ArticlePDF AvailableLiterature Review

The Transporter-Mediated Cellular Uptake and Efflux of Pharmaceutical Drugs and Biotechnology Products: How and Why Phospholipid Bilayer Transport Is Negligible in Real Biomembranes

Authors:

Abstract and Figures

Over the years, my colleagues and I have come to realise that the likelihood of pharmaceutical drugs being able to diffuse through whatever unhindered phospholipid bilayer may exist in intact biological membranes in vivo is vanishingly low. This is because (i) most real biomembranes are mostly protein, not lipid, (ii) unlike purely lipid bilayers that can form transient aqueous channels, the high concentrations of proteins serve to stop such activity, (iii) natural evolution long ago selected against transport methods that just let any undesirable products enter a cell, (iv) transporters have now been identified for all kinds of molecules (even water) that were once thought not to require them, (v) many experiments show a massive variation in the uptake of drugs between different cells, tissues, and organisms, that cannot be explained if lipid bilayer transport is significant or if efflux were the only differentiator, and (vi) many experiments that manipulate the expression level of individual transporters as an independent variable demonstrate their role in drug and nutrient uptake (including in cytotoxicity or adverse drug reactions). This makes such transporters valuable both as a means of targeting drugs (not least anti-infectives) to selected cells or tissues and also as drug targets. The same considerations apply to the exploitation of substrate uptake and product efflux transporters in biotechnology. We are also beginning to recognise that transporters are more promiscuous, and antiporter activity is much more widespread, than had been realised, and that such processes are adaptive (i.e., were selected by natural evolution). The purpose of the present review is to summarise the above, and to rehearse and update readers on recent developments. These developments lead us to retain and indeed to strengthen our contention that for transmembrane pharmaceutical drug transport “phospholipid bilayer transport is negligible”.
Content may be subject to copyright.
Molecules 2021, 26, 5629. https://doi.org/10.3390/molecules26185629 www.mdpi.com/journal/molecules
Article
The Transporter-Mediated Cellular Uptake and Efflux
of Pharmaceutical Drugs and Biotechnology Products: How
and Why Phospholipid Bilayer Transport Is Negligible
in Real Biomembranes
Douglas B. Kell 1,2,3
1 Department of Biochemistry and Systems Biology, Institute of Systems, Molecular and Integrative Biology,
University of Liverpool, Crown St, Liverpool L69 7ZB, UK; dbk@liv.ac.uk
2 Novo Nordisk Foundation Centre for Biosustainability, Technical University of Denmark, Building 220,
Kemitorvet, 2800 Kgs Lyngby, Denmark
3 Mellizyme Biotechnology Ltd., IC1, Liverpool Science Park, Mount Pleasant, Liverpool L3 5TF, UK
Abstract: Over the years, my colleagues and I have come to realise that the likelihood of pharma-
ceutical drugs being able to diffuse through whatever unhindered phospholipid bilayer may exist
in intact biological membranes in vivo is vanishingly low. This is because (i) most real biomembranes
are mostly protein, not lipid, (ii) unlike purely lipid bilayers that can form transient aqueous chan-
nels, the high concentrations of proteins serve to stop such activity, (iii) natural evolution long ago
selected against transport methods that just let any undesirable products enter a cell, (iv) transport-
ers have now been identified for all kinds of molecules (even water) that were once thought not to
require them, (v) many experiments show a massive variation in the uptake of drugs between dif-
ferent cells, tissues, and organisms, that cannot be explained if lipid bilayer transport is significant
or if efflux were the only differentiator, and (vi) many experiments that manipulate the expression
level of individual transporters as an independent variable demonstrate their role in drug and nu-
trient uptake (including in cytotoxicity or adverse drug reactions). This makes such transporters
valuable both as a means of targeting drugs (not least anti-infectives) to selected cells or tissues and
also as drug targets. The same considerations apply to the exploitation of substrate uptake and
product efflux transporters in biotechnology. We are also beginning to recognise that transporters
are more promiscuous, and antiporter activity is much more widespread, than had been realised,
and that such processes are adaptive (i.e., were selected by natural evolution). The purpose of the
present review is to summarise the above, and to rehearse and update readers on recent develop-
ments. These developments lead us to retain and indeed to strengthen our contention that for trans-
membrane pharmaceutical drug transport phospholipid bilayer transport is negligible”.
Keywords: membrane transport; pharmaceuticals; drugs; energy coupling; biotechnology; ADME;
DMPK; transporter engineering
1. Introduction
Over the years, two main ideas have been used to explain the mechanisms by which
pharmaceutical drugs and/or substrates and products of biotechnological interest pass
through the plasma (or other) membranes of the relevant organism: in a more classical
analysis, it is assumed (and it really is purely an assumption [1]) that molecules can dif-
fuse through the core of the lipid bilayer by some means. In an alternative and really en-
tirely opposite view, which we refer to as PBIN for phospholipid bilayer diffusion is
negligible[2], we have argued [112] (and these constitute background that we mainly
do not rehearse again here) that this does not in fact occur in real biomembranes to any
Citation:
Kell, D.B. The Transporter-
Mediated Cellular Upt ake and Efflux
of Pharmaceutical Drugs and
Biotechnology Products: How and
Why Phospholipid Bilayer Transport
Is Negligible in Real Biomembranes.
Molecules
2021, 26, 5629.
https://doi.org/
10.3390/molecules26185629
Academic
Editors: Kristiina
Huttunen and Santosh Kumar Adla
Received:
17 August 2021
Accepted:
14 September 2021
Published:
16 September 2021
Publisher’s Note:
MDPI stays neu-
tral with regard to jurisdictional
claims in published maps and institu-
tional affiliations.
C
opyright: © 2021 by the author. Li-
censee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and con-
ditions of the Creative
Commons At-
tribution (CC BY) license (http://crea-
tivecommons.org/licenses/by/4.0/).
Molecules 2021, 26, 5629 2 of 39
significant extent at all. Some of this background material is also available in a webinar
(https://bit.ly/3yQJ1FG, accessed on 15 September 2021). The two modes are illustrated in
Figure 1. One may debate what is significantand/or negligible, but anything less than
5% of a total flux really is not much of a contribution, and plenty of evidence implies that
it is almost certainly less than 1%. In our view, the term “negligible” covers this more than
adequately. The purpose of this review and commentary is to provide an update on some
of the salient issues. Examples of the importance of transporters continue to grow apace,
whilewhatever may or may not happen in artificial membranesit remains the case
that the actual evidence for significant transmembrane transport solely through any bi-
layer portions of intact biological membranes is non-existent. We shall start by looking at
why this is the case, initially by focusing on some of the major differences between real
biomembranes and those artificial ones made from phospholipid bilayers.
