which are then imported into cells by the fructose transporter
Frt1p and catabolized through the glycolytic pathway. Regulation of the expression
of INU1 in K. marxianus has been studied using various strains, and it has been
shown that the production of inulinase depends on the carbon source (Sokolenko
Kluyveromyces marxianus as a Platform in Synthetic Biology … 297
↑Bbxfpk Bifidobacterium
breve
↑Septa Salmonella
enterica
(continued)
299
300
Table 1 (continued)
Aimed product Modification Source Parental strain Cultivation condition Titer References
↑EcppsA Escherichia coli
↑ScARO10 ↑ScADH2 S. cerevisiae DMKU 3-1042 Glucose, 30 °C 1.0 g/L Kim et al. (2014)
Hexanoic acid ↑atoB E. coli ATCC17555 Galactose, 37 °C 154 mg/L Cheon et al. (2014)
↑bktB Ralstonia eutropha
↑crt ↑hbd Clostridium
acetobutylicum
↑ter Treponema
denticola
↑MCT1 S. cerevisiae
Pyruvate KmPDC1 KmGPD1 K. marxianus YZJ051 Glucose/xylose, 42 °C 29.21 g/L Zhang et al. (2017a,
↑KmMTH1-ΔT b)
↑SsXYL2 Scheffersomyces
stipitis
↑ScGAL2 (N376F) S. cerevisaie
β-Carotene ↑HpCHYb Haematococcus KY3 Galactose, 30 °C 224.4 ± 9.9 mg/g DCW Chang et al. (2015)
pluvialis
Triacetic acid lactone ↑G2PS1 Gerbera hybrida CBS 6556 Xylose, 37 °C 1.24 g/L McTaggart et al.
(2019)
D-Allulose ↑dpe Agrobacterium CICC 1911 Fructose, 55 °C 190 g/L Yang et al. (2018)
tumefaciens
Fatty acid ethyl esters ↑KmATF1 K. marxianus UMPe-1 Glucose, 30 °C 1.53% Campos-García
et al. (2018)
Cellulases ↑EGI ↑EGIA Aspergillus niger KY3 Glucose – Chang et al. (2017)
(continued)
N. Lertwattanasakul et al.
Table 1 (continued)
Aimed product Modification Source Parental strain Cultivation condition Titer References
↑EGIII ↑CBHI Trichoderma
reesei
↑CBHII Chemically
synthesized and
optimized for K.
marxianus
↑NpaBGS Neocallimastix
patriciarum
Tannase – – NRRL Y-8281 Olive pomace, 45 °C 1026.12 U/mg Mahmoud et al.
(2018)
α-Galactosidase – Cyamopsis CBS 6556 Starch 153 mg/L Bergkamp et al.
tetragonoloba (1993)
α-Amylase ↑TAA A. oryzae DMKU 3-1042 – – Hoshida et al.
(2014)
β-Galactosidase ↑KlLAC4 Kluyveromyces NBRC1777 Glucose, 45 °C 20 U/mg Yang et al. (2015)
lactis
β-Glucuronidase ↑gusA E. coli NBRC1777 Glucose, 45 °C 0.8 U/mg Yang et al. (2015)
Superoxide dismutase ↑KlSOD1 K. lactis L3 Glucose, 30 °C 24 kU/L Raimondi et al.
Kluyveromyces marxianus as a Platform in Synthetic Biology …
(2010)
β-Glucuronidase ↑gusA E. coli KM1 Glucose, 37 °C 50 nmole 4-MU/min Pecota and Da silva
/mL (2005)
Glucose oxidase ↑GOX A. niger CBS 6556 Glucose, 30 °C 1722 U/g DCW Rocha et al. (2010)
(continued)
301
302
Table 1 (continued)
Aimed product Modification Source Parental strain Cultivation condition Titer References
Esterase ↑est Thermus CBS 6556 Sucrose 56.9 U/g DCW Rocha et al. (2011)
thermophilus
Glucoamylase ↑GAA Arxula L3 Glucose, 40 °C 80 U/mL Raimondi et al.
adeninivorans (2013)
Dengue virus type I – Dengue virus UFV-3 Galactose, 37 °C 1.2 mg/mL Bragança et al.
non-structural protein 1 (2015)
Cytochrome P450 ↑CYP505A1 Fusarium UOFS Y1185 Glucose, 28 °C – Theron et al. (2014)
monooxygenase oxysporum
↑CYP102A1 B. megaterium
Xylose isomerase ↑XYLA Orpinomyces NBRC1777 Xylose – Wang et al. (2013)
Porcine circovirus type 2 – Porcine circovirus FIM-1 Glucose, 30 °C 1.91 g/L Duan et al. (2019)
virus-like particles type 2
Porcine parvovirus – Parvovirinae virus FIM-1 Glucose, 30 °C 2.5 g/L Yang et al. (2021)
virus-like particles
Single-chain antibody ↑scFv HyHEL-10 scFv NBRC1777 Xylose, 30 °C – Nambu-Nishida
et al. (2018)
Ethanol ↑KmXYL2 (NADP+ -dependent) K. marxianus DMB13 Xylose/glucose, 40 °C – Suzuki et al. (2019)
↑PsXYL1 (N272D) Pichia stipitis NBRC1777 Xylose, 42 °C 3.55 g/L Zhang et al. (2013)
↑NcXYL1 N. crassa NBRC1777 Xylose/glucose, 42 °C 6.22 g/L Zhang et al. (2015a,
b)
(continued)
N. Lertwattanasakul et al.
Table 1 (continued)
Aimed product Modification Source Parental strain Cultivation condition Titer References
↑PsXYL2 P. stipitis
↑PsXYL1 ↑PsXYL2 P. stipitis DMB1 Xylose, 45 °C 3.3 g/L Goshima et al.
↑ScXYL3 S. cerevisiae (2013a)
↑BGL A. aculeatus
↑FLO1 ↑FLO5 ↑FLO9 or S. cerevisiae DMKU 3-1042 Glucose, 40 °C 48 g/L Nonklang et al.
↑FLO10 (2009)
(continued)
303
304
Table 1 (continued)
Aimed product Modification Source Parental strain Cultivation condition Titer References
↑BGL Thermoascus NBRC1777 Cellobiose, 45 °C 29.5 g/L Matsuzaki et al.
aurantiacus (2012)
↑ScGAL2 ↑ScGAL2 (N376F) S. cerevisiae KCTC 17555 Glucose/galactose, 30 °C 36.14 g/L Kwon et al. (2020)
↑olpB-ScGPI ↑cipA C. thermocellum 4G5 Avicel, 30 °C 3.09 g/L Anandharaj et al.
S. cerevisiae Phosphoric acid-swollen 8.61 g/L (2020)
cellulose, 30 °C
LDH lactate dehydrogenase, JEN1 proton-coupled monocarboxylate transporter, PFK 6-phosphofructokinase, dld1 putative d-lactate dehydrogenase, GAL2 galactose permease, XYL1 xylose
reductase, XYL2 xylitol dehydrogenase, XYL3 xylulokinase, XYLA xylose isomerase, ARO3/ARO4 DAHP synthase, ARO7 chorismate mutase, ARO8 aromatic aminotransferase, ARO10
phenylpyruvate decarboxylase, PHA2 prephenate dehydratase, eat1 ethanol acetyltransferase, TKL1 transketolase, TAL1 transaldolase, xfpk phosphoketolase, pta phosphotransacetylase,
ppsA phosphoenolpyruvate synthase, ADH2 alcohol dehydrogenase 2, atoB acetyl-CoA acetyltransferase, bktB β-ketothiolase, crt crotonase, hbd 3-hydroxybutyryl-CoA dehydrogenase, ter
trans-enoyl-CoA reductase, MCT1 malonyl CoA-acyl carrier protein transacylase, PDC1 pyruvate decarboxylase, GPD1 glycerol-3-phosphate dehydrogenase, MTH1 negative regulator of the
glucose-sensing signal transduction pathway, CHYb β-carotene hydroxylase, G2PS1 2-pyrone synthase, dpe D-psicose-3-epimerase, ATF1 alcohol acetyl-transferase, EG endoglucanase,
CBH cellobiohydrolase BGL β-glucosidase, LAC4 β-galactosidase, gusA β-glucuronidase, SOD1 superoxide dismutase, GOX glucose oxidase, est esterase, GAA glucoamylase, CYP
cytochrome P450 monooxygenase, AMY amylase, GAM1 glucoamylase, CDT cellodextrin transporter, FLO flocculating, GPI glycosylphosphatidylinositol, cipA scaffoldin, olpB anchoring
protein
N. Lertwattanasakul et al.
Kluyveromyces marxianus as a Platform in Synthetic Biology … 305
K. marxianus has the interesting feature of producing several useful materials such
as ethanol as its main product (Table 2). The application of HTF at a temperature
of about 40 °C or higher is expected to provide many advantages over general
fermentation at a temperature of about 30 °C or less including reducing operating
costs of cooling systems, minimizing risk of contamination, reducing viscosity of
the fermentation broth, efficiently achieving simultaneous saccharification and fer-
mentation, robustness process against accidental temperature rise even in tropical
countries, operating continuous ethanol removal, and increasing enzyme activity
for biomass hydrolysis (Banat et al. 1998; Fonseca et al. 2008; Hoshida and Akada
2017; Kosaka et al. 2018). Fermenting microbes for HTF must be thermotolerant,
that is, they must grow sufficiently and ferment efficiently at high temperatures.
Yeast strains that have often been isolated as ethanol-producing thermotoler-
ant yeasts are strains of K. marxianus and S. cerevisiae because they are highly
thermotolerant yeasts and efficient ethanol-producing yeasts, respectively (Hoshida
and Akada 2017). K. marxianus DMKU 3-1042 efficiently produced ethanol from
sugar cane juice at a temperature of 40 °C, and the maximal ethanol concentration
was 67.8 g L−1 and the yield was 60.4% of the theoretical yield (Limtong et al.
2007). Five K. marxianus strains isolated in Laos exhibited strong fermentation
abilities in a 16% sugars-containing medium of glucose, sucrose, sugarcane, or
molasses at 40 °C (Keo-oudone et al. 2016). The evolutionary adapted KM-100d
306 N. Lertwattanasakul et al.
Table.2 Comparison of ethanol production levels and yields among various strains of K. marxi-
anus
Strains Temp Carbon sources Time Ethanol Ethanol References
(°C) (h) production yield (g
(g L−1 ) g−1 )
K. marxianus 30 Xylose 48 1.71 ± 0.45 0.09 ± 0.03 Nitiyon et al.
DMKU 3-1042 (2016)
30 Xylose 72 ~2.60 0.13 Rodrussamee
et al. (2011)
37 Xylose 36 1.29 ± 0.23 0.07 ± 0.01 Nitiyon et al.
(2016)
40 Xylose 72 ~2.20 0.11 Rodrussamee
et al. (2011)
45 Xylose 48 ~0.96 0.06 Rodrussamee
et al. (2011)
40 Sugar cane 60 67.8 0.6 Limtong et al.
juice (2007)
K. marxianus 30 Xylose 48 2.91 ± 0.40 0.15 ± 0.02 Nitiyon et al.
BUNL-21 (2016)
37 Xylose 36 2.58 ± 0.05 0.14 ± 0.00 Nitiyon et al.
(2016)
40 Glucose 24 68.7 0.43 Keo-oudone
et al. (2016)
K. marxianus IMB3 45 Xylose 0.8–1.2 0.08–0.12 Banat et al.
(1998)
K. marxianus FIM1 30 Glucose 48 110 0.55 Mo et al. (2019)
45 Glucose 48 50 0.25 Mo et al. (2019)
K. marxianus 40 Mannose 24 8.52 0.426 Rouhollah et al.
(2007)
K. marxianus SBK1 40 Glucose and 72 23.82 0.35 Kim et al. (2019)
xylose
K. marxianus DMB1 42 Lignocellulosic 25.98 Goshima et al.
hydrolysates (2013b)
K. marxianus CICC 40 Lignocellulosic 42.6 Du et al. (2019)
1727–5 hydrolysates
K. marxianus 37 Inulin 72 ~104.83 0.47 Charoensopharat
DBKKU Y-102 jerusalem et al. (2015)
artichoke
40 Inulin 72 ~97.46 0.45 Charoensopharat
jerusalem et al. (2015)
artichoke
K. marxianus Y179 30 Inulin 36 98 0.43 Gao et al. (2015)
Kluyveromyces marxianus as a Platform in Synthetic Biology … 307
(Quarella et al. 2016), UFS-Y2791 (Schabort et al. 2016), and other nine strains:
L01, L02, L03, L04, L05, CBS397, NBRC0272, NBRC0288, and NBRC0617
(Ortiz-Merino et al. 2018). However, complete genome sequences of only two
strains, DMKU 3-1042 and NBRC1777, are currently available. The former
genome of 11.0 Mb is composed of 8 chromosomes in total including mitochon-
drial DNA. Annotation of the genome revealed a total of 4952 genes. A total of
202 tRNAs and 8 rDNAs were identified.
A major potential future application of K. marxianus may be ethanol production
from lignocellulosic biomass, which is an anaerobic or oxygen-limited process
in which both glucose and xylose are present. Detailed transcription start site
sequencing (TSS Seq) to explore the response of K. marxianus DMKU 3-1042
was thus performed under four different conditions: shaking condition in a rich
medium at 30 °C (30D) or 45 °C (45D), a static condition in a rich medium at
30 °C (30DS), and shaking condition in a xylose-containing rich medium at 30 °C
(30X) (Lertwattanasakul et al. 2015).
Under the 30DS condition, K. marxianus may increase the turnover of RNAs
and proteins in addition to suppression of transporters that depend on the mito-
chondrial respiratory activity. Most of the genes for several oxygen-dependent
biosynthetic pathways, such as those for heme, sterols, unsaturated fatty acids,
pyrimidine, and deoxyribonucleotides (Ishtar Snoek and Yde Steensma 2006), are
crucial for cellular metabolism under a static condition. Under the 45D condition,
K. marxianus seems to drastically change metabolic pathways, that is, enhance-
ment of the pentose phosphate pathway (PPP) and attenuation of the TCA cycle
after the fumarate-producing step. Several genes involved in both the DNA repair
pathways of homologous recombination (HR) and non-homologous end-joining
(NHEJ) are upregulated. Heat shock proteins and chaperones, such as Hsp26,
Hsp60, Hsp78, Hsp82, Ssa3, and Cpr6, are crucial for survival at high temper-
atures. The thermotolerance of K. marxianus is thus likely to be achieved by
systematic mechanisms consisting of various strategies. The yeast prevents the
generation of reactive oxygen species (ROS) by minimizing mitochondrial activ-
ity and mainly acquires ATP from glycolysis rather than from the TCA cycle at
high temperatures. Fu et al. (2019) also reported that excess ROS generated dur-
ing a high temperature fermentation condition could be neutralized by NADPH
in K. marxianus. The degree of fatty acid unsaturation may be reduced to adapt
to high temperatures. Genes associated with DNA repair or lipid composition of
the plasma membrane are upregulated. The yeast also produces more ergosterol
to deal with ethanol stress. Under the 30X condition, degradation of lipids in the
peroxisome seems to be stimulated and amino acid synthesis is kept at a low level,
indicating the possibility that fatty acids could be a subsidiary intracellular carbon
source in the xylose medium. Consistently, Schabort et al. (2016) also reported
that peroxisomal fatty acid catabolism is dramatically upregulated in a defined
xylose mineral medium without fatty acids. They also described mechanisms by
which fatty acids are activated and products of β-oxidation are transferred to the
mitochondria.
Kluyveromyces marxianus as a Platform in Synthetic Biology … 311
Notably, oxidative stress-response genes were highly induced under the three
conditions tested, indicating that ROS accumulated in the cytoplasm, mitochondria,
and peroxisome under the 30DS and 30X conditions and in the cytoplasm and
mitochondria under the 45D condition. In conclusion, K. marxianus likely adapts
to the three different growth conditions by distinctive metabolic pathways from
the control condition. Interestingly, the yeast appears to overcome the problem of
ROS, which tend to accumulate under all three conditions. Nicotinamide adenine
dinucleotide phosphate (NADPH) synthesis from several reactions is the key for
cells to cope with ROS.
Marcišauskas et al. (2019) reported the first genome-scale metabolic model of
K. marxianus, iSM996, using TSS data reported by Lertwattanasakul et al. (2015).
The model includes 1913 reactions associated with 996 genes and 1531 metabo-
lites. The iSM996 was used to construct three condition-specific models in YPD
medium while considering the low oxygen and high temperature conditions. The
results suggest that at a high temperature, the cell turns off more genes, thereby
introducing new auxotrophies and utilizing as many resources as possible from the
medium. These findings may be used in the design of growth media at low levels
of oxygen and/or high temperatures.
The global transcriptional response of K. marxianus to multiple inhibitors
including acetic acid, phenols, furfural, and HMF at 42 °C was also studied via
RNA-seq technology (Wang et al. 2018). Genes involved in the glycolysis pathway,
fatty acid metabolism, ergosterol metabolism, and vitamin B6 and B1 metabolic
process were enriched in the downregulated gene set, while genes involved in the
TCA cycle, respiratory chain and detoxification of ROS and transporter coding
genes were enriched in the upregulated gene set in response to the stress with
multiple inhibitors. Redox balance and NAD(P)+ /NAD(P)H homeostasis play an
important role in tolerance to lignocellulose-derived inhibitors.
K. marxianus can grow well at temperatures over 45 °C, unlike K. lactis, which
belongs to the same genus, or S. cerevisiae, which is a closely related yeast in
hemiascomycetous yeasts. K. marxianus may thus have an intrinsic mechanism to
survive at high temperatures. K. marxianus strains are relatively resistant against
hydrogen peroxide, furfural, and hydroxymethyl furfural (Nitiyon et al. 2016).
Thermotolerance of the yeast may thus overlap with other stress tolerances, assum-
ing a common mechanism of robustness against stressors. K. marxianus, which is
one of the mesophilic yeasts, may have evolved into yeast with thermal resis-
tance due to natural selection pressure in a high-temperature environment, and the
acquisition of thermotolerance might have allowed it to withstand other stresses.
A clue for the mechanism of thermotolerance might be provided by tran-
scriptome analysis, flux analysis, or comparison with a non-thermotolerant yeast.
Transcriptome analysis (Lertwattanasakul et al. 2015) revealed that there are a
tremendous number of significantly upregulated and downregulated genes in K.
312 N. Lertwattanasakul et al.
an increase in respiration rate (Abbott et al. 2009). Low levels of ROS are scav-
enged by non-enzymatic and enzymatic anti-oxidizing agents such as glutathione
(GSH), thioredoxin (TRX), superoxide dismutase, catalase, and peroxidases, but
high levels of ROS cause oxidation of intracellular components, such as DNA,
protein, and lipid and induce apoptosis (Madeo et al. 1999; Scherz-Shouval and
Elazar 2007). Increased activity of the PPP supports the high demand of NADPH
at high temperatures for glutathione reductase-mediated protection against oxida-
tive stress (Konings 1988; Grant 2001; Sugiyama et al. 2000). However, it is not
likely that a higher level of NADPH production is sufficient for thermotolerance
because increased PPP activity (Celton et al. 2012; Frick and Wittmann 2005) in S.
cerevisiae is not capable of supporting growth at elevated temperatures, and ther-
mosensitive K. lactis (Tarrío et al. 2006) and Pichia pastoris (Jorda et al. 2014)
have relatively high PPP activity. Notably, genome-wide analysis of thermotoler-
ant genes supporting cell survival at a critical high temperature (CHT), an upper
limit of temperature, in three bacteria, Escherichia coli (Murata et al. 2011, 2018),
thermotolerant Acetobacter tropicalis (Soemphol et al. 2011), and thermotolerant
Zymomonas mobilis (Charoensuk et al. 2017), revealed that they share thermo-
tolerant genes for membrane stabilization, protection against oxidative stress, and
repair of damage of DNA or proteins, which may contribute to minimization of
ROS generation, avoidance of ROS damage, and recovery from ROS damage,
respectively.
As for thermotolerance, microbes seem to have repair mechanisms for damage
of proteins, lipid, or DNA caused by ROS (Piper 1993) and structural specialties
of the membranes and enzymes, which may be crucial for metabolic activity at
high temperatures (Mejía-Barajas et al. 2018; Fields 2001). Heat shock proteins
including proteinases participate in the renaturation or degradation of unfolded or
damaged proteins. Enhanced expression of genes for heat shock proteins as well
as genes for ROS-scavenging enzymes improves thermotolerance in Z. mobilis
(Anggarini et al. 2016). The susceptibility of lipids to oxidation depends on the
lipid composition and degree of unsaturation (Catalá 2012). The extent of cellular
damage under heat shock conditions is correlated positively with increasing unsat-
uration of fatty acids (Suryawati et al. 2008). Consistently, Steels et al. (1994)
found that the most stress-resistant yeast membranes were enriched in saturated
fatty acids. Membrane fluidity is related to the ratio of saturated to unsaturated
fatty acids (Los and Murata 2004).
To further understand the mechanism of thermotolerance in yeasts, thermal
adaptation followed by detailed analysis of causative mutations or enhanced ther-
motolerance of a thermosensitive yeast by heterologous expression of genes from a
thermotolerant yeast may be useful. Heterologous expression of heat shock genes
from thermophiles has been shown to improve thermotolerance of S. cerevisiae
(Liu et al. 2014). The transcription factors KmHsf1 and KmMsn2 of K. marxianus
can promote both cell growth and ethanol fermentation of S. cerevisiae at high tem-
peratures (Li et al. 2017). These two transcription factors might increase ethanol
production by different mechanisms. In addition, KmMsn2 might also help to cope
314 N. Lertwattanasakul et al.
Escherichia coli has so far been primarily used as a host for the production of use-
ful proteins by genetic engineering, probably due to the accumulation of advanced
methodologies including the development of various vectors that are superior to
other microbes (Rosano and Ceccarelli 2014). Although various kinds of cytokines
or growth factors are produced with E. coli and are already commercially available,
E. coli has some drawbacks (Nausch et al. 2013; Feng et al. 2015). For example,
lipopolysaccharide as a constituent of the membrane in E. coli is harmful to the
human body, and great care must therefore be taken, especially in the purifica-
tion of pharmaceutical products (Storeng et al. 1987). The surface antigen protein
of hepatitis B virus (HBsAg) or tissue plasminogen activators (TPA) cannot be
Kluyveromyces marxianus as a Platform in Synthetic Biology … 315
produced as active proteins, and small substances such as peptide hormones are
decomposed in E. coli cells (Pumpen et al. 1984; Elghanam et al. 2012; Xu et al.
2017). Therefore, attempts have been made to use other hosts to compensate for the
disadvantage of E. coli, and yeast is drawing attention as one of the candidates. The
utilization of S. cerevisiae, which is generally used as a host for DNA recombina-
tion research (Mattanovich et al. 2012; Porro et al. 2005), has several advantages:
1. a wealth of genetic knowledge has accumulated, 2. the fundamental mecha-
nisms of replication, transcription and translation have been elucidated, making it
a suitable model of analysis of biological phenomena in higher organisms, 3. the
yeast has a long history of being used industrially as a useful microorganism for
food, feed, and pharmaceutical raw materials, and 4. there is abundant information
on fermentation engineering and culture engineering. However, S. cerevisiae also
has drawbacks for protein production. S. cerevisiae as a Crabtree-positive yeast
consumes oxygen in a limited range in the presence of glucose above a certain
density, despite high levels of dissolved O2 (van Urk et al. 1990). As a result,
S. cerevisiae specializes in ethanol production, reducing the yield of the target
product per sugar. Therefore, the Crabtree effect may be one of the causes of the
small applicable range of material production of S. cerevisiae. In addition, since the
yeast growth temperature is around 30 °C, temperature control during fermentation
is indispensable and costly.
Thermotolerant K. marxianus, which is generally accepted as safe (GRAS), is
a Crabtree-negative yeast that acts on the metabolism from a TCA cycle with a
priority independent of glucose concentration and performs aerobic alcohol fer-
mentation (Blank et al. 2005; Vandijken et al. 1993). Additionally, K. marxianus
is the fastest-growing yeast known so far (Groeneveld et al. 2009), reported growth
rates of K. marxianus, K. lactis, S. cerevisiae and P. pastoris being 0.80 h−1 ,
0.50 h−1 , 0.37 h−1 , and 0.18 h−1 , respectively (Lane and Morrissey 2010; Gao
et al. 2012), and has prominent features including efficient fermentation at high
temperatures, assimilation of various sugars, and advanced genetic tools as men-
tioned above. Therefore, K. marxianus is one of the alternative yeasts that have
great potential in technology to produce various useful proteins under aerobic
conditions.
K. marxianus has promising properties for producing various enzymes such
as β-glucosidase, β-galactosidase, inulinase, endopolygalacturonase, and xylosi-
dase (Fonseca et al. 2008; Lertwattanasakul et al. 2015) (Fig. 1). β-glucosidase
of Kluyveromyces fragilis ATCC 12424 has been cloned and expressed in S. cere-
visiae, and it has been applied for hydrolysis of cellulosic materials (Raynal et al.
1987). The ability to hydrolyze cellulose is useful for further application in ethanol
production from cellulosic biomass. β-Glucosidase is one of three major cellulase
enzymes that are responsible for the regulation of the whole cellulolytic process
to produce fermentable sugars (Zhou et al. 2018). β-Galactosidase can be pro-
duced from a MIG1 mutant of K. marxianus KM-15 for hydrolysis of lactose in
foods to become glucose and galactose (Zhou et al. 2013). This property is very
attractive because it produces lactose-free or low-lactose foods and reduces health
316 N. Lertwattanasakul et al.
issues of people who suffered from lactose intolerance (Wolf et al. 2018). Inuli-
nase can be produced from the disrupted MIG1 gene and the overexpressed INU1
gene (Zhou et al. 2014). The enzyme is used to hydrolyze inulin by degrading the
β-2,1-fructosyl bond to become fructose (Hoshida et al. 2018). The enzyme can
be applied in food, fermentation, pharmaceutical, and chemical industries (Chi
et al. 2011). Endopolygalacturonase produced by K. marxianus BKM Y-719 is
used for the reduction of viscosity in fruit processing products (Šiekštele et al.
1999). Xylosidase is related to catalysis of the reducing site of xylooligosaccha-
rides and liberates xylose. Efficient liberation of xylose is related to the ability
of a non-conventional yeast to convert xylose to ethanol. Xylosidase activity of
thermotolerant yeast K. marxianus strains NIRE-K1 and NIRE-K3 was found to
be higher than that of Candida tropicalis. An increased amount of xylosidase
was found in adapted K. marxianus cells (Behera et al. 2016). Other possible
enzymes related to biomass utilization have been predicted from the whole genome
sequence (Lertwattanasakul et al. 2015), and they include β-glucosidase (LAC4),
endo-1,3(4)-β-glucanase 1 (DSE4), and endo-1,3(4)-β-glucanase 2 (ACF2) in addi-
tion to three lactose permeases (3 copies of LAC12). Phosphatase, aminopeptidase,
and carboxypeptidase have also been reported as possible enzymes related to
biomass utilization (Nurcholis et al. 2020).
Production of proteins with K. marxianus that have been investigated so far
include enzymes (β-galactosidase, β-glucosidase, inulinase, polygalacturonase, and
others) (Topete et al. 1997; Su et al. 2021; Yarimizu et al. 2015; Hoshida et al.
2018), single-cell proteins (Rajkumar and Morrissey 2020), and antibodies (Duan
et al. 2019; Yang et al. 2021). However, there is a problem in producing a protein
using K. marxianus. K. marxianus cells have a strong cell wall, and it takes a lot
of labor to purify the target protein produced inside cells, and thus the production
amount is limited. Therefore, if the yeast can secrete the protein outside the cell
body, application of highly efficient production and continuous in vitro cultivation
as well as simplification of the industrial process by the ease of purification would
be possible. In order to release the produced protein extracellularly, the N-terminal
signal sequence of a known secretory protein must be connected to the protein of
interest. Zhou et al. (2018) reported that a P10L substitution in the signal sequence
of the INU1 gene increased the secretory expression of lignocellulolytic enzymes
in K. marxianus. The P10L substitution extended the hydrophobic core of the sig-
nal sequence and promoted secretion of mature proteins. Raimondi et al. (2010)
succeeded in extracellular secretion of superoxide dismutase (SOD) by fusion with
an appropriate signal sequence with SOD. In addition to enzymes, the production
of a single-chain Fv antibody (Nambu-Nishida et al. 2018) and virus-like particles
of porcine parvovirus (Yang et al. 2021) and porcine circovirus (Duan et al. 2019)
has been reported.
Kluyveromyces marxianus as a Platform in Synthetic Biology … 317
There are two major mechanisms by which cells can repair DNA double-strand
breaks (DSBs), which are intimately related to the classes of genetic recombina-
tion, homologous and non-homologous recombinations (Daley et al. 2005). The
core components of NHEJ required for rejoining of any DSB can be experimen-
tally defined as the proteins needed for simple religation. In the NHEJ pathway,
binding of the Ku heterodimer (Ku70/Ku80) to both ends of the DNA DSB with
the participation of additional proteins, such as Lig4, Nej1, and Lif1, promotes
repair of the DSB (Kooistra et al. 2004; Palmbos et al. 2005; Kegel et al. 2006;
Maassen et al. 2008). Ku, which is conserved from bacteria to humans, is an
indispensable protein for NHEJ (Doherty et al. 2001), and disruption of KU70 or
KU80, results in efficient gene targeting in K. lactis (Kooistra et al. 2004), P. stipitis
(Maassen et al. 2008), and K. marxianus (Abdel-Banat et al. 2010). In yeast lacking
KU70, a high frequency of non-homologous gene integration was abolished and
KmKU70 mutants showed 82–95% homologous gene targeting efficiency using
homologous sequences of 40–1000 bp, indicating that the highly efficient NHEJ
pathway can be used in random gene disruption techniques such as transposon
mutagenesis and plasmid-free gene manipulations in K. marxianus (Abdel-Banat
et al. 2010). In addition, the linear DNA integrative technique can eliminate the
burden of plasmid construction for targeted gene manipulation in K. marxianus.
in one of two ways: heterologous DNA is either incorporated into the genome
at a random locus (Kegel et al. 2006) by NHEJ, or the cassette is targeted to
a specific site on the genome by homologous recombination (HR) to the site of
interest (Lieber 2010; Löbs et al. 2017). In S. cerevisiae, HR is the primary DNA
repair pathway, and its high capabilities have made genomic engineering relatively
efficient and have facilitated the development of a wide range of in vivo DNA
assembly tools (Shao et al. 2009; Horwitz et al. 2015). However, in most other
yeasts, NHEJ is the preferred DNA repair pathway and genome engineering with
HR is inefficient. The high functionality of K. marxianus NHEJ can limit genome
editing in many applications. However, NHEJ-mediated functional marker selec-
tion as a novel DNA cloning method has been developed in the yeast (Hoshida
et al. 2014). Some researchers have taken advantage of this ability for multiplexed
gene integration (Cheon et al. 2014), but successful transformants differ signifi-
cantly in hexanoic acid-producing ability, probably due to gene insertion at crucial
genomic loci. Based on available genomic and transcriptomic data, Rajkumar et al.
(2019) characterized a set of native sequences in K. marxianus, including consti-
tutive and inducible promoters and terminators for expression of multiple genes,
for metabolic engineering and synthetic biology. Several known centromeres and
autonomous replication sequences (ARS) are also included in their collection.
These tools will serve as the basis for efficiently building next-generation cell
factories from this alternative yeast. A widely used strategy for enhancing HR in
non-conventional yeasts is the disruption of genes essential for the NHEJ path-
way such as KU70 or KU80. In K. marxianus, the disruption of KU70 and KU80
increased HR rates to 95% and 70%, respectively (Abdel-Banat et al. 2010; Choo
et al. 2014).
An alternative strategy for achieving efficient HR is the introduction of a
genomic DSB using a programmable endonuclease in the presence of a homol-
ogous repair template (Liu et al. 2017). Several programmable tools exist for a
targeted DSB, including dimeric meganucleases, zinc finger nucleases (ZFNs),
transcription activator-like effector nucleases (TALENs), and clustered regu-
larly interspaced short palindromic repeats (CRISPR) and CRISPR-associated 9
(CRISPR-Cas9) (Liu et al. 2017). The CRISPR-Cas9 gene-editing system has
been widely used in many yeasts including S. cerevisiae, Schizosaccharomyces
pombe, Yarrowia lipolytica, and Kluyveromyces lactis and recently in K. marxi-
anus (Nambu-Nishida et al. 2017; Löbs et al. 2017; Lee et al. 2018; Juergens
et al. 2018). The CRISPR-Cas9 system has been developed for use with K. marx-
ianus, enabling both NHEJ-based and HR-based genome editing (Cernak et al.
2018; Rajkumar et al. 2019). The genetic loci, ALPHA3 and KAT1, responsible for
mating-type switching in K. marxianus were identified and heterothallic haploid
strains were constructed using the CRISPR-Cas9 system. This is indispensable
for genetic manipulation of the most desired traits, which are likely to depend on
multiple unlinked genetic loci, and remain difficult to identify without the ability
to carry out genetic crosses. Three complex traits, the ability to take up exoge-
nous DNA, thermotolerance, and high lipid production, have been successfully
combined into a single K. marxianus isolate (Cernak et al. 2018).
Kluyveromyces marxianus as a Platform in Synthetic Biology … 319
A non-conventional yeast can increase the variety and complexity of aroma pro-
files of bakery products such as nuts and fruity aroma (Aslankoohi et al. 2016)
and alcoholic beverages such as wine and beer (Gamero et al. 2020). In addition
to producing ethanol, K. marxianus is capable of producing a variety of volatile
molecules or aromatic esters used as fragrances or flavors (Table 1). The ability
of the yeast to produce acetate esters such as 2-phenyl ethyl acetate (2-PEA) and
isoamyl acetate from 2-phenyl ethanol (2-PE) and isoamyl alcohol is due to the
presence of alcohol acetyltransferase or AATase (Gethins et al. 2015). The ability
to produce 2-phenylethanol from glucose without an additional L-phenylalanine
supplement can be achieved in K. marxianus by genetic engineering via over-
expression of ARO10 for phenylpyruvate decarboxylase and ADH2 for alcohol
dehydrogenase II from S. cerevisiae (Kim et al. 2014). Genes involved in ethyl
acetate biosynthesis in K. marxianus were identified (Löbs et al. 2017). KmAdh2
was found to be critical for aerobic and anaerobic ethanol production. Aerobically
produced ethanol is supplied for the biosynthesis of ethyl acetate catalyzed by
KmAtf. KmAdh7 was found to exhibit activity toward the oxidation of hemiacetal,
a possible alternative route for the synthesis of ethyl acetate.
9.2 Fructose
Xylitol has gained increasing attention in recent years due to its use in several
industries such as food, dental products, and pharmaceuticals (Ravella et al. 2012).
The ability of K. marxianus to ferment xylose under oxygen-limited conditions
is weak due to its redox imbalance. Xylose consumption and fermentation can
320 N. Lertwattanasakul et al.
10 Conclusions
80
70
60
50
40
30
20
10
0
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010
2011
2012
2013
2014
2015
2016
2017
2018
2019
2020
Fig. 2 Changes in the number of publications related to K. marxianus (straight lines) and the
number of publications related to K. lactis (broken lines). Analysis was performed using PubMed
Acknowledgements This study was supported by the Advanced Low Carbon Technology Research
and Development Program, which was granted by the Japan Science and Technology Agency
(JPMJAL1106) (MM, TK, and MY) and e-ASIA Joint Research Program, which was granted by
Japan Science (JPMJSC16E5) (MM, TK, and MY) and Technology Agency, Ministry of Research,
Technology and Higher Education of the Republic of Indonesia, Agricultural Research Development
322 N. Lertwattanasakul et al.
Agency of Thailand and Ministry of Science and Technology of Laos, and partially supported by the
Core to Core Program A. Advanced Research Networks, which was granted by the Japan Society
for the Promotion of Science, the National Research Council of Thailand, Ministry of Science and
Technology in Vietnam, National Univ. of Laos, Univ. of Brawijaya and Beuth Univ. of Applied
Science Berlin (NL, MM, TK, and MY), and the Japan Society for the Promotion of Science,
MEXT/JSPS Kakenhi (25250028 and 16H02485 to MY).
References
Abbott DA, Zelle RM, Pronk JT, van Maris AJA (2009) Metabolic engineering of Saccharomyces
cerevisiae for production of carboxylic acids: current status and challenges. FEMS Yeast Res
9(8):1123–1136. https://doi.org/10.1111/j.1567-1364.2009.00537.x
Abdel-Banat BMA, Nonklang S, Hoshida H, Akada R (2010) Random and targeted gene integra-
tions through the control of non-homologous end joining in the yeast Kluyveromyces marxi-
anus. Yeast 27(1):29–39. https://doi.org/10.1002/yea.1729
Anandharaj M, Lin Y-J, Rania RP, Nadendla EK, Ho M-C, Huang C-C, Cheng J-F, Chang J-J, Li
W-H (2020) Constructing a yeast to express the largest cellulosome complex on the cell surface.