Figure 1. A drug D might pass through a biological membrane in one of two main ways conceptually (we do not here
discuss endo- and exocytosis; the focus is only on cases where the drug is considered to cross through the membrane
barrier itself). On the left is illustrated transport through phospholipid bilayers, while on the right we illustrate the use of
proteinaceous solute carriers to effect entry and exit of the drug. The crux of this review is that the mode on the left does
not take place at any meaningful rate in intact biological membranes (since they have a high protein content).
1.1. Biological Membrane Structure
The textbook view of biological membranes is that they are to be seen as a fluid
mosaic[13,14] of proteins embedded within a phospholipid bilayer, along with other
small molecules such as sterols. The original article [13] featured a now iconic illustration
(shown in Figure 3 of said article) in which a small number of proteins were embedded in
How drugs might pass through
cellular membranes
OR
via
phospholipid
bilayer
via one or more
of hundreds of
transporters
Molecules 2021, 26, 5629 3 of 39
and on a seaof phospholipids, with the proteins representing (by eye) approximately
one-seventh of both the total mass and area. Unfortunately, this picture is very mislead-
ing, since biological membranes are commonly 3:1 protein:lipid by mass, and maybe 1:1
by area (not 1:7) [1528] (see Figure 2). This widespread but erroneous mental picture, of
proteins floating in a sea of phospholipids, has led many to suppose that from a bio-
physical point of view biomembranes are thus essentially just like pure lipid bilayers, in
that the supposedly sparse proteins would do little or nothing to affect the kinds of prop-
erties that can be seen in purely phospholipid bilayers such as those studied in vitro [29
31]. However, this is emphatically not the case.
Purely artificial (solely phospholipid) bilayer membranes are not terribly stable and
admit the passage of small molecules (and even ions) via transient aqueous pores or chan-
nels (e.g., [3243]). This mode of transfer clearly has to be the case if ostensibly ‘transmem-
brane’ transport ‘through’ them is observed, as it can easily be calculated that the free
energy necessary for passing a small cation through a dielectric (such as that represented
by a membrane interior of phospholipid tails) with a permittivity of ca 2 is so great that it
would be unlikely to occur even over millions of years [44]. More specifically, the lateral
flexing of phospholipids needed to make such transient aqueous pores simply cannot oc-
cur when the phospholipids are either bound directly to a protein or are strongly influ-
enced thereby (as they are, e.g., [4562]). This is likely reinforced by the well-established,
substantial and dynamic lipid asymmetry between the inner and outer halves of mem-
branes [6365]. Note too that ABC transporters are responsible for moving lipids them-
selves around cells and within membranes [47,6669]. In short, studying an aardvaark
tells you about an aardvark, not a langoustine. In a similar way, studying sodium chloride
tells you about sodium chloride, not strychnine chloride, even if they both contain chlo-
ride. Thus, in a similar way, studying pure phospholipid bilayers tells you about pure
phospholipid bilayers, and not biomembranes that may happen to contain relatively small
amounts phospholipids in a bilayer form. This may seem obvious when set out in this
way, which is why we do it.
Figure 2. Cartoon of a typical biomembrane indicating the relative paucity of phopholipid bilayer
that is uninfluenced by proteins. Taken from an Open Access animation covering some of this
ground and related transporter matters at https://www.youtube.com/watch?v=s23vNwLE-Jw.
Protein and lipid amounts as they
are in a typical biomembrane
Molecules 2021, 26, 5629 4 of 39
1.2. Transport of Drugs through Biological Membranes and the (Mis)Use of the Term “Passive”
It is now well-established that cells require proteinaceous transporters to effect the
transmembrane transport of the nutrients they need to survive and to grow. Since these
nutrients are mainly small molecules, it might be thought that a similar degree of ac-
ceptance would accord to the assumption that this held true for pharmaceutical drugs too,
as well as for the uptake and efflux of substances of interest to the biotechnology industry.
Surprisingly, as mooted above, and for some comparable reasons with bioenergetics more
generally [70], this has not largely been the case. Indeed, why these and other beliefs per-
sist despite the facts, and what to do about it, is of general philosophical and psychological
interest [7176]. In this case, it seems from our experience that it is to do with culture and
education; those versed in physical organic chemistry (who are often predominant in the
DMPK/ADMET sections of pharmaceutical companies but also in more traditional phar-
macology schools in academia) tend not to know much about enzymology, while molec-
ular biologistswho tend to know little of physical organic chemistryexpress extreme
surprise when told of the widespread belief of physical organic chemists that drugs
simply pass through the bilayer portions of biomembranes. There is clearly value for in-
dividuals from both backgrounds in learning a little of each topic.
The term passive, as employed for instance in the phrase passive diffusion, is
also widely misused, not least because it can be (and is) used to cover and then conflate
two separate concepts. The first represents a thermodynamic meaning, where passive
is used to mean equilibrative: the transporter requires no free energy and simply lets
molecules pass down their concentration gradients until their concentrations (strictly,
thermodynamic activities) are the same on each side of the membrane of interest (i.e., that
in which the transporter is embedded). Clearly, the thermodynamic usage can (or should)
have no mechanistic implications. The problem is that the same term (as in passive dif-
fusionor passive permeability) is also used to imply a mechanism, viz that such equili-
brative diffusion occurs through the phospholipid bilayer. In the worst cases, demonstra-
tion of the thermodynamic property is used to imply or even to claim the demonstration
of a mechanism as transbilayer diffusion. In our view, the term passivehas acquired so
much baggage that the only solution to this (Figure 3 and reference [2]), in addition to
education, is to avoid the term passivecompletely, and thus be forced to be more ex-
plicit about precisely what it is that is being claimed in any particular case.
Figure 3. Suggested terminologies to avoid the use of the term “passive”, which is still widely mis-
used to conflate two entirely separate concepts, one thermodynamic (hence independent of mecha-
nism) and one mechanistic.