PNAS 117(5):2385–2394. https://doi.org/10.1073/pnas.1916529117
Anderson P, McNeil K, Watson, K (1986) High-efficiency carbohydrate fermentation to ethanol
at temperatures above 40 ◦ C by Kluyveromyces marxianus var. marxianus isolated from sugar
mills. Appl Environ Microbiol (USA) 51(6):1314–1320
Anggarini S, Murata M, Kido K, Kosaka T, Sootsuwan K, Thanonkeo P, Yamada M (2019)
Improvement of thermotolerance of Zymomonas mobilis by genes for reactive oxygen species-
scavenging enzymes and heat shock proteins. Front Microbiol 10:3073. https://doi.org/10.3389/
fmicb.2019.03073
Aslankoohi E, Herrera-Malaver B, Rezaei MN, Steensels J, Courtin CM, Verstrepen KJ (2016)
Non-conventional yeast strains increase the aroma complexity of bread. PLoS ONE 11(10):1–
18. https://doi.org/10.1371/journal.pone.0165126
Ballesteros I, Ballesteros M, Cabanas A, Carrasco J, Martin C, Negro M, Saez R (1991) Selection
of thermotolerant yeasts for simultaneous saccharification and fermentation (SSF) of cellulose
to ethanol. Appl Biochem Biotechnol (USA) 28:307–315
Banat I, Nigam P, Singh D, Marchant R, McHale A (1998) Ethanol production at elevated tem-
peratures and alcohol concentrations: part I—yeasts in general. World J Microbiol Biotechnol
(united Kingdom) 14(6):809–821
Banat IM, Nigam P, Marchant R (1992) Isolation of thermotolerant, fermentative yeasts growing
at 52 °C and producing ethanol at 45 and 50 °C. World J Microbiol Biotechnol 8(3):259–263.
https://doi.org/10.1007/BF01201874
Behera S, Sharma N, Arora R, Kumar S (2016) Effect of evolutionary adaption on xylosidase activ-
ity in thermotolerant yeast isolates Kluyveromyces marxianus NIRE-K1 and NIRE-K3. Appl
Biochem Biotechnol 179(7):1143–1154. https://doi.org/10.1007/s12010-016-2055-2
Bergkamp RJM, Bootsman TC, Toschka HY, Mooren ATA, Kox L, Verbakel JMA, Geerse RH,
Planta RJ (1993) Expression of an alpha-galactosidase gene under control of the homolo-
gous inulinase promoter in Kluyveromyces marxianus. Appl Microbiol Biotechnol 40:309–317.
https://doi.org/10.1007/BF00170386
Blank LM, Lehmbeck F, Sauer U (2005) Metabolic-flux and network analysis in fourteen hemias-
comycetous yeasts. FEMS Yeast Res 5:545–558
Bollók M, Réczey K, Zacchi G (2000) Simultaneous saccharification and fermentation of steam-
pretreated spruce to ethanol. Appl Biochem Biotechnol 84–6:69–80. https://doi.org/10.1385/
ABAB:84-86:1-9:69
Bragança CRS, Colombo LT, Roberti AS, Alvim MCT, Cardoso SA, Reis KCP, de Paula SO, da
Silveira WB, Passos FML (2015) Construction of recombinant Kluyveromyces marxianus UFV-
3 to express dengue virus type 1 nonstructural protein 1 (NS1). Appl Microbiol Biotechnol
99:1191–1203. https://doi.org/10.1007/s00253-014-5963-5
Kluyveromyces marxianus as a Platform in Synthetic Biology … 323
Camargo D, Gomes S, Sene L (2014) Ethanol production from sunflower meal biomass by simul-
taneous saccharification and fermentation (SSF) with Kluyveromyces marxianus ATCC 36907.
Bioprocess Biosyst Eng 37(11):2235–2242. https://doi.org/10.1007/s00449-014-1201-x
Campos-García J, Vargas A, Farías-Rosales L, Miranda AL, Meza-Carmen V, Díaz-Pérez AL
(2018) Increase in fatty acid ethyl ester content through ATF1 expression in an engineered
Kluyveromyces marxianus UMPe-1 yeast improves the organoleptic properties of a craft Mez-
cal beverage. J Agric Food Chem 66(17):4469–4480. https://doi.org/10.1021/acs.jafc.8b00730
Castro R, Roberto I (2014) Selection of a thermotolerant Kluyveromyces marxianus strain with
potential application for cellulosic ethanol production by simultaneous saccharification and fer-
mentation. Appl Biochem Biotechnol 172(3):1553–1564. https://doi.org/10.1007/s12010-013-
0612-5
Catalá A (2012) Lipid peroxidation modifies the picture of membranes from the “Fluid Mosaic
Model” to the “Lipid Whisker Model.” Biochimie 94(1):101–109. https://doi.org/10.1016/j.bio
chi.2011.09.025
Celton M, Sanchez I, Goelzer A, Fromion V, Camarasa C, Dequin S (2012) A comparative tran-
scriptomic, fluxomic and metabolomic analysis of the response of Saccharomyces cerevisiae to
increases in NADPH oxidation. BMC Genomics 13(1):317–330. https://doi.org/10.1186/1471-
2164-13-317
Cernak P, Estrela R, Poddar S, Skerker JM, Cheng Y-F, Carlson AK, Chen B, Glynn VM, Furlan M,
Ryan OW, Donnelly MK, Arkin AP, Taylor JW, Cate JHD (2018) Engineering Kluyveromyces
marxianus as a robust synthetic biology platform host. mBio 9(5):e01410– e01418.https://doi.
org/10.1128/mBio.01410-18
Chamnipa N, Thanonkeo S, Klanrit P, Thanonkeo P (2018) The potential of the newly isolated
thermotolerant yeast Pichia kudriavzevii RZ8-1 for high-temperature ethanol production. Braz
J Microbiol 49(2):378–391. https://doi.org/10.1016/j.bjm.2017.09.002
Chang J-J, Ho C-Y, Ho F-J, Tsai T-Y, Ke H-M, Wang CH, Li W-H (2012) PGASO: A synthetic
biology tool for engineering a cellulolytic yeast. Biotechnol Biofuels 5(1):53–64. https://doi.
org/10.1186/1754-6834-5-53
Chang J-J, Ho F-J, Ho C-Y, Wu Y-C, Hou Y-H, Huang C-C, Shih M-C, Li W-H (2013) Assembling
a cellulase cocktail and a cellodextrin transporter into a yeast host for CBP ethanol production.
Biotechnol Biofuels 6(1):1–13. https://doi.org/10.1186/1754-6834-6-19
Chang J-J, Lin Y-J, Lay C-H, Thia C, Wu Y-C, Hou Y-H, Huang C-C, Li W-H (2017) Constructing
a cellulosic yeast host with an efficient cellulase cocktail. Biotechnol Bioeng 1–11. https://doi.
org/10.1002/bit.26507
Chang J-J, Thia C, Lin H-Y, Liu H-L, Ho F-J, Wu J-T, Shih M-C, Li W-H, Huang C-C (2015) Inte-
grating an algal β-carotene hydroxylase gene into a designed carotenoid-biosynthesis pathway
increases carotenoid production in yeast. Biores Technol 184:2–8. https://doi.org/10.1016/j.bio
rtech.2014.11.097
Charoensopharat K, Thanonkeo P, Thanonkeo S, Yamada M (2015) Ethanol production from
Jerusalem artichoke tubers at high temperature by newly isolated thermotolerant inulin-
utilizing yeast Kluyveromyces marxianus using consolidated bioprocessing. Antonie Van
Leeuwenhoek 108(1):173–190. https://doi.org/10.1007/s10482-015-0476-5
Charoensuk K, Sakurada T, Tokiyama A, Murata M, Kosaka T, Thanonkeo P, Yamada M (2017)
Thermotolerant genes essential for survival at a critical high temperature in thermotolerant
ethanologenic Zymomonas mobilis TISTR 548. Biotechnol Biofuels 10(1):1–11. https://doi.
org/10.1186/s13068-017-0891-0
Cheon Y, Kim J-S, Park J-B, Heo P, Lim JH, Jung GY, Seo J-H, Park JH, Koo HM, Cho KM,
Park J-B, Ha S-J, Kweon D-H (2014) A biosynthetic pathway for hexanoic acid production
in Kluyveromyces marxianus. J Biotechnol 182–183:30–36. https://doi.org/10.1016/j.jbiotec.
2014.04.010
Chi Z-M, Zhang T, Cao T-S, Liu X-Y, Cui W, Zhao C-H (2011) Biotechnological potential of
inulin for bioprocesses. Biores Technol 102(6):4295–4303. https://doi.org/10.1016/j.biortech.
2010.12.086
324 N. Lertwattanasakul et al.
Choo JH, Han C, Kim J-Y, Kang HA (2014) Deletion of a KU80 homolog enhances homolo-
gous recombination in the thermotolerant yeast Kluyveromyces marxianus. Biotechnol Lett
36(10):2059–2067. https://doi.org/10.1007/s10529-014-1576-4
Choudhary J, Singh S, Nain L (2016) Thermotolerant fermenting yeasts for simultaneous saccha-
rification fermentation of lignocellulosic biomass. Electron J Biotechnol 21:82–92. https://doi.
org/10.1016/j.ejbt.2016.02.007
Cruz-Guerrero A, Garcia-Peña I, Barzana E, Garcia-Garibay M, Gomez-Ruiz L (1995)
Kluyveromyces marxianus CDBB-L-278: a wild inulinase hyperproducing strain. J
Fermentation Bioeng (japan) 80(2):159–163
Daley JM, Palmbos PL, Wu D, Wilson TE (2005) Nonhomologous end joining in yeast. Annu Rev
Genet 39:431–451
Doherty AJ, Jackson SP, Weller GR (2001) Identification of bacterial homologues of the Ku DNA
repair proteins. FEBS Lett 500:186–188
Du C, Li Y, Zhao X, Pei X, Yuan W, Bai F, Jiang Y (2019) The production of ethanol from ligno-
cellulosic biomass by Kluyveromyces marxianus CICC 1727–5 and Spathaspora passalidarum
ATCC MYA-4345. Appl Microbiol Biotechnol 103(6):2845–2855. https://doi.org/10.1007/s00
253-019-09625-1
Duan J, Yang D, Chen L, Yu Y, Zhou J, Lu H (2019) Efficient production of porcine circovirus
virus-like particles using the nonconventional yeast Kluyveromyces marxianus. Appl Environ
Microbiol 103:833–842
Elghanam S-M, Attia S-A, Shoeb A-H, Hashem MA-E (2012) Expression and purification of
hepatitis B surface antigen S from Escherichia coli; a new simple method. BMC Res Notes
5:125–133
Feng J, Wan R, Yi Q, He L, Yang L, Tang L (2015) Examination of alternate codon bias solu-
tions for expression and purification of recombinant mechano-growth factor in Escherichia coli.
Biotechnol Appl Biochem 62(5):690–698. https://doi.org/10.1002/bab.1312
Fields PA (2001) Review: protein function at thermal extremes: balancing stability and flexibil-
ity. Comparative Biochem Physiol Part A 129(2):417–431. https://doi.org/10.1016/S1095-643
3(00)00359-7
Fonseca GG, Heinzle E, Wittmann C, Gombert AK (2008) The yeast Kluyveromyces marxianus
and its biotechnological potential. Appl Microbiol Biotechnol 79(3):339–354. https://doi.org/
10.1007/s00253-008-1458-6
Frick O, Wittmann C (2005) Characterization of the metabolic shift between oxidative and fermen-
tative growth in Saccharomyces cerevisiae by comparative 13 C flux analysis. Microb Cell Fact
4:1–16. https://doi.org/10.1186/1475-2859-4-30
Fu X, Li P, Zhang L, Li S (2019) Understanding the stress responses of Kluyveromyces marxianus
after an arrest during high-temperature ethanol fermentation based on integration of RNA-Seq
and metabolite data. Appl Microbiol Biotechnol 103(6):2715–2729. https://doi.org/10.1007/
s00253-019-09637-x
Galbe M, Zacchi G (2007) Pretreatment of lignocellulosic materials for efficient bioethanol pro-
duction. Adv Biochem Eng Biotechnol 108:41–65. https://doi.org/10.1007/10_2007_070
Gamero A, Dijkstra A, Smit B, de Jong C (2020) Aromatic potential of diverse non-conventional
yeast species for winemaking and brewing. Fermentation 6(2):50. https://doi.org/10.3390/fer
mentation6020050
Gao J, Yuan W, Li Y, Xiang R, Hou S, Zhong S, Bai F (2015) Transcriptional analysis of
Kluyveromyces marxianus for ethanol production from inulin using consolidated bioprocessing
technology. Biotechnol Biofuels 8(1):1–17. https://doi.org/10.1186/s13068-015-0295-y
Gao M-J, Zheng Z-Y, Wu J-R, Dong S-J, Li Z, Jin H, Zhan X-B, Lin C-C (2012) Improvement
of specific growth rate of Pichia pastoris for effective porcine interferon-α production with an
on-line model-based glycerol feeding strategy. Appl Microbiol Biotechnol 93(4):1437–1445.
https://doi.org/10.1007/s00253-011-3605-8
Gethins L, Guneser O, Demirkol A, Rea MC, Stanton C, Ross RP, Yuceer Y, Momissey JP
(2015) Influence of carbon and nitrogen sources on production of volatile fragrance and flavour
Kluyveromyces marxianus as a Platform in Synthetic Biology … 325
Kim S-B, Kwon D-H, Park J-B, Ha S-J (2019) Alleviation of catabolite repression in
Kluyveromyces marxianus: the thermotolerant SBK1 mutant simultaneously coferments
glucose and xylose. Biotechnol Biofuels 12:90. https://doi.org/10.1186/s/s13068-019-1431-x
Kim S, Kim CH (2014) Evaluation of whole Jerusalem artichoke (Helianthus tuberosus L.) for a
consolidated bioprocessing ethanol production. Renewable Energy 65:83–91
Kim S, Park JM, Kim CH (2013) Ethanol production using whole plant biomass of jerusalem arti-
choke by Kluyveromyces marxianus CBS1555. Appl Biochem Biotechnol 169(5):1531–1545.
https://doi.org/10.1007/s12010-013-0094-5
Kim T-Y, Lee S-W, Oh M-K (2014) Biosynthesis of 2-phenylethanol from glucose with genetically
engineered Kluyveromyces marxianus. Enzyme Microb Technol 61–62:44–47. https://doi.org/
10.1016/j.enzmictec.2014.04.011
Kong X, Zhang B, Hua Y, Zhu Y, Li W, Wang D, Hong J (2019) Efficient L-lactic acid production
from corncob residue using metabolically engineered thermo-tolerant yeast. Biores Technol
273:220–230. https://doi.org/10.1016/j.biortech.2018.11.018
Konings AW (1988) Importance of the glutathione level and the activity of the pentose phosphate
pathway in cellular heat sensitivity. Recent Results in Cancer Research Fortschritte Der Krebs-
forschung Progres Dans Les Recherches Sur Le Cancer 109:109–125. https://doi.org/10.1007/
978-3-642-83263-5_14
Kooistra R, Hooykaas PJJ, Steensma HY (2004) Efficient gene targeting in Kluyveromyces lactis.
Yeast 21:781–792
Kosaka T, Lertwattanasakul N, Rodrussamee N, Nurcholis M, Dung NTP, Keo-Oudone C, Murata
M, Götz P, Theodoropoulos C, Suprayogi; Maligan JM, Limtong S, Yamada M (2018) Potential
of thermotolerant ethanologenic yeasts isolated from ASEAN countries and their application in
high-temperature fermentation. Book Chapter IntechOpen. https://doi.org/10.5772/intechopen.
79144
Kosaka T, Nakajima Y, Ishii A, Yamashita M, Yoshida S, Murata M, Kato K, Shiromaru Y, Kato
S, Kanasaki Y, Yoshikawa H, Matsutani M, Thanonkeo P, Yamada M (2019) Capacity for sur-
vival in global warming: adaptation of mesophiles to the temperature upper limit. PLoS ONE
14(5):1–15. https://doi.org/10.1371/journal.pone.0215614
Koutinas M, Patsalou M, Stavrinou S, Vyrides I (2016) High temperature alcoholic fermentation
of orange peel by the newly isolated thermotolerant Pichia kudriavzevii KVMP10. Lett Appl
Microbiol 62(1):75–83. https://doi.org/10.1111/lam.12514
Krishna SH, Reddy TJ, Chowdary G (2001) Simultaneous saccharification and fermentation of
lignocellulosic wastes to ethanol using a thermotolerant yeast. Biores Technol 77(2):193–196
Kuhn A, Van Zyl C, Van Tonder A, Prior BA (1995) Purification and partial characterization of an
aldo-keto reductase from Saccharomyces cerevisiae. Appl Environ Microbiol 61(4):1580–1585.
https://doi.org/10.1128/aem.61.4.1580-1585.1995
Kumar R, Tabatabaei M, Karimi K, Sárvári Horváth I (2016) Recent updates on lignocellulosic
biomass derived ethanol—a review. Biofuel Res J 3(1):347–356
Kumar S, Singh SP, Mishra IM, Adhikari DK (2009) Ethanol and xylitol production from glucose
and xylose at high temperature by Kluyveromyces sp. IIPE453. J Industr Microbiol Biotechnol
36(12):1483–1489. https://doi.org/10.1007/s10295-009-0636-6
Kwon D-H, Kim S-B, Park J-B, Ha S-J (2020) Overexpression of mutant galactose permease
(ScGal2_N376F) effective for utilization of glucose/xylose or glucose/galactose mixture by
engineered Kluyveromyces marxianus. J Microbiol Biotechnol 30(12):1944–1949. https://doi.
org/10.4014/jmb.2008.08035
Kwon D-H, Park J-B, Hong E, Ha S-J (2019) Ethanol production from xylose is highly increased
by the Kluyveromyces marxianus mutant 17694-DH1. Bioprocess Biosyst Eng 42(1):63–70.
https://doi.org/10.1007/s00449-018-2014-0
Lane MM, Morrissey JP (2010) Kluyveromyces marxianus: a yeast emerging from its sister’s
shadow. Fungal Biol Rev 24(1):17–26. https://doi.org/10.1016/j.fbr.2010.01.001
Lee JW, In JH, Park J-B, Shin J, Park JH, Sung BH, Sohn J-H, Seo J-H, Park J-B, Kim SR, Kweon
D-H (2017) Co-expression of two heterologous lactate dehydrogenases genes in Kluyveromyces
328 N. Lertwattanasakul et al.
Löser C, Urit T, Keil P, Bley T (2015) Studies on the mechanism of synthesis of ethyl acetate in
Kluyveromyces marxianus DSM 5422. Appl Microbiol Biotechnol 99(3):1131–1144. https://
doi.org/10.1007/s00253-014-6098-4
Löser C, Urit T, Stukert A, Bley T (2013) Formation of ethyl acetate from whey by Kluyveromyces
marxianus on a pilot scale. J Biotechnol 163(1):17–23. https://doi.org/10.1016/j.jbiotec.2012.
10.009
Lulu L, Dongmei W, Xiaolian G, Jiong H, Ling Z, Hisanori T, Hidehiko K (2013) Identification
of a xylitol dehydrogenase gene from Kluyveromyces marxianus NBRC1777. Mol Biotechnol
53(2):159–169. https://doi.org/10.1007/s12033-012-9508-9
Maassen N, Freese S, Schruff B, Passoth V, Klinner U (2008) Nonhomologous end joining
and homologous recombination DNA repair pathways in integration mutagenesis in xylose-
fermenting yeast Pichia stipitis. FEMS Yeast Res 8:735–743
Madeo F, Fröhlich KU, Fröhlich E, Ligr M, Wolf DH, Grey M, Sigrist SJ (1999) Oxygen stress:
a regulator of apoptosis in yeast. J Cell Biol 145(4):757–767. https://doi.org/10.1083/jcb.145.
4.757
Mahmoud AE, Fathy SA, Rashad MM, Ezz MK, Mohammed AT (2018) Purification and character-
ization of a novel tannase produced by Kluyveromyces marxianus using olive pomace as solid
support, and its promising role in gallic acid production. Int J Biol Macromol 107:2342–2350.
https://doi.org/10.1016/j.ijbiomac.2017.10.117
Marcišauskas S, Ji B, Nielsen J (2019) Reconstruction and analysis of a Kluyveromyces marxi-
anus genome-scale metabolic model. BMC Bioinform 20(1). https://doi.org/10.1186/s12859-
019-3134-5
Margeot A, Hahn-Hagerdal B, Edlund M, Slade R, Monot F (2009) New improvements for lig-
nocellulosic ethanol. Curr Opin Biotechnol 20(3):372–380. https://doi.org/10.1016/j.copbio.
2009.05.009
Martins DBG, de Souza Jr CG, Simões DA, de Morais Jr MA (2002) The β-galactosidase activ-
ity in Kluyveromyces marxianus CBS6556 decreases by high concentrations of galactose. Curr
Microbiol 44(5):379–382. https://doi.org/10.1007/s00284-001-0052-2
Matsushika A, Inoue H, Kodaki T, Sawayama S (2009) Ethanol production from xylose in
engineered Saccharomyces cerevisiae strains: current state and perspectives. Appl Microbiol
Biotechnol 84(1):37–53. https://doi.org/10.1007/s00253-009-2101-x
Matsushita K, Azuma Y, Kosaka T, Yakushi T, Hoshida H, Akada R, Yamada M (2016) Genomic
analyses of thermotolerant microorganisms used for high-temperature fermentations. Biosci
Biotechnol Biochem 80(4):655–668. https://doi.org/10.1080/09168451.2015.1104235
Matsuzaki C, Nakagawa A, Koyanagi T, Tanaka K, Minami H, Tamaki H, Katayama T, Yamamoto
K, Kumagai H (2012) Kluyveromyces marxianus-based platformfor direct ethanol fermenta-
tion and recovery from cellulosic materials under air-ventilated conditions. J Biosci Bioeng
113:604–607. https://doi.org/10.1016/j.jbiosc.2011.12.007
Mattanovich D, Branduardi P, Dato L, Gasser B, Sauer M, Porro D (2012) Recombinant protein
production in yeasts. Methods Mol Biol 824:329–358. https://doi.org/10.1007/978-1-61779-
433-9_17
McTaggart TL, Bever D, Bassett S, Da Silva NA (2019) Synthesis of polyketides from low
cost substrates by the thermotolerant yeast Kluyveromyces marxianus. Biotechnol Bioeng
116:1721–1730. https://doi.org/10.1002/bit.26976
Mejía-Barajas J, Montoya-Pérez R, Manzo-Avalos S, Cortés-Rojo C, Riveros-Rosas H, Cervantes
C, Saavedra-Molina A (2018) Fatty acid addition and thermotolerance of Kluyveromyces marx-
ianus. FEMS Microbiol Lett 365(7):1–5. https://doi.org/10.1093/femsle/fny043
Mejía-Barajas JA, Montoya-Pérez R, Salgado-Garciglia R, Aguilera-Aguirre L, Cortés-Rojo C,
Mejía-Zepeda R, Arellano-Plaza M, Saavedra-Molina A (2017) Oxidative stress and antiox-
idant response in a thermotolerant yeast. Braz J Microbiol 48(2):326–332. https://doi.org/10.
1016/j.bjm.2016.11.005
Mo W, Wang M, Zhan R, Yu Y, Lu H, He Y (2019) Kluyveromyces marxianus developing
ethanol tolerance during adaptive evolution with significant improvements of multiple path-
ways. Biotechnol Biofuels 12(1). https://doi.org/10.1186/s13068-019-1393-z
330 N. Lertwattanasakul et al.
Rocha SN, Abrahão-Neto J, Cerdán ME, González-Siso MI, Gombert AK (2010) Heterologous
expression of glucose oxidase in the yeast Kluyveromyces marxianus. Microb Cell Factories
9:1–12. https://doi.org/10.1186/1475-2859-9-4
Rodrussamee N, Lertwattanasakul N, Hirata K, Limtong S, Kosaka T, Yamada M (2011) Growth
and ethanol fermentation ability on hexose and pentose sugars and glucose effect under vari-
ous conditions in thermotolerant yeast Kluyveromyces marxianus. Appl Microbiol Biotechnol
90(4):1573–1586. https://doi.org/10.1007/s00253-011-3218-2
Rollero S, Bauer FF, Divol B, Bloem A, Camarasa C, Ortiz-Julien A (2019) A comparison of
the nitrogen metabolic networks of Kluyveromyces marxianus and Saccharomyces cerevisiae.
Environ Microbiol 21(11):4076–4091. https://doi.org/10.1111/1462-2920.14756
Roohina F, Mohammadi M, Najafpour GD (2016) Immobilized Kluyveromyces marxianus cells in
carboxymethyl cellulose for production of ethanol from cheese whey: experimental and kinetic
studies. Bioprocess Biosyst Eng 39(9):1341–1349. https://doi.org/10.1007/s00449-016-1610-0
Rosano GL, Ceccarelli EA (2014) Recombinant protein expression in Escherichia coli: advances
and challenges. Front Microbiol 5:172. https://doi.org/10.3389/fmicb.2014.00172
Rouhollah H, Iraj N, Giti E, Sorah A (2007) Mixed sugar fermentation by Pichia stipitis, Sac-
charomyces cerevisiae, and an isolated xylose-fermenting Kluyveromyces marxianus and their
cocultures. Afr J Biotech 6(9):1110–1114
Rouwenhorst RJ, Visser LE, Van Der Baan AA, Scheffers WA, Van Dijken JP (1988) Production,
distribution, and kinetic properties of inulinase in continuous cultures of Kluyveromyces marx-
ianus CBS 6556. Appl Environ Microbiol 54(5):1131–1137. https://doi.org/10.1128/AEM.54.
5.1131-1137.1988
Ryabova OB, Chmil OM, Sibirny AA (2003) Xylose and cellobiose fermentation to ethanol by the
thermotolerant methylotrophic yeast Hansenula polymorpha. FEMS Yeast Res 4(2):157–164.
https://doi.org/10.1016/S1567-1356(03)00146-6
Saet-Byeol K, Deok-Ho K, Jae-Bum P, Suk-Jin H (2019) Alleviation of catabolite repression in
Kluyveromyces marxianus: the thermotolerant SBK1 mutant simultaneously coferments glu-
cose and xylose. Biotechnol Biofuels 12(1):1–9. https://doi.org/10.1186/s13068-019-1431-x
Sansonetti S, Curcio S, Calabrò V, Iorio G (2009) Bio-ethanol production by fermentation of ricotta
cheese whey as an effective alternative non-vegetable source. Biomass Bioenerg 33(12):1687–
1692. https://doi.org/10.1016/j.biombioe.2009.09.002
Schabort DTW, Letebele PK, Steyn L, Kilian SG, du Preez JC (2016) Differential RNA-seq, multi-
network analysis and metabolic regulation analysis of Kluyveromyces marxianus reveals a
compartmentalised response to xylose. PLoS ONE 11(6):1–31. https://doi.org/10.1371/journal.
pone.0156242
Scherz-Shouval R, Elazar Z (2007) ROS, mitochondria and the regulation of autophagy. Trends
Cell Biol 17(9):422–427. https://doi.org/10.1016/j.tcb.2007.07.009
Shallom D, Shoham Y (2003) Microbial hemicellulases. Curr Opin Microbiol 6(3):219–228.
https://doi.org/10.1016/S1369-5274(03)00056-0
Shao Z, Zhao H, Zhao H. (2009) DNA assembler, an in vivo genetic method for rapid construction
of biochemical pathways. Nucleic Acids Res 37(2):e16
Sharma NK, Behera S, Arora R, Kumar S (2016) Enhancement in xylose utilization using
Kluyveromyces marxianus NIRE-K1 through evolutionary adaptation approach. Bioprocess
Biosyst Eng 39(5):835–843. https://doi.org/10.1007/s00449-016-1563-3
Sharma NK, Behera S, Arora R, Kumar S (2017) Evolutionary adaptation of Kluyveromyces marx-
ianus NIRE-K3 for enhanced xylose utilization. Front Energy Res 5(DEC). https://doi.org/10.
3389/fenrg.2017.00032
Šiekštele R, Bartkevičiute D, Sasnauskas K (1999) Cloning, targeted disruption and heterol-
ogous expression of the Kluyveromyces marxianus endopolygalacturonase gene (EPG1).
Yeast 15(4):311–322. https://doi.org/10.1002/(SICI)1097-0061(19990315)15:4%3c311::AID-
YEA379%3e3.0.CO;2-9
Silveira WB, Diniz RH, Cerdán ME, González-Siso MI, Souza RA, Vidigal PM, Brustolini OJ,
de Almeida Prata ER, Medeiros AC, Paiva LC, Nascimento M, Ferreira EG, Dos Santos VC,
Bragança CR, Fernandes TA, Colombo LT, Passos FM (2014) Genomic sequence of the yeast
Kluyveromyces marxianus as a Platform in Synthetic Biology … 333
Kluyveromyces marxianus CCT 7735 (UFV-3), a highly lactose-fermenting yeast isolated from
the Brazilian dairy industry. Genome Announcements 2. https://doi.org/10.1128/genomeA.011
36-14
Simoness O, Murilol B, Carlosr R, Paulalde A, Mariackv R, Franciscorde A-N, Soreleb F, Luizars
D (2017) Asymmetric bioreduction of β-ketoesters derivatives by Kluyveromyces marxianus:
Influence of molecular structure on the conversion and enantiomeric excess. An Acad Bras
Ciênc 89:1403–1415. https://doi.org/10.1590/0001-3765201720170118
Sivarathnakumar S, Jayamuthunagai J, Baskar G, Praveenkumar R, Selvakumari IAE, Bharathiraja
B (2019) Bioethanol production from woody stem Prosopis juliflora using thermo tolerant yeast
Kluyveromyces marxianus and its kinetics studies. Biores Technol 293. https://doi.org/10.1016/
j.biortech.2019.122060
Soemphol W, Deeraksa A, Matsutani M, Yakushi T, Toyama H, Adachi O, Yamada M, Matsushita
K (2011) Global analysis of the genes involved in the thermotolerance mechanism of thermo-
tolerant Acetobacter tropicalis SKU1100. Biosci Biotechnol Biochem 75(10):1921–1928
Sokolenko G, Karpechenko N (2015) Expression of inulinase genes in the yeasts Saccharomyces
cerevisiae and Kluyveromyces marxianus. Microbiology (00262617) 84(1):23–27.https://doi.
org/10.1134/S0026261715010142
Steels EL, Learmonth RP, Watson K (1994) Stress tolerance and membrane lipid unsaturation in
Saccharomyces cerevisiae grown aerobically or anaerobically. Microbiology 140(3):569–576
Storeng R, Johne B (1987) Toxic effects of lipopolysaccharide from Bacteroides intermedius and
Escherichia coli assessed in the pre-implantation mouse embryo culture system. Acta Pathol
Microbiol Immunol Scand B 95:135–139
Su M, Hu Y, Cui Y, Wang Y, Yu H, Liu J, Dai W, Piao C (2021) Production of b-glucosidase from
okara fermentation using Kluyveromyces marxianus. J Food Sci Technol 58:366–376. https://
doi.org/10.1007/s13197-020-04550-y
Sugiyama KI, Izawa S, Inoue Y (2000) The Yap1p-dependent induction of glutathione synthesis in
heat shock response of Saccharomyces cerevisiae. J Biol Chem 275(20):15535–15540. https://
doi.org/10.1074/jbc.275.20.15535
Suryawati L, Wilkins MR, Bellmer DD, Huhnke RL, Maness NO, Banat IM (2008) Simultaneous
saccharification and fermentation of Kanlow switchgrass pretreated by hydrothermolysis using
Kluyveromyces marxianus IMB4. Biotechnol Bioeng 101(5):894–902. https://doi.org/10.1002/
bit.21965
Suzuki T, Hoshino T, Matsushika A (2014) Draft genome sequence of Kluyveromyces marxianus
strain DMB1, isolated from sugarcane bagasse hydrolysate. Genome Announcements 2. https://
doi.org/10.1128/genomeA.00733-14
Suzuki T, Hoshino T, Matsushika A (2019) High-temperature ethanol production by a series of
recombinant xylose-fermenting Kluyveromyces marxianus strains. Enzyme Microbial Technol
129. https://doi.org/10.1016/j.enzmictec.2019.109359
Szczodrak J, Targonski Z (1988) Selection of thermotolerant yeast strains for simultaneous saccha-
rification and fermentation of cellulose. Biotechnol Bioeng (USA) 31(4):300–303
Tarrío N, Becerra M, Cerdan ME, González-Siso MI (2006) Reoxidation of cytosolic NADPH in
Kluyveromyces lactis. FEMS Yeast Res 6(3):371–380
Tarrío N, García-Leiro A, Cerdán ME, González-Siso MI (2008) The role of glutathione reduc-
tase in the interplay between oxidative stress response and turnover of cytosolic NADPH in
Kluyveromyces lactis. FEMS Yeast Res 8(4):597–606. https://doi.org/10.1111/j.1567-1364.
2008.00366.x
Theron CW, Labuschagné M, Gudiminchi R, Albertyn J, SmitMS, (2014) A broad-range yeast
expression system reveals Arxula adeninivorans expressing a fungal self-sufficient cytochrome
P450 monooxygenase as an excellent whole-cell biocatalyst. FEMS Yeast Res 14:556–566.
https://doi.org/10.1111/1567-1364.12142
Topete M, Casas TL, Galindo E (1997) β-Galactosidase production by Kluyveromyces marxianus
cultured in shake flasks. Rev Latinoam Microbiol 39:101–107
334 N. Lertwattanasakul et al.
Van Urk H, Voll WS, Scheffers WA, Van Dijken JP (1990) Transient-state analysis of metabolic
fluxes in crabtree-positive and crabtree-negative yeasts. Appl Environ Microbiol 56(1):281–
287. https://doi.org/10.1128/AEM.56.1.281-287.1990
Vandijken JP, Weusthuis RA, Pronk JT (1993) Kinetics of growth and sugar consumption in yeasts.
Antonie Van Leeuwenhoek Int J Gen Mol Microbiol 63:343–352
Varela JA, Gethins L, Stanton C, Ross P, Morrissey JP (2017) Applications of Kluyveromyces marx-
ianus in biotechnology. In: Satyanarayana T, Kunze G (eds) Yeast diversity in human welfare.
Springer, p 439e53
Wang D, Wu D, Yang X, Hong J (2018) Transcriptomic analysis of thermotolerant yeast:
Kluyveromyces marxianus in multiple inhibitors tolerance. RSC Adv 8(26):14177–14192.
https://doi.org/10.1039/c8ra00335a
Wang R, Li L, Zhang B, Gao X, Wang D, Hong J (2013) Improved xylose fermentation of
Kluyveromyces marxianus at elevated temperature through construction of a xylose iso-
merase pathway. J Ind Microbiol Biotechnol 40(8):841–854. https://doi.org/10.1007/s10295-
013-1282-6
Wang R, Wang D, Gao X, Hong J (2014) Direct fermentation of raw starch using a Kluyveromyces
marxianus strain that express glucoamylase and alpha-amylase to produce ethanol. Biotechnol
Prog 30:338–347. https://doi.org/10.1002/btpr.1877
Wang YJ, Ying BB, Shen W, Zheng RC, Zheng YG (2017) Rational design of Kluyveromyces
marxianus ZJB14056 aldo–keto reductase KmAKR to enhance diastereoselectivity and activ-
ity. Enzyme Microb Technol 107:32–40. https://doi.org/10.1016/j.enzmictec.2017.07.012
Wolf M, Gasparin BC, Paulino AT (2018) Hydrolysis of lactose using β-d-galactosidase immobi-
lized in a modified Arabic gum-based hydrogel for the production of lactose-free/low-lactose
milk. Int J Biol Macromol 115:157–164. https://doi.org/10.1016/j.ijbiomac.2018.04.058
Wu WH, Cheng KC, Hung WC, Wan HP, Lo KY, Chen YH (2016) Bioethanol production from taro
waste using thermo-tolerant yeast Kluyveromyces marxianus K21. Biores Technol 201:27–32.
https://doi.org/10.1016/j.biortech.2015.11.015
Xu F, Wang KY, Wang N, Li G, Liu D (2017) Modified human glucagon-like peptide-1 (GLP-
1) produced in E. coli has a long-acting therapeutic effect in type 2 diabetic mice. PLoS One
12(7):e0181939. https://doi.org/10.1371/journal.pone.0181939
Yanase S, Yamada R, Ogino C, Kondo A, Hasunuma T, Tanaka T, Fukuda H (2010) Direct
ethanol production from cellulosic materials at high temperature using the thermotolerant
yeast Kluyveromyces marxianus displaying cellulolytic enzymes. Appl Microbiol Biotechnol
88(1):381–388. https://doi.org/10.1007/s00253-010-2784-z
Yang C, Hu S, Zhu S, Wang D, Gao X, Hong J (2015) Characterizing yeast promoters used in
Kluyveromyces marxianus. World J Microbiol Biotechnol 31:1641–1646. https://doi.org/10.
1007/s11274-015-1899-x
Yang D, Chen L, Duan J, Yu Y, Zhou J, Lu H (2021) Investigation of Kluyveromyces marxianus as
a novel host for large-scale production of porcine parvovirus-like particles. Microb Cell Fact
20:24. https://doi.org/10.1186/s12934-021-01514-5
Yang P, Zhu X, Zheng Z, Mu D, Jiang S, Luo S, Wu Y, Du M (2018) Cell regeneration and
cyclic catalysis of engineered Kluyveromyces marxianus of a D-psicose-3-epimerase gene from
Agrobacterium tumefaciens for D-allulose production. World J Microbiol Biotechnol 34:65.
https://doi.org/10.1007/s11274-018-2451-6
Yarimizu T, Nakamura M, Hoshida H, Akada R (2015) Synthetic signal sequences that enable effi-
cient secretory protein production in the yeast Kluyveromyces marxianus. Microb Cell Fact
14:20–34
Yuan WJ, Chang BL, Ren JG, Bai FW, Li YY, Liu JP (2012) Consolidated bioprocessing strat-
egy for ethanol production from Jerusalem artichoke tubers by Kluyveromyces marxianus under
high gravity conditions. J Appl Microbiol 112(1):38–44. https://doi.org/10.1111/j.1365-2672.