Means of transport
Purely lipoidal Transporter-mediated
Thermodynamics
Concentra�ve
Equilibra�ve
Concentra�ve
bilayer diffusion
(coupled e.g. to a
pH gradient)
Equilibra�ve
Bilayer diffusion
Facilitated
diffusion
The terminology of transport reac�ons
Ac�ve transport
Molecules 2021, 26, 5629 5 of 39
One feature of transport across membrane systems such as those of the popular Caco-
2 cell monolayers [7784], that (like other human tissues [85,86]; https://www.proteinat-
las.org/search/slc), express hundreds of transporter proteins [8791], is that the initial rate
of transport of a given substrate at the same initial concentration from side A side B can
differ from that when the test is made from side B side A, even when the transport is
purely equilibrative. This has no ready explanation by or for those schooled purely in
physico-chemical concepts and who believe that diffusion explains everything. By con-
trast, enzymologists have a perfectly straightforward explanation of this, which comes
from the well-known thermodynamic Haldane relationship relating the Km and Vmax val-
ues in the forward (Km,f and Vmax,f) and reverse (Km,r and Vmax,r) directions of a reaction to
its equilibrium constant Keq. While this can be found in any textbook of enzyme kinetics
(e.g., [9294]), or even of biochemistry, we reproduce it below:
(Vmax,f · Km,r)/(Vmax,r · Km,f) = Keq (1
)
Thus, even for Keq values of unity, it is easily possible to have rates that differ
manyfold in the two directions for the same external concentration, just by manipulating
the other values in Equation (1) while keeping their ratios consistent with it. Of course,
and I stress this purposely, this is true only for enzyme- (transporter)-mediated transport,
not diffusionthrough lipid bilayers [5].
1.3. Untestability of Bilayer Transport Models in Real Biomembranes
Perhaps surprisingly, the view that molecules pass through phospholipid bilayers in
real biological membranes is presently untestable, because we do not have the ability to
image small molecules and membranes with the atomic resolution that would be neces-
sary to observe their transport directly. What is presently carried out most commonly is
to observe molecules on one side of a membrane and later on the other side, and simply
assume that they appeared there via transbilayer transport. This is a well-established and
classical logical fallacy known as affirming the precedentor post hoc ergo propter hoc
[1]. In a different vein, observing a molecule transferring from one side of a pure phos-
pholipid bilayer simply does not tell you what it might do in a real membrane where there
is very little bilayer that is not affected by the presence of protein. Similarly, changing the
type or amount of phospholipid does not simplydo that, because the activities of mem-
brane proteins, including transporters [95], can vary quite substantially with changes in
their adjacent lipids. Consequently, any changes in transport induced by changing lipids
can perfectly well (and more accurately) be explained by their influences on the proteins
embedded in the membrane.
1.4. Testability of Transporter-Mediated Models in Real Biomembranes
By contrast, those of us who claim that bilayer transport is negligible in real biomem-
branes can easily change the expression levels or activities of transporters of interest (e.g.,
[96,97]) and observe the concomitant effects on the transport of the small molecules of
interest (as in Figure 4). (The same strategy can also be used to detect inhibitors of the
transporter, e.g., [98,99].) In the paper of Winter et al. [97], the uptake and toxicity of an
anticancer compound called YM155 or sepantronium bromide was decreased some 500-
fold when the single transporter SLC35F2 was knocked out, and the toxicity of the com-
pound correlated closely with that transporter’s expression level over some four orders of
magnitude when explored in some 15 separate cell lines (that, as for many transporters
[86,100], showed equivalently massively varying expression levels). The only reasonable
(one might say plausible) explanation is that the SLC35F2 transporter normally carried
some 99.5% of the flux of YM155 into the cell, and that consequently the transmembrane
flux occurring by any other means, including via bilayer transport, is indeed negligible.
This simple phenomenon served to explain entirely the highly variable efficacy of YM155
in a series of clinical trials.
Molecules 2021, 26, 5629 6 of 39
Figure 4. Principle of determining the substrate of a drug at toxic concentrations (e.g., [8,96,97,101,102]) by assessing the
ability of cells lacking a particular transporter gene to survive its presence, while the wild-type cells, or cells knocked out
for other genes not involved in the drug’s transport, are killed. Obviously, this is the extreme; there may also be degrees
of resistance.
Typically, the question of the interaction between substrates and transporters can be
phrased in two complementary ways: the transporter-centric question is given a trans-
porter of interest X, what are its substrates?”. This is the problem of de-orphanisation
[103109]. The complementary, substrate-centric, question is given a substrate Y, which
transporter(s) is/are responsible for its cellular uptake and/or efflux?. The latter is the
more important one for assessing the mechanisms of drug transport [8], and there are
many examples (e.g., [110118]) where transporter activity has been identified but where
the transporters involved have not. Of course, while a known result is a partial answer to
each question, the strategies for tackling them are slightly different (for a recent overview
of cell-based assay methods for SLCs, see [119]). Note, in particular, that simple biophysics
plus the re-use of protein motifs in evolution means that most small molecules bind to
multiple targets, and most proteins can bind small molecules promiscuously (e.g.,
[8,9,120130]).
As part of the EU-IMI ReSolute project [131] (https://re-solute.eu/), the CEMM group
studied 60 cytotoxic compounds. By using an SLC-focused CRISPR-Cas9 library, they
identified a series of transporters whose absence induced resistance to the drugs tested
[101], much as had been achieved in smaller numbers before [96,97]. This kind of sub-
strate-centric strategy clearly represents a potent means of identifying drug transporters,
and notably it also illuminated cases of interactions. Such interactions scale exponentially
with the number of candidates, but high-throughput CRISPR-Cas methods are enabling
the identification of all kinds of genes involved in complex biological processes (e.g., [132
137]).
D
D
D
DEAD
Gene knockout -based strategy for
solute carrier identification
KO2
SURVIVES
D
LACKS CARRIER Y
KO1
KO3
KO4
Add toxic substrate t hat if transported kills ce lls
Molecules 2021, 26, 5629 7 of 39
1.5. Heterogeneity of Transport and Transporters in Different Cells and Tissues
If trans-bilayer permeability were significant in real biomembranes, the free concen-
trations of drugs inside cells and organelles (modulo pH and potential gradients) would
be more or less homogeneous. Of course, they are not, as is well-known (and as can easily
be determined by chemical imaging methods, e.g., [138159]). A particularly well-estab-
lished case is that of the bloodbrain barrier, where, despite the lipids not being noticeably
different from those in other cells, the permeability is negligible if specific transporters are
not involved [160164]. What does differ greatly between different cells and tissues, of
course, is the expression of particular proteins such as transporter proteins [85,86,165],
with any number of large datasets now becoming available.
1.6. Role of Transporters in Biotechnology
In addition to activities mentioned in our previous reviews that were focused on
transporters and biotechnology [4,12,166], a number of other authors (as reviewed, e.g., in
[167178]) have also highlighted the importance of transporters in the uptake of CO2 [179]
and in the biotechnological production of substances of interest in the BioEconomy. Some
recent examples include the production of amorphadiene [180], citrate [181], L-malate
[182184], antibiotics [185], fatty acids [186], fatty alcohols [187], olefins [188], various or-
ganic acids [176], other dicarboxylic acids [189,190], sophorolipids [173,186,191,192], in
microbial fluorination (by removing a fluoride effluxer from E. coli [193]), and in the pro-
duction of a variety of hydrophobic substances [191,194]. Promiscuity can be quite signif-
icant for biotechnology [182,195199]. In particular, here, the promiscuity of some trans-
porters, especially under biotechnological conditions of unphysiologically high intracel-
lular concentrations of small molecules, means that cells can have a tendency to leak path-
way intermediates with structural similarities (as judged by various means [200,201]) to
products, rather than simply just excreting the desired product; this too can be manipu-
lated via transporter engineering [171].
1.7. Adaptive Laboratory Evolution and Membrane Transporters
While some studies are more binary, looking for resistance to a toxic substrate via
survival or death “in one go, other strategies (such as the very elegant “variable dose
analysis[202]) are more graded. A particularly nice example is that of adaptive labora-
tory evolution (ALE), illustrated in Figure 5. While many very elegant examples of long-
term laboratory-based bacterial evolution exist (e.g., [203210]), a more directed focus has
been where it is exploited for biotechnology (e.g., [190,211223]). However, it is perfectly
applicable in drug transporter studies too. As with any phenotypic selection of this type
[224231], it relies on the fact that, in a heterogeneous population, the faster-growing
strains will tend to become more prevalent at the expense of the slower-growing ones.
Sequencing those that take over indicates where favourable mutational events have oc-
curred, and essentially identifies the relevant genes. It can best be run as a hypothesis-free
strategy [232], given that some mutations cannot reasonably be predicted (e.g., the role of
ribosomal subunits in improving methanol tolerance in E. coli [233]). In one example in-
volving transporters [234], ALE was used to increase the tolerance to aromatic amino acids
of baker’s yeast. Here, [234], the major mutation was in a transcriptional activator called
Aro80, that served to increase the activity of an efflux transporter Espb6, a role (in efflux-
ing aromatic compounds) shared with the previously known Pdr12. In another example
[235], E. coli cells were made more tolerant to ionic liquids; in this case, the chief mutations
in tolerant clones occurred in transport processes in the shape of mdtJI, a multidrug efflux
pump, and yhdP, a largely uncharacterised transporter possibly involved [236,237] in
phospholipid transport to the outer membrane.
Molecules 2021, 26, 5629 8 of 39
Figure 5. Illustration of the principle of adaptive laboratory evolution (ALE). Cells are exposed to a toxic substance that
causes them to grow sub-optimally. Selection leads to strains that can revert to rates and extents of growth shown by the
wild type when inoculated into fresh cultures. The stress level is increased and the process continued. Sampling and
growth rate measurement can be completely automated.
As illustrated in Figure 5, the commonest means of performing ALE is in a semi-
batchtype of mode in which small inocula are used to grow cells in batch mode, while
they are then sampled at the end of growth and a new inoculum introduced to a separate
batch culture. What is then selected, in part, is cells that as well as having a higher growth
rate also have lowered lag phases and the ability to survive better in stationary phase.
Truly continuous cultures provide for a much more stringent selection, especially when
carried out in a turbidostat. In a turbidostat (e.g., [227,238245]), the growth medium is
arranged such thatunlike in a chemostatcells can grow at the maximum rate they are
able to within it. The biomass in a fermentor of working volume V is controlled at a set
point via a suitable probe (see, e.g., [246248]). As it exceeds this level, fresh growth me-
dium (including any inhibitors) is pumped in at an average rate over a period of v mL ·
min−1. Cells are washed out at the same rate. As usual [249], this rate is numerically equiv-
alent to the growth rate = the dilution rate = v/V min−1, and may be recorded continuously.
Rather surprisingly, the method has not been widely used, although McGeachy et al. [227]
give a nice example that illustrated selection of mutations in the Mep3p ammonium trans-
porter. We predict that these kinds of strategies may have much more impact in the future.
Molecules 2021, 26, 5629 9 of 39
1.8. Substrate Misannotations and the Importance of Antiporter Activity in Drug Influx and
Efflux
Most transporters are discovered via their effects on the uptake or efflux of a partic-
ular substrate of interest, and they are often codified accordingly. Our experience is that—
just as with enzyme annotation more generally [250]this leads to misannotation in that
anyactivity discovered first is typically seen the main or even only activity. A classic
example is SLC22A4, previously known as OCTN1. It had been found to catalyse the up-
take of carnitine, and also that of the the non-physiological tetraethylammonium cation.
However, the rates were in fact quite miserable, and it was not until Gründemann and
colleagues used what was effectively an untargeted metabolomics approach [103] that it
was discovered that it was in fact reallya concentrative, sodium-coupled transporter
for the nutraceutical ergothioneine [251,252] and also for the related stachydrine (proline
betaine) [103,253,254]. Equivalently, this example also serves to illustrate how finding a
transporter with one activity does not mean that it is the only such capability, and it is
now known that at least one more transporter, viz SLC22A15 [109], can also catalyse er-
gothioneine uptake.
Another means of misannotation is that based on an assumption of unidirectionality
(as an influxeror an effluxer) that is not warranted and follows from the fact that
most kinetic assays are set up to measure either only an influx or an efflux. At one level,
bidirectional transport is inevitably the case, in that most equilibrative transporters are
necessarily functionally perfectly reversible; if a transporter has two uniported substrates
A and B, both can pass in either direction depending on their relative concentration on
either side of the membrane. More significantly, however, is the case in which the trans-
porter is in fact an antiporter, where the transport of A in one direction is coupled to the
transport of B in the other direction, whether B is measured or not. In radioisotopic assays
for the uptake of A, B is usually unlabelled and hence unobserved. By contrast, so-called
“untargeted” [255257] mass spectrometric methods show clearly that there is a massive
amount of efflux as well as influx when cells are exposed to new drug and nutrient sources
[258]. In addition, the energy coupling elements of concentrative or efflux transporters are
largely separate from those responsible for conducting the passage of the substrate
through the membrane or can be made so [259].