2011.05171.x
Yuangsaard N, Yongmanitchai W, Limtong S, Yamada M (2013) Selection and characterization
of a newly isolated thermotolerant Pichia kudriavzevii strain for ethanol production at high
Kluyveromyces marxianus as a Platform in Synthetic Biology … 335
temperature from cassava starch hydrolysate. Antonie Van Leeuwenhoek Int J General Mol
Microbiol 103(3):577–588. https://doi.org/10.1007/s10482-012-9842-8
Zafar S, Owais M (2006) Ethanol production from crude whey by Kluyveromyces marxianus.
Biochem Eng J 27(3):295–298
Zhang B, Li L, Zhang J, Gao X, Wang D, Hong J (2013) Improving ethanol and xylitol fer-
mentation at elevated temperature through substitution of xylose reductase in Kluyveromyces
marxianus. J Industr Microbiol Biotechnol 40(3–4):305–316. https://doi.org/10.1007/s10295-
013-1230-5
Zhang B, Zhang L, Wang D, Gao X, Hong J (2011) Identification of a xylose reductase gene
in the xylose metabolic pathway of Kluyveromyces marxianus NBRC1777. J Ind Microbiol
Biotechnol 38(12):2001–2010. https://doi.org/10.1007/s10295-011-0990-z
Zhang B, Zhu Y, Zhang J, Wang D, Sun L, Hong J (2017a) Engineered Kluyveromyces marxianus
for pyruvate production at elevated temperature with simultaneous consumption of xylose and
glucose. Biores Technol 224:553–562. https://doi.org/10.1016/j.biortech.2016.11.110
Zhang G, Lu M, Wang J, Wang D, Gao X, Hong J (2017b) Identification of hexose kinase genes in
Kluyveromyces marxianus and thermo-tolerant one step producing glucose-free fructose strain
construction. Scientific Reports 7. https://doi.org/10.1038/srep45104
Zhang J, Zhang B, Wang D, Gao X, Hong J (2014) Xylitol production at high temperature by engi-
neering Kluyveromyces marxianus. Bioresour Technol 152:192–201. https://doi.org/10.1016/j.
biortech.2013.10.109
Zhang J, Zhang B, Wang D, Gao X, Sun L, Hong J (2015a) Rapid ethanol production at ele-
vated temperatures by engineered thermotolerant Kluyveromyces marxianus via the NADP(H)-
preferring xylose reductase-xylitol dehydrogenase pathway. Metab Eng 31:140–152. https://
doi.org/10.1016/j.ymben.2015.07.008
Zhang M, Jiang L, Shi J (2015b) Modulation of mitochondrial membrane integrity and ROS forma-
tion by high temperature in Saccharomyces cerevisiae. Electron J Biotechnol 18(3):202–209.
https://doi.org/10.1016/j.ejbt.2015.03.008
Zhou HX, Xin FH, Chi Z, Liu GL, Chi ZM (2014) Inulinase production by the yeast
Kluyveromyces marxianus with the disrupted MIG1 gene and the over-expressed inulinase
gene. Process Biochem 49(11):1867–1874. https://doi.org/10.1016/j.procbio.2014.08.001
Zhou HX, Xu JL, Chi Z, Liu GL, Chi ZM (2013) β-Galactosidase over-production by a mig1
mutant of Kluyveromyces marxianus KM for efficient hydrolysis of lactose. Biochem Eng J
76:17–24. https://doi.org/10.1016/j.bej.2013.04.010
Zhou J, Zhu P, Hu X, Lu H, Yu Y (2018) Improved secretory expression of lignocellulolytic
enzymes in Kluyveromyces marxianus by promoter and signal sequence engineering. Biotech-
nol Biofuels 11(1). https://doi.org/10.1186/s13068-018-1232-7
Zoppellari F, Bardi L (2013) Production of bioethanol from effluents of the dairy industry
by Kluyveromyces marxianus. New Biotechnol 30(6):607–613. https://doi.org/10.1016/j.snbt.
2012.11.017
Synthetic Biology in the Candida
(CTG) Clade
Abstract
The continuous bio-race to find an ideal microbial chassis opens the path
toward the Candida CTG clade to enter the competition. Unexpectedly, this
unique CUG-serine coding clade successfully mastered the production of var-
ious nutraceutical, commercial, and pharmaceutical valuable compounds. The
following chapter aims to snapshot the Candida CTG clade’s bioengineering
properties and their biotechnological applications in synthetic biology.
1 Introduction
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2022 337
F. Darvishi Harzevili (ed.), Synthetic Biology of Yeasts,
https://doi.org/10.1007/978-3-030-89680-5_12
338 D. Kasir et al.
behaviors can generate synthetic bio-factories with new desirable properties (Lam
et al. 2010). The decrease in DNA synthesis and sequencing cost, the improvement
of the gene-to-protein relationship decoding, the standardization of DNA assem-
bly modalities, and the implementation of detector–reporter systems are considered
key players in this modern sector (Jensen and Keasling 2015). Various microbial
factories are considered promising candidates in several biotechnological appli-
cations. Among these, the mycotic CTG clade corresponds to a subdivision of
restricted ascomycetous yeast group that displays a distinctive genetic code (Papon
et al. 2014). In these species, a mistranslation reassigned the universal canon-
ical leucine CUG codon to serine predominantly (Papon et al. 2014). Despite
the presence of numerous Candida species that pose a significant health chal-
lenge (due to their ability to develop muco-cutaneous and/or systemic infections),
the CTG clade encompasses several species with strong biotechnological applica-
tion prepotency (Defosse et al. 2018a; Turner and Butler 2014). Currently, these
mycotic microorganisms are considered as suitable bio-producer candidates for
industrial enzymes, single-cell protein, bioethanol, vitamins, sweeteners, lipids,
and metabolites with pharmaceutical and nutritional values (Johnson 2013). Dis-
playing extraordinary potentials is one of the reasons that bring these yeasts into
the discipline of synthetic biology, namely tolerate niche osmolarity changes, grow
on a variety of carbon sources (including pentoses, n-alkanes, fatty acids, and
phenols), metabolize inexpensive substrate, and stand for extreme environmental
stresses (Papon et al. 2014; Johnson 2013). The genetic toolbox of CTG clade
yeasts is progressively prepared to generate a stand-alone bio-factory via recipient
strain construction, selectable marker adaptation, control (inducible or repress-
ible) gene expression system formation, and transformation protocol optimization
(Papon et al. 2012). Such efforts with the recent advancements in omics studies and
metabolic modeling improve the understanding of CTG clade metabolic behavior,
bringing it to the list of microbial competitors in the field of synthetic biology
(Papon et al. 2014). The following chapter will cover the taxonomy, the main
applied classical genetic tools, omics resources, and the major biotechnological
potentials in Candida CTG clade.
2 Taxonomy
Fig. 1 Phylogeny of Saccharomycotina and CTG clade yeasts. The topology represents the cur-
rent vision of the relations between clades and other species [according to (Dujon and Louis 2017;
Shen et al. 2016; Chen et al. 2000; Tsui et al. 2008)]. The human opportunistic species are shown
in red. The phylogeny of Saccharomycotina presented here in 8 clades is not definitive nor used
by all authors (Dujon and Louis 2017). Only a few species of the CTG clade are indicated in the
phylogenetic tree on the right. Adapted from Defosse et al. (2018a)
stipitis (Fig. 1) (Fitzpatrick et al. 2006; Butler et al. 2009; Nosek et al. 2009;
Defosse et al. 2018b). The unique genetic decoding ability of CTG clade was
characterized for the first time by Kawaguchi et al. (1989) in D. rugosa. Gener-
ally, the deviation of this clade’s coding system greatly supports the “ambiguous
intermediate theory,” which hypothesized that codons are reassigned through enig-
matic decoding in which a mutation in tRNA can expand its decoding array
capacity (Santos et al. 2004). Mühlhausen et al. (2014) described the cause for
the translation codon shift in CTG clade by demonstrating the polyphyletic rela-
tionship among the Candida genus using phylogenomic analyses of 26 motor
and cytoskeletal proteins. The evolutionary driving force mainly led to the emer-
gence of two groups with codon bias; one used the standard CUG translation,
while the other is an alternative yeast codon usage (AYCU) (Mühlhausen and
Kollmar 2014). Comparative and molecular phylogenomics studies showed that
tRNACAG Ser emerged (272 ± 25 million years ago) prior to the separation between
the Saccharomyces and Candida genera (170 ± 27 million years ago); such appear-
ance drives the common ancestor of CTG clade to lose the cognate tRNACAG leu
to the newly mutant tRNACAG Ser (Defosse et al. 2018b; Santos et al. 2011).
Comparative genomics analyses of CTG-Ser clade species revealed that CTG
codons are most frequently aligned with Ser rather than Leu amino acid (Riley
et al. 2016). Furthermore, their predicted tRNACAG Ser mainly has the three sery-
lation features: (i) a guanine at position 33 (G33) in the anticodon loop (may
lessen the rates of leucylation), (ii) Ser identity element (positioned in the vari-
able loop), and (iii) G discriminator base at position 73 (Riley et al. 2016). The
genomic comparison of ascomycete yeasts gives a snapshot of their biotechnolog-
ical exploitations, which reveals a heterogeneous distribution of metabolic traits,
with restriction to a single clade (such as methylotrophy), and patches distribu-
tion for others (such as D-xylose utilization) (Riley et al. 2016). Functionally
speaking, the incorporation of serine in polypeptides originating from CUG-loaded
340 D. Kasir et al.
mRNA drives the diversification in the primary structure of translated proteins that
possibly promotes the CTG clade yeasts to adapt to challenging environmental
conditions (Defosse et al. 2018b; Santos et al. 2011).
Reporter genes are considered a robust tool in genetic analysis, reporting success-
ful transformation and monitoring protein interaction and localization (Papon et al.
2012; Expression et al. 2002; Jones and Thomas 2003). Furthermore, apart from
reporting successful recipients of an episomal system, various reporter genes facil-
itate the dual-labeling approach in different Candida CTG clade yeasts which aids
in studying protein co-expression and co-localization (Courdavault et al. 2011; Rei-
jnst et al. 2011). Among the reporter genes used in Candida genetic manipulation
are:
(continued)
344
Table 1 (continued)
Candida CTG Selectable Reporter Genetic editing tool References
clade spp. marker genes Plasmid Recombinase Tet-system CRISPR-Cas
L. elongisporus IMH3.2 YFP pAYCU254 – – – Defosse et al. (2018a, b)
M. guilliermondii MET2 yEGFP3 pAYCU211 – – – Defosse et al. (2018a), Courdavault et al.
HYG# YFP pAYCU256 (2011), Millerioux et al. (2011), Obando
SAT1 CFP pAYCU281 Montoya et al. (2014), Foureau et al. (2013)
URA5 yEmRFP pAYCU230
pAYCU210
pAYCU212
M. farinosa ble ND pAYCU230 – – – Defosse et al. (2018a), Wang et al. (2006)
pAYCU210
pAYCU212
M. fructicola ND CFP pAYCU211 – – – Defosse et al. (2018a)
pAYCU256
pAYCU230
pAYCU210
pAYCU212
S. stipitis ble yEGFP3 pAYCU257 Cre – CRISPR/Cas9 Defosse et al. (2018a, b), Laplaza et al. (2006),
IMH3.2 pAYCU211 Cao et al. (2017b), Passoth et al. (2003)
pAYCU256
pAYCU230
pAYCU210
pAYCU212
S. passalidarum IMH3.2 GFP pAYCU270 – – – Defosse et al. (2018a, b)
pAYCU271
pAYCU229
Y. tenuis ND ND pCtAKR – – – Wohlbach et al. (2011)
D. Kasir et al.
Synthetic Biology in the Candida (CTG) Clade 345
Fig. 2 Schematic illustration of the standard pAYCU plasmid constructs (Defosse et al. 2018a)
2018a). Regarding reporter genes, the synthetic construct offers a broad spectrum
of ORF-adapted fluorescent protein alternatives including yeast-enhanced green
(yeGFP), yellow (yeYFP), cyan (yeCFP), monomeric cherry (yemCherry), the
surface-exposed Gaussia princeps luciferase (gLUC59), and lacZ gene (StlacZ)
(Courdavault et al. 2011; Enjalbert et al. 2009; Uhl and Johnson 2001) (Fig. 2).
Various combinations of pAYCU episome construct are offered to adapt the genetic
transformation and manipulation in Candida spp. (Defosse et al. 2018a) (Table
1). Nowadays, the plasmid transformation is no longer restricted to the conven-
tional ectopic gene manipulation but rather is extended to markerless gene editing
modalities (Lombardi et al. 2019; Wang et al. 2018). Wang et al. (2018) recently
reported the construction of a “suicide plasmid” named pPICPJ-mazF, facilitating
scarless metabolic manipulation in C. tropicalis to increase the conversion rate of
oils into long-chain dicarboxylic acids (DCAs). The future perspective for Candida
gene manipulation will be mainly shifting from standardized conventional plasmid
transformation to genetic circuit design with directed functional behavior (Papon
et al. 2014; Røkke et al. 2014).
3.4 Recombinases
2012). The intensive efforts over the last two decades facilitated the application of
two main recombinase systems in the CTG clade, the flippase (FLP) recognition
target (FRT) and the Cre recombinase (Cre) locus of crossing over (x), P1 (loxP)
site-specific recombinases (Papon et al. 2012; Nunes-Düby et al. 1998; Austin
et al. 1981; Sternberg and Hamilton 1981).
genome editing (Heidari et al. 2017; Stoneman et al. 2020). Typically, the selec-
tion of a single-guide RNA (sgRNA) target site is considered as the prime step
in designing each CRISPR–Cas9 experiment (Stoneman et al. 2020). Upon the
assembly of the sgRNA and Cas9 (endonuclease) enzyme complex, the guide RNA
directs Cas9 for manipulating its complementary region in the genome (Stoneman
et al. 2020).
Experimentally speaking, this novel tool demonstrates its ability to meticulously
reduce the transformation steps and minimize the off-target effects and the overall
number of “unwanted” genomic changes, especially hyperploidy events (Marton
et al. 2020; Chandrasegaran and Carroll 2016). Cao et al. (2017b) demonstrated
the potency of using CRISPR technology in non-conventional yeasts, namely
S. stipitis to overcome the genome editing’s limited efficacy, where the rate of
ade2-knockout and trp1-knockout was boosted up from <1% to 80%. In addi-
tion, both centromere (CEN) and autonomously replicating sequence (ARS) of the
350 D. Kasir et al.
CRISPR–Cas9 episome system were able to trick the yeast cell and solve both,
segregation bias and directing plasmid replication (Cao et al. 2017b). The overall
titer of the target component was improved 3-folds through using the above-
mentioned stable minichromosome-like expression construct (Cao et al. 2017b).
Recently, a new species, “the Cinderella of the non-conventional yeast,” namely
C. famata, joined the list of CRISPR–Cas9-manipulated microorganisms (Prista
et al. 2016; Spasskaya et al. 2021). Spasskaya et al. (2021) demonstrated the
integral role of 26S proteasome in the extremophilic nature of this interesting
species (halo- and osmotolerant capacity) by developing scarless highly efficient
single (pDhCRISPR-1)- and dual-guide (pDhCRISPR-2) CRISPR systems. The
targeted mutagenesis of DhRpn4-binding sites (Rpn4: induced transcription factor,
regulates proteasomal gene expression under stress conditions) leads to C. famata
proteasomal deregulation, which generated strains sensitive to various environ-
mental stresses (such as gene-proteotoxic, oxidative stress, and conditions of high
salinity and osmolarity) (Spasskaya et al. 2021). Moreover, this cutting-edge tool
allowed mastering the generation of diploid mutant strains such as C. albicans,
rather than only disrupting a single allele as a result of conventional mutagenesis
strategies that limit the efficacy of constructing null mutations (both alleles are
modified at once) (Vyas et al. 2015). However, the application of CRISPR/Cas9-
based gene editing in CTG clade is somehow cumbersome because of the
prerequisite of Cas9 gene recoding to overcome the “leucine/serine translation
exchange” issue; thus, codon optimization is crucially needed in CRISPR-CTG
clade applications (Vyas et al. 2015). Different studies have focused on the Can-
dida codon optimization; for instance, Vyas et al. (2015) constructed two systems:
solo and duet, expressing Candida/Saccharomyces codon-optimized version of
Cas9 (CaCas9) that evades the usage of CUG codon with compatibility to all
CTG species. Ironically, Grahl et al. (Grahl et al. 2017) described an alternative
approach to circumvent the need for species-specific manipulation constructs by
applying expression-free CRISPR genome editing in C. lusitaniae, C. glabrata, and
C. auris. The alternative construct is composed of purified Cas9 protein and gene-
specific and scaffold RNAs referred as RNA–protein complexes (RNPs) (Grahl
et al. 2017).
The transient system is typically used as a safeguard to avoid the nega-
tive feedback (cell toxicity and strain fitness drop) of CRISPR/Cas9 constitutive
expression (Min et al. 2016). In 2017, Norton et al. (2017) developed the first
transient CRISPR–Cas9 expression with gene deletion efficiencies up to 81%
in Ku70 and LIG4 (non-homologous end joining (NHEJ) factors) in C. lusita-
niae-deficient strains. Huang and Mitchell (2017) generated a marker recycling
model called CRISPR–Cas9-induced marker excision (CRIME) in C. albicans
to avoid the paucity of drug-resistant makers. Methodologically speaking, the
selection marker that is used to delete the target gene is flanked by directly
repeated sequences. The CRISPR–Cas9 system is used dually to delete the gene
of interest subsequently to the deletion of the marker itself by double-strand break
(DSB) induction, leading to marker excision by recombination between the direct
Synthetic Biology in the Candida (CTG) Clade 351
repeats (Huang and Mitchell 2017). Furthermore, Nguyen et al. (2017) devel-
oped two recyclable CRISPR-mediated approaches, namely the LEUpOUT system
(applicable to LEU2/leu2 parental strain) and the HIS–FLP system (applicable
theoretically to any nourseothricin-sensitive C. albicans strain), to allow scarless
genome manipulation of C. albicans without the integration of permanent markers
or the application of a cloning step. Furthermore, Lombardi et al. (2017) applied
an episome-based CRISPR–Cas9 system in C. parapsilosis; the adopted plasmid
system called pRIBO theoretically offered the ability to manipulate any genetic
background with the possibility of sequential genome editing in the same back-
ground since the plasmid is escaped easily in the absence of selection. Afterward,
the former system was further improved into markerless-gene manipulation sys-
tem adapting various Candida spp., namely pCP-tRNA and pCT-tRNA applied
on C. parapsilosis (and their sister species, Candida orthopsilosis and Candida
metapsilosis) and C. tropicalis, respectively (Lombardi et al. 2019).
Interestingly, since bioinformatics is the keystone in scientific speaking lan-
guages, a breakthrough in CRISPR–Cas9 application area was made by Stoneman
et al. (2020) who developed recently CRISpy-pop, a flexible Python-based Web
application with a user-friendly graphical interface. The substantial aim of this
application is to theoretically design CRISPR/Cas9-driven genetic manipulations
on individual strain or population of strains to predict the most efficient combina-
tion (sgRNA and strain) (Stoneman et al. 2020). Currently, this powerful tool can
cover both bacteria (Zymomonas mobilis) and yeasts (S. cerevisiae), with a future
perspective to integrate new biological models (Stoneman et al. 2020).
4 Omics Resources
The complexity of the living cell hindered the scientists’ ability to see the
whole picture behind any given phenotype for a long period of time, since they
mainly adopted the “one–one analysis” (one gene/protein at a time) in different
biomedical researches, neglecting the myriad interactions between genes, pro-
teins, carbohydrates, lipids and metabolites that are eventually translated into a
certain phenotype (Stagljar 2016). The innovations of technological techniques
have significantly facilitated the conjugation of the “omics” suffix to both, the
genomic and post-genomic levels (transcriptomics, proteomics, and metabolomics)
that decode holistically the physiological and biochemical framework of numer-
ous organisms (Hasin et al. 2017; Babar et al. 2018; Narad et al. 2018; Patra
et al. 2021). Since omics experiments (high-dimensional biology) are hypothesis-
generating rather than hypothesis-driven studies (traditional experiment approach),
it could aid in opening the path toward building a nonbiased picture of the biolog-
ical system (Horgan and Kenny 2011). Generally, systemic biology represents an
interdisciplinary platform for mathematical analysis and computational modeling
approaches that translate the complexity of a defined biological system via either
top-down or bottom-up modalities (Patra et al. 2021).
352 D. Kasir et al.
4.1 Genomics
4.2 Transcriptomics
4.3 Proteomics
This modality represents the third stage in the holistic understanding of a cell’s
biological processes. It offers in-depth characterization of microorganisms’ pro-
tein profiles (including their location, structure, and function) since the transcript
and protein levels are weakly related due to the specific regulatory mechanisms
Synthetic Biology in the Candida (CTG) Clade 353
4.4 Metabolomics
These approaches constitute the fourth stage of the “biological puzzle-solving” that
provide a snapshot of metabolite concentrations (Pickford et al. 2019). Generally,
the metabolomics approach aims to study quantitatively the microorganism’s cel-
lular metabolite profile under specific culture conditions (Roldão et al. 2012). The
construction of mathematical models is considered one of the efficient tools in plot-
ting and understanding the metabolic profile in a defined organism by integrating
data science platforms such KBase, which involves semi-automated reconstruction
approaches (such as RAVEN, Merlin, and ModelSEED) (Arkin et al. 2018). It
displays a detailed map for the underlying molecular interaction that can evalu-
ate the efficacy of future network response under specific genetic/environmental
interventions (Roldão et al. 2012). A corrected meticulously comprehensive view
of microbe’s metabolic trait could be achieved by adding species-specific software
databases such as BRENDA, KEGG, and MetaCyc (Patra et al. 2021).
4.5 Fluxomics
Since metabolomics neglects the control and the functional regulation of metabolic
networks, the fluxomics platform acts as a complementary approach to intensively
understand the microorganism’s phenotype by determining the rates of metabolic
reactions (Pickford et al. 2019; Ratcliffe and Shachar-Hill 2005). Thereby improv-
ing our ability to predict the effect of genetic manipulation on the entity of the
microorganism’s physiological behavior, which leads to the formation of micro-
bial systems with improved metabolic capacity (Yu et al. 2012). The offered flux
balance analysis (FBA) in the fluxomics discipline greatly facilitates the reading
of metabolite flow through a defined metabolic model (Orth et al. 2010; Price
et al. 2004). The FBA analysis involves several advantages, such as the analysis
of phenotypic behaviors and flux coupling, the evolution of metabolic status with
environmental and genetic changes, and the prediction of metabolic engineering
foci (Patra et al. 2021).
The high-throughput potency of omics to comprehensively study the microbial
system is directly dependent on the available pool of technologies to scientifically
decode such huge data. The genomics platform constituted various technologies to
systematically analyze the Candida genome such as Candida Gene Order Browser
(CGOB), multiplex PCR, whole-genome sequencing, flow cytometry, slot blot
hybridization, microarray, Avadis, FuncExpression, High-Throughput GoMiner,
and GREAT (Roldão et al. 2012; Sampaio et al. 2005; Asadzadeh et al. 2020;
354 D. Kasir et al.
Maldonado et al. 2018; Chang et al. 2015; Loeffler et al. 2000; Campa et al. 2008;
Maguire et al. 2013).
Regarding transcriptomics, numerous techniques and sites are established to
facilitate integrative analysis of Candida transcripts, e.g., dual RNA sequenc-
ing, CLASSIFY, FunCluster, GARBAN, GOdist, dChip, ErmineJ, GoMiner,
Matrix2png, TM4, MAGIC, GeneSpring, Array-Pro Analyzer, and ArrayStar
(Roldão et al. 2012; Schulze et al. 2016).
For the Candida proteomics and metabolomics profiling, different method-
ological approaches are developed to perform a holistic analysis of yeast
proteomic and metabolomic repertoire, e.g., sodium dodecyl sulfate polyacry-
lamide gel electrophoresis (SDS-PAGE), two-dimensional gel electrophoresis
(2DE), differential in-gel electrophoresis (DIGE), electrospray ionization-mass
spectrometry (ESI–MS), matrix-assisted laser desorption/ionization-MS (MALDI-
MS), MALDI-time-of-flight (MALDI-TOF), liquid chromatography-MS (LC–
MS), high-performance liquid chromatography (HPLC), gas chromatography-MS
(GC–MS), ultra-performance liquid chromatography-MS (UPLC-MS), AMDIS
software, Pathway Activity Profiling (PAPi) algorithm, UPLC-QTOF-MS Perseus
software, SIMCA, and FiatFlux (Wang et al. 2009, 2020; Larbi and Jefferies 2009;
Gräslund et al. 2008; Kim et al. 2005; Awad et al. 2018; Karkowska-Kuleta et al.
2020; Cabrera et al. 2010; Christen and Sauer 2011; Han et al. 2012).
The fluxomics yeast profiling is still restricted due to the disciplines nascent
state, leaving behind some constructed tools to examine the microbial pheno-
typic expression such as 13 C metabolic flux analysis (13 C MFA), OptKnock,
OptReg, OptStrain, optimal metabolic network identification (OMNI), OptGene,
thermodynamics-based MFA, COBRA toolbox, and OpenFLUX (Pharkya and
Maranas 2006; Henry et al. 2007; Feng et al. 2010; Pharkya et al. 2004; Her-
rgård et al. 2006; Blazeck and Alper 2010; Schellenberger et al. 2011; Veras et al.
2019).
The microbial community takes advantage of “the power of the group” which
offers multiple boosters for a higher production rate such as having a robust bio-
community, tolerating various environmental challenges, reducing the “metabolic
load,” allowing the exchange of resources, activating bio-communication (between
species), and expanding the metabolic fitness (Stenuit and Agathos 2015; Tsoi
et al. 2018; Bassler and Losick 2006). Interestingly, C. famata was recently
reported as a potential candidate to enter the microbial consortia construction for
the industrial production of riboflavin (Bhatia et al. 2018). The application of such
bio-consortium is expected to increase in the near future especially with the suc-
cessful integration of CRISPR/Cas9 system in the consortium building process
(Wang et al. 2016a). The following section will cover the current main metabolic
engineering applications and biotechnological potentials of Candida CTG clade
yeasts (Table 2).
The world’s modern and technological lifestyle posed many challenges in sus-
taining energy sources especially with the existence of ever-increasing concerns
about economic and environmental consequences (Pereira et al. 2015). This stim-
ulates collaborative research efforts aiming to produce renewable, eco-friendly
biofuel sources using microbial factories (Du et al. 2019). Lignocellulose’s (con-
sists of cellulose, hemicellulose, and lignin) bioconversion into ethanol is currently
being considered as an alternative source to petroleum-based fuels (Rodrussamee
et al. 2018). The efficient fermentation of lignocellulose sugars represents a lim-
iting factor in the production process, especially xylose sugar (most abundant in
hemicellulose) (Rodrussamee et al. 2018). However, since CTG clade species are
considered as xylose growers (utilized as their sole carbon source) with some hav-
ing the ability to ferment xylose naturally into ethanol, they represent an attractive
superior microbial platform in such an industrial section (Papon et al. 2014). The
comparative genomic studies for xylose fermenters revealed the presence of ampli-
fication in sugar transporters and cell surface proteins, which may explain their
exceptional sugar environment (Wohlbach et al. 2011). Interestingly, Wohlbach
et al. reported an increase in the flux of the xylose assimilation pathway with
the absence of xylitol accumulation in strains engineered with pCtAKR (Candida
tenuis aldo/keto reductase), suggesting their involvement in stimulating NADH
recycling and glycerol production (Wohlbach et al. 2011). The insertion of opti-
mized (to frequently used codons in S. cerevisiae) exo-inulinase gene INU1 from
M. guilliermondii (INU1Y ) in S. cerevisiae resulted in Y13 recombinant strains
with inulinase activity up to 43.84 U/mL and 126.30 mg/mL ethanol production
from 300.0 g/L inulin (Liu et al. 2014). In addition, the fusion of S. stipitis with
S. cerevisiae (the microbe of choice in the ethanol industry) protoplasts gener-
ates a hybrid (unstable) strain with enhanced xylose-based ethanol production in
comparison with S. stipitis wild strain (Ruchala et al. 2020; Yoon et al. 1996). Fur-
thermore, the generation of S. stipitis Δhxk1 and 2-deoxyglucose-resistant mutants
has resulted in the derepression of xylose utilization in both engineered strains
356
The microbial lipid production has revolutionized both the industrial sector that
relies intensively on non-renewable, environmentally hazardous resources of fos-
sil fuel and the bio-medicinal area that recently integrated the lipo-molecules in
numerous scientific applications (Li et al. 2020; Younes et al. 2020). Naturally
speaking, the lipid accumulation in oleaginous microorganisms can take two path-
ways, the “de novo” (fermentation on hydrophilic substrates such as sugars and
related substrates) and the “ex novo” lipid synthesis (fermentation on hydrophobic
substrates such as oils and alkane; greatly depends on the carbon source) (Yan et al.
2020; Huang et al. 2017). For example, sophorolipids, and low-toxic, biodegrad-
able, and ecofriendly biosurfactants have gained interest in various therapeutic
Synthetic Biology in the Candida (CTG) Clade 359
KU141F1 showed higher DCA degradation (relative to the parental strain), sug-
gesting the coexistence of non-peroxisomal fatty acid catabolism pathway (Werner
et al. 2017).
The biotechnological potentials of Candida CTG clade are not constrained only to
the former cited applications; rather, numerous species show additional capabilities
in producing valuable bio-molecules such as industrial enzymes and therapeutic
compounds (Papon et al. 2013; Roa Engel et al. 2008; Križanović et al. 2015).
Various Candida CTG clade yeasts demonstrate a remarkable capacity in pro-
ducing citric acid, an organic acid commonly used in food and pharmaceutical
industries (Singh Dhillon et al. 2011). Anastassiadis et al. (2005) reported a suc-
cessful continuous citric acid fermentation by Candida oleophila ATCC 20177
strain under nitrogen limitation condition with synthesis rate up to 74.2 g/L. Fur-
thermore, Kim et al. described considerable citric acid productivity by submerged
fermentation of C. oleophila ATCC 20177, C. tropicalis ATCC 20115, and C.
oleophila ATCC 20373 strains up to 20.7 g/L, 45.0 g/L, and 60.1 g/L, respectively
(Kim et al. 2015).
M. guilliermondii appears to be one of the most efficient candidates in the
xylose-xylitol conversion process with considerable variation in xylitol yield based
on the accredited strain (Papon et al. 2013). The generated xylitol is beneficial in
various nutraceuticals (chewing gum, candies, wafer fillings, and chocolate) and
pharmaceuticals (protein extraction stabilizer and antineoplastic properties) indus-
tries (Papon et al. 2013; Ur-Rehman et al. 2015). Cortez et al. (2016) reported
an effective xylitol production from D-xylose using M. guilliermondii (permeabi-
lized with Triton X-100) with a yield up to 0.80 g/g and 2.65 g/L volumetric
productivity. Furthermore, M. guilliermondii showed high potentials in producing
2-phenylethanol (2-PE), aromatic compound (rose scent) used in cosmetics, per-
fume, and food industries (Yan et al. 2020). Yan et al. (2021) recently reported
newly identified M. guilliermondii strain YLG18 as a promising candidate for 2-
PE high production from L-phe. Upon optimization process, this strain was able
to synthesize 2-PE up to 3.20 g/L (Yan et al. 2021).
S. stipitis shows promising potentials in producing various valuable compounds
among which is fumaric acid, a C4-dicarboxylic acid applied in nutraceutical,
pharmaceutical, and chemical industries (Roa Engel et al. 2008). Wei et al. (2015)
described the development of an optimized strain PSYPMFfS with the following
properties, overexpression of heterologous reductive fumaric acid synthetic path-
way (Rhizopus oryzae FM19), codon modification, blockage of the conversion of
fumaric acid (double-deletion of fumarase genes (Psfum1 and Psfum2), and over-
expression of fumaric acid heterologous transporter (YMAE1). The obtained strain
produced fumaric acid titer up to 4.67 g/L from xylose (Wei et al. 2015). Addi-
tionally, Ilmén et al. (2007) described an efficient production of l-lactic acid by
an engineered S. stipitis strain; the random integration of heterologous LDH gene
enhanced the yield of lactate up to 58 g/L from 100 g/L xylose. The produced lactic
acid can be applied in various industrial applications such as biodegradable plas-
tics and textile fiber manufacturers (Ilmén et al. 2007). The systematic metabolic
engineering driven by flux facilitates the design of S. stipitis-mutant strains with
improved biomass yield up to 44% (in ZWF1 overexpressed mutant) (Unrean et al.
362 D. Kasir et al.
5.5 Bioremediation
parapsilosis ranged from 50 to 64% in removing aflatoxin B1, and the combina-
tion of the aforementioned species with C. tropicalis displayed a removal capacity
ranging from 56 to 68% for zinc metal.
5.6 Biocontrol
The importance of CTG clade yeasts in the field of synthetic biology is not
doubtable anymore; all the identified capabilities in this unique clade are just pre-
senting the tip of submerged potentials. The future vision of Candida metabolic
engineering will mainly deviate to the Candida consortium rather than the sin-
gle strain notion with the application of more species-directed (species–hybrid)
and synthetic standardized gene rewiring modules. The unlimited efforts to stan-
dardize the advanced gene-editing tools are the keystone in deciphering the
competent species. Thus, the continuous optimization and integration of both,
gene-engineering strategies and omics platforms, will facilitate building special-
ized potentially unexplored CTG clade yeast strains with desirable properties in
additional biotechnological fields.