“Multi-drug transporters” (MDTs) represent a particularly clear example of this; of-
ten, a transporter is labelled as an MDT involved in efflux because lowering its activity
makes an organism more sensitive to a cytotoxic drug. However, in some cases (e.g.,
[260]), the removal of such a transporter actually makes the organism more sensitive to
some other substances! Only omicsmethods in which many substances are measured
simultaneously can easily disentangle this kind of behaviour.
1.9. Genome-Wide Analysis of Drug Uptake and Efflux in E. coli
In recent work [261263], we have recognised that the existence of genome-wide
knockout (and overexpression) collections allows for the high-throughput assessment of
the uptake of small molecule substrates. Since fluorophores are perfectly good surrogates
for these purposes [264], and their uptake admits easy assessment using flow cytometry
[265272], we have been able to assess the influence of the expression of individual trans-
porters on the uptake of various fluorophores. Figure 6 gives an example from E. coli [261],
using a wild type and some 530 knockouts, of the effect of such knockouts on the steady-
state uptake of two dyes, viz DiSC3(5), a carbocyanine dye responsive to membrane ener-
gisation [70], and SYBR Green, an intercalating dye whose fluorescence is massively en-
hanced upon binding to (especially double stranded [273]) DNA [274,275]. We here focus
on the so-called multidrug transporter genes of the mdt family. Although the effect of
knocking them out individually is, as expected [276], mainly (for mdt B,C,D,G,I, and J) to
increase the steady-state uptake of DiSC3(5), that of mdtH and mdtK has the opposite
effect, and uptake is very significantly decreased, indicating that normally these can act
Molecules 2021, 26, 5629 10 of 39
as influxers for this molecule. In a similar vein, the uptake of SYBR Green is lowered when
mdtF and mdtL are knocked out.
While one can never exclude the contribution of pleiotropic effects (e.g., [277]), the
obvious conclusion from these kinds of genome-wide study is that many individual trans-
porters can potentially contribute to the uptake and efflux of any stated substrate of inter-
est. Where this is the case, knocking a single one out may not have measurable effects,
since the others can take up the slack[276]; this does not of course then mean that the
gene knocked out did not catalyse the flux of the substrate of interestabsence of evi-
dence is not evidence of absence.
Figure 6. Dual roles of so-called multidrug transport proteins in E. coli. Data redrawn from those published with Open
Access at [261]. Plotted are the median uptakes of SYBR Green and DiSC3(5) by various knockout strains relative to the
Wild Type (WT). The ranges are, respectively, 70-fold and 36-fold. So-called y-genes (genes of nominally unknown func-
tion [278,279]) are encoded in red. Mdt gene knockouts are labelled (and have green symbols).
A convenient method for invoking genome-wide diversity via gene disruption is the
use of transposon insertion, but this has the disadvantage that the inactivation of essential
genes in haploid organisms is missed. Thus, Webber and colleagues [280] have developed
a system (TraDIS-Xpress) based on a transposon linked to an inducible promoter and used
this to determine transporters (and other genes) involved in resistance to triclosan [280]
and to fosfomycin [281]. This kind of strategy will be vital in further uncovering trans-
porter-mediated contributions to antimicrobial resistance (AMR).
Uptake of two dyes in E. coli gene knockouts
Median DiSC3(5) uptake
Median SYBR Green uptake
Molecules 2021, 26, 5629 11 of 39
1.10. Recent Approaches to the De-Orphanisation of Mammalian OrphanTransporters
Some 1000 genes in the human genome encode transporters [7], of which the largest
class, amounting to roughly half [282288], is represented by the SoLute Carrier (SLC)
class (http://slc.bioparadigms.org/, accessed on 15 September 2021). As noted above, even
for those that have at least one known substrate, there are doubtless many more to be
found. More significantly, at this stage of knowledge, many of them are complete or-
phansin that not a single substrate is in fact known. One recent example of de-orphani-
sation is that of the mitochondrial transporter SLC25A51, which turns out [108,289] to be
an NAD+ importer. This de-orphanisation hinged upon a successful combination of ge-
nomics, metabolomics, CRISPR-Cas9-mediated gene editing, and genetics. Other strate-
gies include binding assays [107], the use of direct assays in Xenopus oocytes [290], and
others that are covered in a recent and comprehensive review [119].
1.11. Selectivity and Drug Targeting by the Use and Exploitation of Varying Expression Profiles
Given the need for transporters if drugs are to cross cell membranes, one can make a
virtue of necessity [284] by seeking either to exploit their natural expression profiles [291
293] or to vary them explicitly [294], so as to target them to particular tissues [295]. The
latter has obvious benefits in oncology [296300], where the necessary cytotoxicity of
many drugs, such as nucleoside analogues, is rather non-specific. The strategy can be used
for modifying both influx and efflux transporters, though the latter is likely to prove more
efficacious [276]. In one example [294] (Figure 7), the second drug in the binary weapon
strategy decreased the expression of a relevant efflux transporter some 12-fold, in a cell-
selective manner. This binary weapon strategy potentially holds much promise for target-
ing cytotoxic anti-cancer drugs.
Figure 7. Binary weaponsbased on drug transporters [294]. The cytotoxic drug gemcitabine
(GEM, a fluorinated cytosine nucleoside analogue) when added at a certain low concentration to
Panc1 pancreatic cancer cells was barely cytotoxic (left panel). However, when a second drug was
added, which was itself also non-toxic, the combination was substantially more toxic. What had
occurred was that, in response to the GEM, the cells had increased the expression of the efflux trans-
porter ABCC2 (MRP2) some 12-fold; the second drug inhibited this process, and in a cell-selective
manner.