References
Abbas CA, Sibirny AA (2011) Genetic control of biosynthesis and transport of riboflavin and flavin
nucleotides and construction of robust biotechnological producers. Microbiol Mol Biol Rev
75(2):321–360
Agirman B, Erten H (2020) Biocontrol ability and action mechanisms of Aureobasidium pullu-
lans GE17 and Meyerozyma guilliermondii KL3 against Penicillium digitatum DSM2750 and
Penicillium expansum DSM62841 causing postharvest diseases. Yeast 37(9–10):437–448
Anastassiadis S, Wandrey C, Rehm HJ (2005) Continuous citric acid fermentation by Can-
dida oleophila under nitrogen limitation at constant C/N ratio. World J Microbiol Biotechnol
21(5):695–705
Andreieva Y, Petrovska Y, Lyzak O, Liu W, Kang Y, Dmytruk K et al (2020) Role of the regulatory
genes SEF1, VMA1 and SFU1 in riboflavin synthesis in the flavinogenic yeast Candida famata
(Candida flareri). Yeast (chichester, England) 37(9–10):497–504
Arkin AP, Cottingham RW, Henry CS, Harris NL, Stevens RL, Maslov S et al (2018) KBase:
the United States department of energy systems biology knowledgebase. Nat Biotechnol
36(7):566–569
Asadzadeh M, Dashti M, Ahmad S (2020) Whole genome and targeted-amplicon sequencing of
fluconazole-susceptible and -resistant Candida parapsilosis isolates from Kuwait reveals a
previously undescribed N1132D polymorphism in CDR1. Antimicrob Agents Chemother
Austin S, Ziese M, Sternberg N (1981) A novel role for site-specific recombination in maintenance
of bacterial replicons. Cell 25(3):729–736
Awad A, El Khoury P, Wex B, Khalaf RA (2018) Proteomic analysis of a Candida albicans pga1
Null Strain. EuPA Open Proteom 18:1–6
Babar MM, Afzaal H, Pothineni VR, Zaidi N-u-SS, Ali Z, Zahid MA et al (2018) Omics
approaches in industrial biotechnology and bioprocess engineering ((Chap 14)). In: Barh D,
Azevedo V (eds) Omics technologies and bio-engineering. Academic Press, pp 251–69
Babbal, Adivitiya, Khasa YP (2017) Microbes as biocontrol Agents. In: Kumar V, Kumar M,
Sharma S, Prasad R (eds) Probiotics and plant health. Springer, Singapore, pp 507–552
Bahafid W, Tahri Joutey N, Sayel H, Boularab I, Ghachtouli N (2013) Bioaugmentation
of chromium-polluted soil microcosms with Candida tropicalis diminishes phytoavailable
chromium. J Appl Microbiol 115(3):727–734
Bassler BL, Losick R (2006) Bacterially speaking. Cell 125(2):237–246
Basso LR Jr, Bartiss A, Mao Y, Gast CE, Coelho PSR, Snyder M et al (2010) Transformation of
Candida albicans with a synthetic hygromycin B resistance gene. Yeast (chichester, England)
27(12):1039–1048
Synthetic Biology in the Candida (CTG) Clade 365
Cormack BP, Bertram G, Egerton M, Gow NA, Falkow S, Brown AJ (1997) Yeast-enhanced green
fluorescent protein (yEGFP): a reporter of gene expression in Candida albicans. Microbiology
143(2):303–311
Cortez DV, Mussatto SI, Roberto IC (2016) Improvement on D-xylose to xylitol biotransformation
by Candida guilliermondii using cells permeabilized with triton X-100 and selected process
conditions. Appl Biochem Biotechnol 180(5):969–979
Courdavault V, Millerioux Y, Clastre M, Simkin AJ, Marais E, Crèche J et al (2011) Fluorescent
protein fusions in Candida guilliermondii. Fungal Genet Biol 48(11):1004–1011
Dangi AK, Sharma B, Hill RT, Shukla P (2019) Bioremediation through microbes: systems biology
and metabolic engineering approach. Crit Rev Biotechnol 39(1):79–98
Dashtban M, Wen X, Bajwa PK, Ho CY, Lee H (2015) Deletion of hxk1 gene results in derepres-
sion of xylose utilization in Scheffersomyces stipitis. J Ind Microbiol Biotechnol 42(6):889–896
Defosse TA, Courdavault V, Coste AT, Clastre M, de Bernonville TD, Godon C et al (2018a)
A standardized toolkit for genetic engineering of CTG clade yeasts. J Microbiol Methods
144:152–156
Defosse TA, Le Govic Y, Courdavault V, Clastre M, Vandeputte P, Chabasse D et al (2018b) Les
levures du clade CTG (clade Candida): biologie, incidence en santé humaine et applications en
biotechnologie. Med Mycol J 28(2):257–268
Defosse TA, Mélin C, Clastre M, Besseau S, Lanoue A, Glévarec G et al (2016) An additional
Meyerozyma guilliermondii IMH3 gene confers mycophenolic acid resistance in fungal CTG
clade species. FEMS Yeast Res 16(6)
Dennison PM, Ramsdale M, Manson CL, Brown AJ (2005) Gene disruption in Candida albicans
using a synthetic, codon-optimised Cre-loxP system. Fungal Genet Biol 42(9):737–748
DiCarlo JE, Norville JE, Mali P, Rios X, Aach J, Church GM (2013) Genome engineering in
Saccharomyces cerevisiae using CRISPR-Cas systems. Nucleic Acids Res 41(7):4336–4343
Ding C, Butler G (2007) Development of a gene knockout system in Candida parapsilosis reveals
a conserved role for BCR1 in biofilm formation. Eukaryot Cell 6(8):1310–1319
Dmytruk KV, Yatsyshyn VY, Sybirna NO, Fedorovych DV, Sibirny AA (2011) Metabolic engi-
neering and classic selection of the yeast Candida famata (Candida flareri) for construction of
strains with enhanced riboflavin production. Metab Eng 13(1):82–88
Dmytruk K, Lyzak O, Yatsyshyn V, Kluz M, Sibirny V, Puchalski C et al (2014) Construction
and fed-batch cultivation of Candida famata with enhanced riboflavin production. J Biotechnol
172:11–17
Doyle TC, Nawotka KA, Purchio AF, Akin AR, Francis KP, Contag PR (2006a) Expression of
firefly luciferase in Candida albicans and its use in the selection of stable transformants. Microb
Pathog 40(2):69–81
Doyle TC, Nawotka KA, Kawahara CB, Francis KP, Contag PR (2006b) Visualizing fungal infec-
tions in living mice using bioluminescent pathogenic Candida albicans strains transformed
with the firefly luciferase gene. Microb Pathog 40(2):82–90
Du C, Li Y, Zhao X, Pei X, Yuan W, Bai F et al (2019) The production of ethanol from ligno-
cellulosic biomass by Kluyveromyces marxianus CICC 1727–5 and Spathaspora passalidarum
ATCC MYA-4345. Appl Microbiol Biotechnol 103(6):2845–2855
Dujon BA, Louis EJ (2017) Genome diversity and evolution in the budding yeasts (Saccharomy-
cotina). Genetics 206(2):717–750
Enjalbert B, Rachini A, Vediyappan G, Pietrella D, Spaccapelo R, Vecchiarelli A et al (2009) A
multifunctional, synthetic Gaussia princeps luciferase reporter for live imaging of Candida
albicans infections. Infect Immun 77(11):4847–4858
Expression G, Sundaresan G, Gambhir SS (2002) Radionuclide imaging of reporter gene expres-
sion (Chap 29). In: Toga AW, Mazziotta JC (eds) Brain mapping: the methods, 2nd edn.
Academic Press, San Diego, pp 799–818
Feng X, Page L, Rubens J, Chircus L, Colletti P, Pakrasi H et al (2010) Bridging the gap between
fluxomics and industrial biotechnology. J Biomed Biotechnol 2010:460717
Fitzpatrick DA, Logue ME, Stajich JE, Butler G (2006) A fungal phylogeny based on 42 complete
genomes derived from supertree and combined gene analysis. BMC Evol Biol 6(1):99
Synthetic Biology in the Candida (CTG) Clade 367
Foureau E, Courdavault V, Navarro Gallón SM, Besseau S, Simkin AJ, Crèche J et al (2013) Char-
acterization of an autonomously replicating sequence in Candida guilliermondii. Microbiol Res
168(9):580–588
Freimoser FM, Rueda-Mejia MP, Tilocca B, Migheli Q (2019) Biocontrol yeasts: mechanisms and
applications. World J Microbiol Biotechnol 35(10):154
Gabriel F, Accoceberry I, Bessoule JJ, Salin B, Lucas-Guérin M, Manon S et al (2014) A Fox2-
dependent fatty acid ß-oxidation pathway coexists both in peroxisomes and mitochondria of the
ascomycete yeast Candida lusitaniae. PLoS One 9(12):e114531
Gácser A, Trofa D, Schäfer W, Nosanchuk JD (2007) Targeted gene deletion in Candida parap-
silosis demonstrates the role of secreted lipase in virulence. J Clin Invest 117(10):3049–3058
Gao M, Cao M, Suástegui M, Walker J, Rodriguez Quiroz N, Wu Y et al (2017) Innovating a
nonconventional yeast platform for producing shikimate as the building block of high-value
aromatics. ACS Synth Biol 6(1):29–38
García-Béjar B, Arévalo-Villena M, Guisantes-Batan E, Rodríguez-Flores J, Briones A (2020)
Study of the bioremediatory capacity of wild yeasts. Sci Rep 10(1):11265
Gerami-Nejad M, Berman J, Gale CA (2001) Cassettes for PCR-mediated construction of green,
yellow, and cyan fluorescent protein fusions in Candida albicans. Yeast 18(9):859–864
Gerami-Nejad M, Dulmage K, Berman J (2009) Additional cassettes for epitope and fluorescent
fusion proteins in Candida albicans. Yeast 26(7):399–406
Gordon ZB, Soltysiak MPM, Leichthammer C, Therrien JA, Meaney RS, Lauzon C et al (2019)
Development of a transformation method for Metschnikowia borealis and other CUG-Serine
Yeasts. Genes (Basel) 10(2)
Grahl N, Demers EG, Crocker AW, Hogan DA (2017) Use of RNA-protein complexes for genome
editing in non-albicans Candida species. mSphere 2(3)
Gräslund S, Sagemark J, Berglund H, Dahlgren LG, Flores A, Hammarström M et al (2008) The
use of systematic N- and C-terminal deletions to promote production and structural studies of
recombinant proteins. Protein Expr Purif 58(2):210–221
Griffith F (1928) The significance of pneumococcal types. J Hyg 27(2):113–159
Guerrero V, Guigón-López C, Berlanga D, Ojeda-Barrios D (2014) Complete control of Penicil-
lium expansum on apple fruit using a combination of antagonistic yeast Candida oleophila.
Chilean J Agric Res 74:427–431
Han T-l, Cannon RD, Villas-Bôas SG (2012) Metabolome analysis during the morphological
transition of Candida albicans. Metabolomics 8(6):1204–1217
Hara A, Arie M, Kanai T, Matsui T, Matsuda H, Furuhashi K et al (2001) Novel and convenient
methods for Candida tropicalis gene disruption using a mutated hygromycin B resistance gene.
Arch Microbiol 176(5):364–369
Hasin Y, Seldin M, Lusis A (2017) Multi-omics approaches to disease. Genome Biol 18(1):83
Heidari R, Shaw DM, Elger BS (2017) CRISPR and the rebirth of synthetic biology. Sci Eng Ethics
23(2):351–363
Henry CS, Broadbelt LJ, Hatzimanikatis V (2007) Thermodynamics-based metabolic flux analysis.
Biophys J 92(5):1792–1805
Herrgård M, Panagiotou G (2012) Analyzing the genomic variation of microbial cell factories in
the era of “New Biotechnology”. Comput Struct Biotechnol J 3(4):e201210012
Herrgård M, Fong S, Palsson B (2006) Identification of genome-scale metabolic network models
using experimentally measured flux profiles. PLoS Comput Biol 2:e72
Hinchliffe E, Kenny E (1993) Yeast as a vehicle for the expression of heterologous genes (Chap
9). In: Rose AH, Stuart Harrison J (eds) The yeasts, 2nd edn. Academic Press, San Diego, pp
325–356
Hirata Y, Ryu M, Oda Y, Igarashi K, Nagatsuka A, Furuta T et al (2009) Novel characteristics
of sophorolipids, yeast glycolipid biosurfactants, as biodegradable low-foaming surfactants. J
Biosci Bioeng 108(2):142–146
Holkers M, Vries AAFd, Gonçalves MAFV (2006) Modular and excisable molecular switch for
the induction of gene expression by the yeast FLP recombinase. BioTech 41(6):711–713
368 D. Kasir et al.
Horgan RP, Kenny LC (2011) ‘Omic’ technologies: genomics, transcriptomics, proteomics and
metabolomics. Obstet Gynaecol 13(3):189–195
Huang C, Luo M-T, Chen X-F, Qi G-X, Xiong L, Lin X-Q et al (2017) Combined “de novo” and
“ex novo” lipid fermentation in a mix-medium of corncob acid hydrolysate and soybean oil by
Trichosporon dermatis. Biotechnol Biofuels 10(1):147
Huang MY, Mitchell AP (2017) Marker recycling in Candida albicans through CRISPR-Cas9-
induced marker excision. mSphere 2(2):e00050-17
Ilmén M, Koivuranta K, Ruohonen L, Suominen P, Penttilä M (2007) Efficient production of lactic
acid from Xylose by Pichia stipitis. Appl Environ Microbiol 73(1):117–123
Ishchuk O, Dmytruk K, Rohulya O, Voronovsky A, Abbas C, Sibirny A (2008) Development of a
promoter assay system for the flavinogenic yeast Candida famata based on the Kluyveromyces
lactis β-galactosidase LAC4 reporter gene. Enzyme Microb Technol 42:208–215
Jenior ML, Moutinho TJ Jr, Dougherty BV, Papin JA (2020) Transcriptome-guided parsimonious
flux analysis improves predictions with metabolic networks in complex environments. PLoS
Comput Biol 16(4):e1007099-e
Jensen MK, Keasling JD (2015) Recent applications of synthetic biology tools for yeast metabolic
engineering. FEMS Yeast Res 15(1):1–10
Jeon WY, Shim WY, Lee SH, Choi JH, Kim JH (2013) Effect of heterologous xylose transporter
expression in Candida tropicalis on xylitol production rate. Bioprocess Biosyst Eng 36(6):809–
817
Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E (2012) A Pro-
grammable Dual-RNA–Guided DNA Endonuclease in Adaptive Bacterial Immunity. Science
337(6096):816–821
Johnson EA (2013) Biotechnology of non-Saccharomyces yeasts-the basidiomycetes. Appl Micro-
biol Biotechnol 97(17):7563–7577
Jones HD (2003) Genetic modification | transformation, general principles. In: Thomas B (ed)
Encyclopedia of applied plant sciences. Elsevier, Oxford, pp 377–382
Ju JH, Oh BR, Heo SY, Lee YU, Shon JH, Kim CH et al (2020) Production of adipic acid by short-
and long-chain fatty acid acyl-CoA oxidase engineered in yeast Candida tropicalis. Bioprocess
Biosyst Eng 43(1):33–43
Juers DH, Matthews BW, Huber RE (2012) LacZ β-galactosidase: structure and function of an
enzyme of historical and molecular biological importance. Protein Sci 21(12):1792–1807
Karkowska-Kuleta J, Kulig K, Karnas E, Zuba-Surma E, Woznicka O, Pyza E et al (2020) Charac-
teristics of extracellular vesicles released by the pathogenic yeast-like fungi Candida glabrata,
Candida parapsilosis and Candida tropicalis. Cells 9(7):1722
Kawaguchi Y, Honda H, Taniguchi-Morimura J, Iwasaki S (1989) The codon CUG is read as serine
in an asporogenic yeast Candida cylindracea. Nature 341(6238):164–166
Keppler-Ross S, Noffz C, Dean N (2008) A new purple fluorescent color marker for genetic studies
in Saccharomyces cerevisiae and Candida albicans. Genetics 179(1):705–710
Kim SH, Shin DH, Liu J, Oganesyan V, Chen S, Xu QS et al (2005) Structural genomics of minimal
organisms and protein fold space. J Struct Funct Genomics 6(2–3):63–70
Kim KH, Lee H-Y, Lee CY (2015) Pretreatment of sugarcane molasses and citric acid production
by Candida zeylanoides. Microbiol Biotechnol Lett 43(2):164–168
Köhler GA, White TC, Agabian N (1997) Overexpression of a cloned IMP dehydrogenase gene
of Candida albicans confers resistance to the specific inhibitor mycophenolic acid. J Bacteriol
179(7):2331–2338
Kosa P, Gavenciakova B, Nosek J (2007) Development of a set of plasmid vectors for genetic
manipulations of the pathogenic yeast Candida parapsilosis. Gene 396(2):338–345
Kouzuma A, Kato S, Watanabe K (2015) Microbial interspecies interactions: recent findings in
syntrophic consortia. Front Microbiol 6(477)
Križanović S, Butorac A, Mrvčić J, Krpan M, Cindrić M, Bačun-Družina V et al (2015) Character-
ization of a S-adenosyl-l-methionine (SAM)-accumulating strain of Scheffersomyces stipitis.
Int Microbiol 18(2):117–125
Synthetic Biology in the Candida (CTG) Clade 369
Lam CMC, Godinho M, dos Santos VAPM (2010) An introduction to synthetic biology. In:
Schmidt M, Kelle A, Ganguli-Mitra A, Vriend H (eds) Synthetic biology: the technoscience
and its societal consequences. Springer, Netherlands, Dordrecht, pp 23–48
Laplaza JM, Torres BR, Jin Y-S, Jeffries TW (2006) Sh ble and Cre adapted for functional
genomics and metabolic engineering of Pichia stipitis. Enzyme Microb Technol 38(6):741–747
Larbi NB, Jefferies C (2009) 2D-DIGE: comparative proteomics of cellular signalling pathways.
Methods in Molecular Biology (clifton, NJ) 517:105–132
Larsson DGJ (2014) Pollution from drug manufacturing: review and perspectives. Philos Trans R
Soc Lond B Biol Sci 369(1656):20130571
Ledesma-Amaro R, Santos MA, Jiménez A, Revuelta JL (2013) Microbial production of vitamins
(Chap 21). In: McNeil B, Archer D, Giavasis I, Harvey L (eds) Microbial production of food
ingredients, enzymes and nutraceuticals. Woodhead Publishing, pp 571–594
Leonard E, Nielsen D, Solomon K, Prather KJ (2008) Engineering microbes with synthetic biology
frameworks. Trends Biotechnol 26(12):674–681
Leuker CE, Hahn AM, Ernst JF (1992) beta-Galactosidase of Kluyveromyces lactis (Lac4p) as
reporter of gene expression in Candida albicans and C. tropicalis. Mol Gen Genet 235(2–
3):235–241
Li Y, Chen Y, Tian X, Chu J (2020) Advances in sophorolipid-producing strain perfor-
mance improvement and fermentation optimization technology. Appl Microbiol Biotechnol
104:10325–10337
Lin C-H, Choi A, Bennett RJ (2011) Defining pheromone-receptor signaling in Candida albicans
and related asexual Candida species. Mol Biol Cell 22(24):4918–4930
Liu S, Hu W, Wang Z, Chen T (2020) Production of riboflavin and related cofactors by biotechno-
logical processes. Microb Cell Fact 19(1):31
Liu GL, Fu GY, Chi Z, Chi ZM (2014) Enhanced expression of the codon-optimized exo-inulinase
gene from the yeast Meyerozyma guilliermondii in Saccharomyces sp. W0 and bioethanol
production from inulin. Appl Microbiol Biotechnol 98(21):9129–9138
Löbs AK, Schwartz C, Wheeldon I (2017) Genome and metabolic engineering in non-conventional
yeasts: current advances and applications. Synth Syst Biotechnol 2(3):198–207
Loeffler J, Hebart H, Magga S, Schmidt D, Klingspor L, Tollemar J et al (2000) Identification of
rare Candida species and other yeasts by polymerase chain reaction and slot blot hybridization.
Diagn Microbiol Infect Dis 38(4):207–212
Lombardi L, Turner SA, Zhao F, Butler G (2017) Gene editing in clinical isolates of Candida
parapsilosis using CRISPR/Cas9. Sci Rep 7(1):8051
Lombardi L, Oliveira-Pacheco J, Butler G (2019) Plasmid-based CRISPR-Cas9 gene editing in
multiple Candida species. mSphere 4(2)
Maguire SL, ÓhÉigeartaigh SS, Byrne KP, Schröder MS, O’Gaora P, Wolfe KH et al (2013) Com-
parative genome analysis and gene finding in Candida Species using CGOB. Mol Biol Evol
30(6):1281–1291
Maier T, Güell M, Serrano L (2009) Correlation of mRNA and protein in complex biological
samples. FEBS Lett 583(24):3966–3973
Maldonado I, Cataldi S, Garbasz C, Relloso S, Striebeck P, Guelfand L et al (2018) Identification of
Candida yeasts: conventional methods and MALDI-TOF MS. Rev Iberoam Micol 35(3):151–
154
Mancera E, Frazer C, Porman AM, Ruiz-Castro S, Johnson AD, Bennett RJ (2019) Genetic mod-
ification of closely related Candida species. Front Microbiol 10(357)
Marian M, Shimizu M (2019) Improving performance of microbial biocontrol agents against plant
diseases. J Gen Plant Pathol 85(5):329–336
Marton T, Maufrais C, d’Enfert C, Legrand M (2020) Use of CRISPR-Cas9 to target homologous
recombination limits transformation-induced genomic changes in Candida albicans. mSphere
5(5)
Masuda Y, Park SM, Ohkuma M, Ohta A, Takagi M (1994) Expression of an endogenous and a het-
erologous gene in Candida maltosa by using a promoter of a newly-isolated phosphoglycerate
kinase (PGK) gene. Curr Genet 25(5):412–417
370 D. Kasir et al.
McCarty NS, Ledesma-Amaro R (2019) Synthetic biology tools to Engineer microbial communi-
ties for biotechnology. Trends Biotechnol 37(2):181–197
McLellan MA, Rosenthal NA, Pinto AR (2017) Cre-loxP-mediated recombination: general prin-
ciples and experimental considerations. Curr Protoc Mouse Biol 7(1):1–12
Messing R, Brodeur J (2018) Current challenges to the implementation of classical biological
control. Biocontrol 63(1):1–9
Michel S, Ushinsky S, Klebl B, Leberer E, Thomas D, Whiteway M et al (2002) Generation of
conditional lethal Candida albicans mutants by inducible deletion of essential genes. Mol
Microbiol 46(1):269–280
Millerioux Y, Clastre M, Simkin AJ, Courdavault V, Marais E, Sibirny AA et al (2011) Drug-
resistant cassettes for the efficient transformation of Candida guilliermondii wild-type strains.