The fact that expression levels of SLCs often vary considerably was illustrated by us
previously [86], using a publicly available dataset [85]. The Gini coefficient (see
[86,100,301,302]) (Figure 8) describes the heterogeneity of a distribution in a simple, non-
Drug 2 varies efflux
transport er
expression
increased toxicity
Drug 2
Drug 2
Proposed partial mode of action of binary weapons in
enhancing gemcitabine toxicity in Panc1 cells
Panc1
GEM
GEM
Nucleoside
influx
transporter
GEM GEM
Influx/efflux transporter
expression ratio
insufficient for full toxicity
Dr ug 2
transporter
Nucleoside
efflux
transporters
GEM
GEM
GEM
Molecules 2021, 26, 5629 12 of 39
parametric manner, and takes values between zero and 1. Some of the cell line data are
replotted in Figure 9, where the Gini coefficient is plotted against their median expression
level as assessed by RNASeq. Specifically, we use these data to illustrate four points: (i)
the very fact that there are a great many uptake transporters (at least 400), that may or
may not be expressed in different cells; (ii) those transporters with the lowest Gini coeffi-
cient, indicating their relative homogeneity of expression between cell lines (which may
be referred to as GiniGenes [86,100]); (iii) an example (in the form of SLC18A2, a vesicular
monoamine transporter [303]) of a transporter with both a high Gini coefficient, indicating
a very restricted expression, and a reasonably high expression level; and (iv) the substan-
tial variation in Gini coefficient for the six members of the SLC35F family, with three being
quite selectively expressed (SLC35F1,3,4) while the others (SLC35F2,5,6) are expressed
over a very broad range of values, as shown explicitly elsewhere for SLC35F2 [97]. Over-
all, 61 of the 410 transporters (15%) plotted in Figure 9 have a Gini coefficient over 0.8, 105
(26%) exceed 0.7 and 173 (42%) exceed 0.5. By contrast, the Gini coefficient for income
inequality in different countries (its usual domain of application [304]) https://www.in-
dexmundi.com/facts/indicators/SI.POV.GINI/rankings shows only 14 countries/159 (9%)
with a Gini coefficient exceeding 0.5.
Figure 8. Illustration of the derivation of the non-parametric Gini coefficient for describing the inequality of a distribution
(here the variation of transcript levels between cell lines). This was achieved by rank ordering the value of the different
examples according to their expression levels, as indicated.
Derivation of the Gini coefficient (index)
and its application to absolute
transcriptomics
Cumul a�ve frac�on of total
transcripts when ordered from
lowest to highest expression levels
01
Cumula�ve frac�on of transcript
types in order of increasing
expression magnitude
Area A
Area B
G = A/(A + B)
1
Molecules 2021, 26, 5629 13 of 39
Figure 9. Variation of Gini coefficient and median expression for 410 SLC transporters in 56 cell lines. Data were obtained
from previous open access publications [85,86] and a subset replotted. A few transporters are labelled to illustrate some
of the SLCs with the lowest value of the Gini coefficient (lower right), one with a high value and a reasonable expression
(SLC18A2), and the six members of the SLC35F family. The transcript expression levels are expressed through a wide-
spread normalisation (see [305], but also [306,307]) as reads per kilobase million (RPKM).
1.12. Transporters and Prodrugs
Another strategy that has been widely used to improve cellular drug uptake is the
use of so-called prodrugs, in which a drug is modified by the addition of a moiety that,
although without direct pharmacological activity at a target receptor, assists the passage
of the drug to its target (e.g., [308318]). These necessarily involve transporters, although
their identity is not always known. Several recent examples showing high efficacy are
based on the SLC7A5 (LAT1) transporter [319324]. In one recent instance [321], the up-
take of the anti-inflammatory salicylic acid was enhanced more than five-fold by fusing it
with a phenylalanine moiety. Sometimes the prodrug is more lipophilic than its parent,
and this is taken in some quarters to mean that it therefore must be passing through bi-
layers. As pointed out before [8], however, “in actual practice, the reformulation of a water
soluble drug with lipidization modifications is difficult to execute successfully, and there
is not a single example of a drug presently sold whereby medicinal chemistry was suc-
cessfully used to convert a non-brain-penetrating drug into a molecule that crosses the
BBB {bloodbrain barrier} in pharmacologically significant amounts[325].
Variation of expression of SLC transporters in cell lines
Gini coefficient
Expression level / RPKM
Molecules 2021, 26, 5629 14 of 39
1.13. Transporters and Adverse Drug Reactions
Notwithstanding all the benefits of small molecule drugs, it remains the case that
they are also widely associated with various morbidities and mortalities, often referred to
as adverse drug reactions(ADRs) (e.g., [326346]). Despite the lengthy and complex
regulatory hurdles that drugs must overcome before being marketed (and toxicity re-
mains a major cause of so-called “attritionwhere drug candidates are pulled before they
even get to market [347353]), ADRs are extremely common (and hard to anticipate, given
the huge genetic and phenotypic variation in human populations [354]). Due to non-line-
arities in biochemical kinetics, averaging across many cells or tissues necessarily hides the
true biology [267,355], and thus lumping heterogeneous cells into tissues will always miss
such problems [2]. Our contention is that in many cases these ADRs are mediated via
transporters, especially since when concentrative transporters can potentially cause mas-
sive accumulation in particular cells. Failure or inhibition of efflux transporters also has a
major role to play in drug toxicity.
Consequently, the involvement of transporters in ADRs and drug cytotoxicity is
well-established (see above and, e.g., [294,336,339,340,356394]), providing further evi-
dence for the major roles of transporters in both pharmacokinetics and pharmacodynam-
ics.
1.14. Transporters, Antibiotics, and Antimicrobial Resistance (AMR)
Antimicrobial resistance (AMR) is a major human health problem (e.g., [272,395
419]). Most pertinently, efflux transporters are well-recognised as a major source of AMR
(e.g., [276,397,420457]). The outer membrane also contributes significantly to the perme-
ability barrier in Gram-negatives (e.g., [265,434,443,444,448,458466]). However, another
specific area in which the role of transporters is largely unrecognised—albeit this is a spe-
cific subset of drug transportpertains to the uptake transport of anti-infectives to their
sites of action [439,467]. This can involve both the targets within the microbe and the host’s
transporters when (as is common, e.g., [468485]) the infective agents reside intracellu-
larly. A particularly clear example is given by Mycobacterium tuberculosis, the causative
agent of TB, where the very striking lack of correlation between in vivo and in vitro drug
potencies is easily and necessarily explained via transporter activities [139,151,486488].
Many orally prescribed antibiotics enter the host via SLC15 family members [489,490],
while some of the relatively few known microbial uptake transporters for anti-infectives
are listed in Table 1. What evidence there is implies that there are multiple means of up-
take, which is why identifying individual transporters for successful antibiotics has
proven difficult [276]. On the flipside, of course, when we recognise the relevant trans-
porters and/or their structureactivity relationships governing cell permeability, we can
exploit them [437,444,491495].