FEMS Yeast Res 11(6):457–463
Min K, Ichikawa Y, Woolford CA, Mitchell AP (2016) Candida albicans gene deletion with a
transient CRISPR-Cas9 system. mSphere 1(3)
Min et al (2016) Candida albicans gene deletion with a transient CRISPR-Cas9 system. https://
doi.org/10.1128/mSphere.00130-16
Mishra P, Park GY, Lakshmanan M, Lee HS, Lee H, Chang MW et al (2016) Genome-scale
metabolic modeling and in silico analysis of lipid accumulating yeast Candida tropicalis for
dicarboxylic acid production. Biotechnol Bioeng 113(9):1993–2004
Möckli N, Auerbach D (2004) Quantitative β-galactosidase assay suitable for high-throughput
applications in the yeast two-hybrid system. Biotechniques 36(5):872–876
Moreno-Ruiz E, Ortu G, de Groot PWJ, Cottier F, Loussert C, Prévost M-C et al (2009) The GPI-
modified proteins Pga59 and Pga62 of Candida albicans are required for cell wall integrity.
Microbiology 155(6):2004–2020
Morschhäuser J, Michel S, Hacker J (1998) Expression of a chromosomally integrated, single-copy
GFP gene in Candida albicans, and its use as a reporter of gene regulation. Mol Gen Genet
MGG 257(4):412–420
Morschhäuser J, Michel S, Staib P (1999) Sequential gene disruption in Candida albicans by FLP-
mediated site-specific recombination. Mol Microbiol 32(3):547–556
Mühlhausen S, Kollmar M (2014) Molecular phylogeny of sequenced Saccharomycetes reveals
polyphyly of the alternative yeast codon usage. Genome Biol Evol 6(12):3222–3237
Nakayama H, Mio T, Nagahashi S, Kokado M, Arisawa M, Aoki Y (2000) Tetracycline-regulatable
system to tightly control gene expression in the pathogenic fungus Candida albicans. Infect
Immun 68(12):6712–6719
Narad P, Kirthanashri SV (2018) Introduction to omics. In: Arivaradarajan P, Misra G (eds) Omics
approaches, technologies and applications: integrative approaches for understanding OMICS
data. Springer Singapore, Singapore, pp 1–10
Nguyen LN, Trofa D, Nosanchuk JD (2009) Fatty acid synthase impacts the pathobiology of
Candida parapsilosis in vitro and during mammalian infection. PLoS One 4(12):e8421
Nguyen N, Quail MMF, Hernday AD (2017) An efficient, rapid, and recyclable system for
CRISPR-mediated genome editing in Candida albicans. mSphere 2(2)
Norton EL, Sherwood RK, Bennett RJ (2017) Development of a CRISPR-Cas9 system for efficient
genome editing of Candida lusitaniae. mSphere 2(3)
Nosek J, Holesova Z, Kosa P, Gacser A, Tomaska L (2009) Biology and genetics of the pathogenic
yeast Candida parapsilosis. Curr Genet 55(5):497–509
Nunes-Düby SE, Kwon HJ, Tirumalai RS, Ellenberger T, Landy A (1998) Similarities and differ-
ences among 105 members of the Int family of site-specific recombinases. Nucleic Acids Res
26(2):391–406
Obando Montoya EJ, Mélin C, Blanc N, Lanoue A, Foureau E, Boudesocque L et al (2014) Dis-
rupting the methionine biosynthetic pathway in Candida guilliermondii: characterization of the
MET2 gene as counter-selectable marker. Yeast 31(7):243–251
Olivares-Hernández R, Usaite R, Nielsen J (2010) Integrative analysis using proteome and tran-
scriptome data from yeast to unravel regulatory patterns at post-transcriptional level. Biotech-
nol Bioeng 107(5):865–875
Synthetic Biology in the Candida (CTG) Clade 371
Olivares-Hernández R, Bordel S, Nielsen J (2011) Codon usage variability determines the corre-
lation between proteome and transcriptome fold changes. BMC Syst Biol 5(1):33
Orr-Weaver TL, Szostak JW, Rothstein RJ (1981) Yeast transformation: a model system for the
study of recombination. Proc Natl Acad Sci U S A 78(10):6354–6358
Orth JD, Thiele I, Palsson BØ (2010) What is flux balance analysis? Nat Biotechnol 28(3):245–248
Papon N, Courdavault V, Clastre M, Simkin AJ, Crèche J, Giglioli-Guivarc’h N (2012) Deus ex
Candida genetics: overcoming the hurdles for the development of a molecular toolbox in the
CTG clade. Microbiology 158(Pt 3):585–600
Papon N, Savini V, Lanoue A, Simkin AJ, Crèche J, Giglioli-Guivarc’h N et al (2013) Candida
guilliermondii: biotechnological applications, perspectives for biological control, emerging
clinical importance and recent advances in genetics. Curr Genet 59(3):73–90
Papon N, Courdavault V, Clastre M (2014) Biotechnological potential of the fungal CTG clade
species in the synthetic biology era. Trends Biotechnol 32(4):167–168
Park Y-N, Morschhäuser J (2005) Tetracycline-inducible gene expression and gene deletion in
Candida albicans. Eukaryot Cell 4(8):1328–1342
Passoth V, Cohn M, Schäfer B, Hahn-Hägerdal B, Klinner U (2003) Analysis of the hypoxia-
induced ADH2 promoter of the respiratory yeast Pichia stipitis reveals a new mechanism for
sensing of oxygen limitation in yeast. Yeast 20(1):39–51
Patra P, Das M, Kundu P, Ghosh A (2021) Recent advances in systems and synthetic biology
approaches for developing novel cell-factories in non-conventional yeasts. Biotechnol Adv
47:107695
Pereira SC, Maehara L, Machado CMM, Farinas CS (2015) 2G ethanol from the whole sugarcane
lignocellulosic biomass. Biotechnol Biofuels 8(1):44
Pharkya P, Maranas C (2006) An optimization framework for identifying reaction activa-
tion/inhibition or elimination candidates for overproduction in microbial systems. Metab Eng
8:1–13
Pharkya P, Burgard A, Maranas C (2004) OptStrain: A computational framework for redesign of
microbial production systems. Genome Res 14:2367–2376
Pickford R (2019) Mass spectrometry-based metabolomic analysis. In: Ranganathan S, Gribskov
M, Nakai K, Schönbach C (eds) Encyclopedia of bioinformatics and computational biology.
Academic Press, Oxford, pp 410–425
Polen T, Spelberg M, Bott M (2013) Toward biotechnological production of adipic acid and pre-
cursors from biorenewables. J Biotechnol 167(2):75–84
Price ND, Reed JL, Palsson B (2004) Genome-scale models of microbial cells: evaluating the
consequences of constraints. Nat Rev Microbiol 2(11):886–897
Prista C, Michán C, Miranda IM, Ramos J (2016) The halotolerant Debaryomyces hansenii, the
Cinderella of non-conventional yeasts. Yeast 33(10):523–533
Raghavachari N (2011) Overview of omics pp 1–20
Ramírez-Ramírez R, Calvo-Méndez C, Avila-Rodriguez M, Lappe-Oliveras P, Ulloa M, Vázques-
Juárez R et al (2004) Cr(VI) reduction in chromate-resistant strain of Candida maltosa isolated
from the leather industry. Antonie Van Leeuwenhoek 85:63–68
Ratcliffe RG, Shachar-Hill Y (2005) Revealing metabolic phenotypes in plants: inputs from NMR
analysis. Biol Rev Camb Philos Soc 80(1):27–43
Reijnst P, Walther A, Wendland J (2011) Dual-colour fluorescence microscopy using yEmCherry-
/GFP-tagging of eisosome components Pil1 and Lsp1 in Candida albicans. Yeast 28(4):331–
338
Ren J, Lee J, Na D (2020) Recent advances in genetic engineering tools based on synthetic biology.
Microbiology 58(1):1–10
Reuss O, Vik A, Kolter R, Morschhäuser J (2004) The SAT1 flipper, an optimized tool for gene
disruption in Candida albicans. Gene 341:119–127
Revuelta JL, Buey RM, Ledesma-Amaro R, Vandamme EJ (2016) Microbial biotechnology for the
synthesis of (pro)vitamins, biopigments and antioxidants: challenges and opportunities. Microb
Biotechnol 9(5):564–567
372 D. Kasir et al.
Riley R, Haridas S, Wolfe KH, Lopes MR, Hittinger CT, Göker M et al (2016) Comparative
genomics of biotechnologically important yeasts. Proc Natl Acad Sci U S A 113(35):9882–
9887
Roa Engel CA, Straathof AJJ, Zijlmans TW, van Gulik WM, van der Wielen LAM (2008) Fumaric
acid production by fermentation. Appl Microbiol Biotechnol 78(3):379–389
Roda A, Pasini P, Mirasoli M, Michelini E, Guardigli M (2004) Biotechnological applications of
bioluminescence and chemiluminescence. Trends Biotechnol 22(6):295–303
Rodrussamee N, Sattayawat P, Yamada M (2018) Highly efficient conversion of xylose to
ethanol without glucose repression by newly isolated thermotolerant Spathaspora passali-
darum CMUWF1–2. BMC Microbiol 18(1):73
Roemer T, Jiang B, Davison J, Ketela T, Veillette K, Breton A et al (2003) Large-scale essen-
tial gene identification in Candida albicans and applications to antifungal drug discovery. Mol
Microbiol 50(1):167–181
Røkke G, Korvald E, Pahr J, Øyås O, Lale R (2014) BioBrick assembly standards and techniques
and associated software tools. In: Valla S, Lale R (eds) DNA cloning and assembly methods.
Humana Press, Totowa, NJ, pp 1–24
Roldão A, Kim I-K, Nielsen J (2012) Bridging omics technologies with synthetic biology in
yeast industrial biotechnology. In: Wittmann C, Lee SY (eds) Systems metabolic engineering.
Springer, Netherlands, Dordrecht, pp 271–327
Ruchala J, Kurylenko OO, Dmytruk KV, Sibirny AA (2020) Construction of advanced produc-
ers of first- and second-generation ethanol in Saccharomyces cerevisiae and selected species
of non-conventional yeasts (Scheffersomyces stipitis, Ogataea polymorpha). J Ind Microbiol
Biotechnol 47(1):109–132
Samaranayake DP, Hanes SD (2011) Milestones in Candida albicans gene manipulation. Fungal
Genet Biol: FG & B 48(9):858–865
Sampaio P, Gusmão L, Correia A, Alves C, Rodrigues AG, Pina-Vaz C et al (2005) New
microsatellite multiplex PCR for Candida albicans strain typing reveals microevolutionary
changes. J Clin Microbiol 43(8):3869–3876
Sánchez-Martínez C, Pérez-Martín J (2002) Site-specific targeting of exogenous DNA into the
genome of Candida albicans using the FLP recombinase. Mol Genet Genomics 268(3):418–
424
Santos MAS, Moura G, Massey SE, Tuite MF (2004) Driving change: the evolution of alternative
genetic codes. Trends Genet: TIG 20(2):95–102
Santos MA, Gomes AC, Santos MC, Carreto LC, Moura GR (2011) The genetic code of the fungal
CTG clade. C R Biol 334(8–9):607–611
Schellenberger J, Que R, Fleming RMT, Thiele I, Orth JD, Feist AM et al (2011) Quantitative pre-
diction of cellular metabolism with constraint-based models: the COBRA Toolbox v2.0. Nature
Protocols 6(9):1290–1307
Schindler D (2020) Genetic engineering and synthetic genomics in yeast to understand life and
boost biotechnology. Bioengineering 7(4)
Schulze S, Schleicher J, Guthke R, Linde J (2016) How to predict molecular interactions between
species? Front Microbiol 7:442
Segal E, Yehuda H, Droby S, Wisniewski M, Goldway M (2002) Cloning and analysis of CoEXG1,
a secreted 1,3-β-glucanase of the yeast biocontrol agent Candida oleophila. Yeast 19(13):1171–
1182
Shahana S, Childers DS, Ballou ER, Bohovych I, Odds FC, Gow NA et al (2014) New clox systems
for rapid and efficient gene disruption in Candida albicans. PLoS One 9(6):e100390
Shen J, Guo W, Köhler JR (2005) CaNAT1, a heterologous dominant selectable marker for transfor-
mation of Candida albicans and other pathogenic Candida species. Infect Immun 73(2):1239–
1242
Shen X-X, Zhou X, Kominek J, Kurtzman CP, Hittinger CT, Rokas A (2016) Reconstructing
the backbone of the saccharomycotina yeast phylogeny using genome-scale data. G3 Genes
Genomes Genet 6(12):3927–3939
Synthetic Biology in the Candida (CTG) Clade 373
Shi J, Zhang M, Zhang L, Wang P, Jiang L, Deng H (2014) Xylose-fermenting Pichia stipitis by
genome shuffling for improved ethanol production. Microb Biotechnol 7(2):90–99
Shin M, Kim JW, Ye S, Kim S, Jeong D, Lee DY et al (2019) Comparative global metabolite pro-
filing of xylose-fermenting Saccharomyces cerevisiae SR8 and Scheffersomyces stipitis. Appl
Microbiol Biotechnol 103(13):5435–5446
Singh Dhillon G, Kaur Brar S, Verma M, Tyagi RD (2011) Recent advances in citric acid bio-
production and recovery. Food Bioprocess Technol 4(4):505–529
Smale ST (2010) Beta-galactosidase assay. Cold Spring Harb Protoc 2010(5):pdb.prot5423
Spasskaya DS, Kotlov MI, Lekanov DS, Tutyaeva VV, Snezhkina AV, Kudryavtseva AV et al
(2021) CRISPR/Cas9-mediated genome engineering reveals the contribution of the 26S pro-
teasome to the extremophilic nature of the yeast Debaryomyces hansenii. ACS Synthetic Biol
Sprengel R, Hasan MT (2007) Tetracycline-controlled genetic switches. Handb Exp Pharmacol
178:49–72
Sreenath HK, Jeffries TW (1999) 2-Deoxyglucose as a selective agent for derepressed mutants of
Pichia stipitis. Appl Biochem Biotechnol 77(1):211–222
Srikantha T, Klapach A, Lorenz WW, Tsai LK, Laughlin LA, Gorman JA et al (1996) The sea
pansy Renilla reniformis luciferase serves as a sensitive bioluminescent reporter for differential
gene expression in Candida albicans. J Bacteriol 178(1):121–129
Stagljar I (2016) The power of OMICs. Biochem Biophys Res Commun 479(4):607–609
Staib P, Kretschmar M, Nichterlein T, Köhler G, Michel S, Hof H et al (1999) Host-induced,
stage-specific virulence gene activation in Candida albicans during infection. Mol Microbiol
32(3):533–546
Staib P, Kretschmar M, Nichterlein T, Hof H, Morschhäuser J (2000a) Differential activation of a
Candida albicans virulence gene family during infection. PNAS 97(11):6102–6107
Staib P, Michel S, Köhler G, Morschhäuser J (2000b) A molecular genetic system for the
pathogenic yeast Candida dubliniensis. Gene 242(1–2):393–398
Stenuit B, Agathos SN (2015) Deciphering microbial community robustness through synthetic
ecology and molecular systems synecology. Curr Opin Biotechnol 33:305–317
Sternberg N, Hamilton D (1981) Bacteriophage P1 site-specific recombination. I. Recombination
between loxP sites. J Mol Biol 150(4):467–486
Stoneman HR, Wrobel RL, Place M, Graham M, Krause DJ, De Chiara M et al (2020) CRISpy-
pop: a web tool for designing CRISPR/Cas9-driven genetic modifications in diverse popula-
tions. G3 Genes Genomes Genet
Sun C, Huang Y, Lian S, Saleem M, Li B, Wang C (2021) Improving the biocontrol efficacy of
Meyerozyma guilliermondii Y-1 with melatonin against postharvest gray mold in apple fruit.
Postharvest Biol Technol 171:111351
Tang S-J, Sun K-H, Sun G-H, Chang T-Y, Wu W-L, Lee G-C (2003) A transformation system
for the nonuniversal CUGSer codon usage species Candida rugosa. J Microbiol Methods
52(2):231–238
Tanner FW Jr, Vojnovich C, Van Lanen JM (1945) Riboflavin production by Candida species.
Science 101(2616):180–1
Treu R, Falandysz J (2017) Mycoremediation of hydrocarbons with basidiomycetes—a review. J
Environ Sci Health B 52(3):148–155
Tsoi R, Wu F, Zhang C, Bewick S, Karig D, You L (2018) Metabolic division of labor in microbial
systems. PNAS 115(10):2526–2531
Tsui CKM, Daniel H-M, Robert V, Meyer W (2008) Re-examining the phylogeny of clinically
relevant Candida species and allied genera based on multigene analyses. FEMS Yeast Res
8(4):651–659
Turner SA, Butler G (2014) The Candida pathogenic species complex. Cold Spring Harb Perspect
Med 4(9):a019778
Uhl MA, Johnson AD (2001) Development of Streptococcus thermophilus lacZ as a reporter gene
for Candida albicans. Microbiology 147(Pt 5):1189–1195
Unrean P, Jeennor S, Laoteng K (2016) Systematic development of biomass overproducing Schef-
fersomyces stipitis for high-cell-density fermentations. Synth Syst Biotechnol 1(1):47–55
374 D. Kasir et al.
Younes S, Bracharz F, Awad D, Qoura F, Mehlmer N, Brueck T (2020) Microbial lipid production
by oleaginous yeasts grown on Scenedesmus obtusiusculus microalgae biomass hydrolysate.
Bioprocess Biosyst Eng 43(9):1629–1638
Yu KO, Jung J, Kim SW, Park CH, Han SO (2012) Synthesis of FAEEs from glycerol in engineered
Saccharomyces cerevisiae using endogenously produced ethanol by heterologous expression of
an unspecific bacterial acyltransferase. Biotechnol Bioeng 109(1):110–115
Zhang C, Konopka JB (2010) A photostable green fluorescent protein variant for analysis of protein
localization in Candida albicans. Eukaryot Cell 9(1):224–226
Zhang Y, Jia D, Sun W, Yang X, Zhang C, Zhao F et al (2018) Semicontinuous sophorolipid fer-
mentation using a novel bioreactor with dual ventilation pipes and dual sieve-plates coupled
with a novel separation system. Microb Biotechnol 11(3):455–464
Zhang X, Li B, Zhang Z, Chen Y, Tian S (2020) Antagonistic yeasts: a promising alternative to
chemical fungicides for controlling postharvest decay of fruit. J Fungi 6(3):158