Molecules 2021, 26, 5629 15 of 39
Table 1. Some examples of differential resistance to anti-microbial drugs involving uptake trans-
porters.
Antibiotic
Transporter
Comments
Selected Reference(s)
Aminoglycosides
[496]
Chloramphenicol YdgR
E. coli. Proton-de-
pendent oligopeptide
transporter analogue
[497]
Cycloserine
[498]
5-fluocytosine
FCY2
Various Candida spp.
[499,500]
Fosfomycin
[501,502]
Pacidamycin
Opp
PA14
Pseudomonas aeru-
ginosa
[503]
[504]
Pentamidine
Three adenosine-
based transporters
[441,505]
Quinoline antimalari-
als
AAT1 [506]
Reviews
[441,507509]
Tetracyclines
Two unknown trans-
porters
[510,511]
1.15. Molecular Dynamics of Transporter Reactions
Given, as mentioned, the present impossibility of detecting the molecular pathway
of drug transport directly, one alternative is to calculate it from first principles, which for
these purposes means via the use of molecular dynamics (MD). MD allows the calculation
ab initioof the molecular motions of molecules during their normal activity. Although
computationally demanding (something that becomes much less of an issue over time,
e.g., [512,513]), it is perfectly suited to calculating the mechanisms of substrates transport
across membranes [514]. We have discussed this in more detail elsewhere [2], and so we
simply include here some of the more recent developments. Thus, Jia et al. [515] could
mirror precisely the experimental findings underpinning the behaviours of a xylose trans-
porter. Other findings uncovered an electrostatic lock in emrE [516] (see also [517,518]),
established the molecular basis of sodium-coupled transporters [519], and illustrated the
mechanism of transporters as varied as acrB [520522], the vitamin B12 importer BtuCD
[523], the maltose transporter ATPase MalK2 [524], McjD (an antibacterial peptide ABC
transporter from E. coli) [525], proton oligopeptide transporters [526], hexameric urea
transporter UreI from Helicobacter pylori [527], and the mammalian transporters SLC1A1
(excitatory amino acid transporter EAAT3) [528], SLC2A1 (glucose transporter GLUT1)
[529], SLC4A1 (band 3” protein, bicarbonate/anion exchanger) [530], SLC6A4 [531],
SLC7A10 [532]. By contrast, studies of membrane-embedded molecules such as aqua-
porins, when conducted with high protein concentrations resembling those in biomem-
branes [533536], show that even water molecules do not pass through the bilayer (see a
direct illustration at http://www3.mpibpc.mpg.de/groups/de_groot/gallery/aqp1_snap-
shot.jpg/). In several cases, the MD simulations are accompanied by confirmatory X-ray
structures (e.g., [530,535,536]).
1.16. Uptake Transporters as Drug Targets
Although our chief interest here relates mainly to the role of solute carriers in drug
disposition, we would be remiss not to mention that, largely because they have been seri-
ously understudied [288], SLCs themselves necessarily constitute potentially valuable and
novel drug targets [11,286,537552]. It is reasonable that the technical improvements in
cryo-EM will contribute to the rational design of such drugs [553], as well as the many
other activities ongoing (e.g., as summarised in [131]).
Molecules 2021, 26, 5629 16 of 39
1.17. What Are the Real” Substrates of Drug Uptake Transporters?
While some drugs that are SLC substrates are, or are semi-synthetic analogues of,
natural products, most modern drugs are entirely synthetic in nature, and so natural evo-
lution had no known exposure to them. It is then at least reasonable to enquire as to what
the normalsubstrates of these molecules are that happen also to allow them to transport
drugs. The principle of molecular similarity (e.g., [201,554564]) suggests that molecules
that have similar structures should tend to have similar activities, so the question then
becomes to which molecules are marketed drugs most similar? This is a cheminformat-
ics question, and the answer depends in part on the nature of the structural encoding,
although most encodings of actuallysimilar molecules show a Tanimoto similarity ex-
ceeding 0.8 or so, a number that may be used as a kind of benchmark. Our initial assump-
tion was that successful, marketed drug should bear structural similarities to endogenous
human metabolites [565568], which have been catalogued in metabolic reconstructions
[569] and elsewhere (e.g., [570]). However, this accounts for only a small percentage (~15%
[83,200,568]), and the true answer—possibly unsurprisingly, post hocis that most drugs
actually bear similarities to natural products, whose uptake via SLCs may be assumed to
be, or to have been during natural selection, of some nutritional or medical benefit to the
host [200,264]. This also brings to the fore the important role of natural products in drug
discovery [9,571586]. Natural products are, of course, famous for breaking [587590]
many of the rule of 5” [591] guidelines, and include many very clear examples of drugs
that cannot possibly diffuse through lipid bilayers in intact biological cells.
1.18. Why Do SOME Solvents Increase the Rate of Drug Uptake?
Anecdotally, there is a common assumption that because solvents such as DMSO can
increase the rate of drug uptake they must be doing so by solubilising the drug, or at least
by assisting its solubilisation, in the phospholipid bilayer part of biomembranes. This
would then be seen as some kind of evidence for the importance of bilayer transport, but
in fact this does not follow at all. The issue with insoluble drugs is that transporters require
their substrates to be bound, and that this normally happens via solubilisation in the aque-
ous phase. Rocks and crystals and amorphous solids are not direct substrates of drug
transporters; molecules are. All that solvents such as DMSO are then doing is increasing
the rate of solubilisation of drug solids and their presentation to the transporters as single
molecules (the necessarily preferred substrates) in solution.
2. Discussion
Real biological membranes possess a high protein:lipid ratio, often as much as 3:1 by
mass. Consequently, they do not remotely behave in a manner similar to pure, artificial,
phospholipid bilayers. Additionally, the transport of small organic molecules (including
drugs) across them commonly requires the intercession of proteinaceous transporters
(e.g., [2,3,6,8,101,287,370,592620]), and there is in fact no actual evidence whatsoever for
any significant flux across native, undamaged biomembranes through whatever phospho-
lipid bilayer may be present. This contrasts with a widely held set of assumptions, based
in part (it is assumed) on what can be observed in pure phospholipid bilayers that admit
the transport of all kinds of small molecules, albeit through transient aqueous pores.
Identifying these transporters of small molecules can be performed (e.g.,
[96,97,99,101,171]) by manipulating their expression, including under conditions in which
their substrates are otherwise toxic. The idea is that removing (or otherwise inhibiting) a
transporter that normally helps a toxic drug to enter a cell should increase the hosts re-
sistance to it, while overexpressing the transporter would make the cells more sensitive.
More generally, understanding the co-variation between the uptake of a molecule and the
expression of its potential transporters gives a strong indication of which they are.
Molecules 2021, 26, 5629 17 of 39
The recognition that in real biological membranes the transbilayer transport through
phospholipids of small molecule drugs, nutrients, and biotechnology products is negligi-
ble also explains straightforwardly the following well-established facts:
The negligible uptake of drugs and substrates in some tissues, including via the
bloodbrain barrier (and equivalents in the retina, testes, and other tissues), where
relevant transporters are absent;
The extreme heterogeneity of uptake of a given molecule in different organs, tissues,
and organisms despite little substantive variation in their lipid physical properties;
The existence of transporters for all kinds of small molecules (even water, acetate,
ammonia, glycerol, etc., as well as entirely hydrophobic molecules such as alkanes
[621623]) that had previously been assumed to lack them;
A variety of cases in which individual defined transporters can be shown to account
for the overwhelming bulk of measured fluxes;
The need for such transporters in order to effect drug uptake, mirroring the wide-
spread recognition that they can serve to efflux them (and thereby created resistance
to their activity);
The role of transporters in drug-mediated toxicity (e.g., [366,624]);
The poor correlation between the uptake of small molecules and simple physico-
chemical properties such as log P or log D (many examples, such as those in
[2,6,83,625627]).
3. Looking to the Future
The two chief questions posed earlier (what are the substrates for a given tra-
sporter?” and what are the transporters for a given substrate?”) are normally addressed
experimentally, using a variety of the methods described above. As intimated above, the
exponential increase in computer power will eventually allow the methods of molecular
dynamics to admit these “measurements” entirely by calculations based on simple force
fields, de novo. In addition, given the success of so-called deep learning [628630] methods
in predicting the structures of proteins [631645], novel receptorligand interactions [646
650], and a variety of other protein and small molecule properties (e.g., [201,651665]), it
seems likely that we shall soon have available in silico methods for predicting transporter
substrates directly from protein sequences, and the likeliest transporters from candidate
substrate structures of interest.
4. Conclusions
On the basis of the present evidence, the transport of drugs and nutrients through
phospholipid bilayers in real biomembranes is negligible. Progress in understanding drug
distribution profiles is thus to be made by establishing which transporters they use, the
expression profiles of those transporters, and the kinetic rate equations that they obey.
Armed with these it will be possible to model, to analyse, to understand, and to exploit
our principled knowledge of drug distributions both mechanistically and with confidence.
Funding: DBK thanks the BBSRC (grant BB/R014426/1) and the Novo Nordisk Foundation (Grant
agreement No. NNF20CC0035580) for financial support.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: I thank many colleagues for useful discussions.
Conflicts of Interest: The author declares no conflict of interest.
Sample Availability: Samples of the compounds are not available from the authors.
Molecules 2021, 26, 5629 18 of 39
References
1. Kell, D.B. How drugs pass through biological cell membranesA paradigm shift in our understanding? Beilstein Mag. 2016, 2,
Available online: http://www.beilstein-institut.de/download/628/09_kell.pdf (accessed on 15/9/21).
2. Kell, D.B.; Oliver, S.G. How drugs get into cells: Tested and testable predictions to help discriminate between transporter-
mediated uptake and lipoidal bilayer diffusion. Front. Pharmacol. 2014, 5, 231.
3. Dobson, P.; Lanthaler, K.; Oliver, S.G.; Kell, D.B. Implications of the dominant role of cellular transporters in drug uptake. Curr.
Top. Med. Chem. 2009, 9, 163–184.
4. Kell, D.B.; Swainston, N.; Pir, P.; Oliver, S.G. Membrane transporter engineering in industrial biotechnology and whole-cell
biocatalysis. Trends Biotechnol. 2015, 33, 237–246.
5. Mendes, P.; Oliver, S.G.; Kell, D.B. Fitting transporter activities to cellular drug concentrations and fluxes: Why the bumblebee
can fly. Trends Pharmacol. Sci. 2015, 36, 710–723.
6. Dobson, P.D.; Kell, D.B. Carrier-mediated cellular uptake of pharmaceutical drugs: An exception or the rule? Nat. Rev .Drug
Discov. 2008, 7, 205–220.
7. Kell, D.B.; Dobson, P.D.; Oliver, S.G. Pharmaceutical drug transport: The issues and the implications that it is essentially carrier-
mediated only. Drug Discov. Today 2011, 16, 704–714.
8. Kell, D.B.; Dobson, P.D.; Bilsland, E.; Oliver, S.G. The promiscuous binding of pharmaceutical drugs and their transporter-
mediated uptake into cells: What we (need to) know and how we can do so. Drug Discov. Today 2013, 18, 218–239.
9. Kell, D.B. Finding novel pharmaceuticals in the systems biology era using multiple effective drug targets, phenotypic screening,
and knowledge of transporters: Where drug discovery went wrong and how to fix it. FEBS J. 2013, 280, 5957–5980.
10. Kell, D.B. Implications of endogenous roles of transporters for drug discovery: Hitchhiking and metabolite-likeness. Nat. Rev.
Drug Discov. 2016, 15, 143–144.
11. Kell, D.B.; Wright Muelas, M.; O’Hagan, S.; Day, P.J. The role of drug transporters in phenotypic screening. Drug Target Rev.
2018, 4, 16–19.
12. Kell, D.B. Control of metabolite efflux in microbial cell factories: Current advances and future prospects. In Fermentation Micro-
biology and Biotechnology, 4th ed.; El-Mansi, E.M.T., Nielsen, J., Mousdale, D., Allman, T., Carlson, R., Eds.; CRC Press: Boca
Raton, FL, USA, 2019; pp. 117–138.
13. Singer, S.J.; Nicolson, G.L. The fluid mosaic model of the structure of cell membranes. Science 1972, 175, 720–731.
14. Nicolson, G.L. The Fluid-Mosaic Model of Membrane Structure: Still relevant to understanding the structure, function and
dynamics of biological membranes after more than 40 years. Biochim. Biophys. Acta 2014, 1838, 1451–1466.
15. Goñi, F.M. The basic structure and dynamics of cell membranes: An update of the Singer-Nicolson model. Biochim. Biophys. Acta
2014, 1838, 1467–1476.
16. Guidotti, G. The composition of biological membranes. Arch. Intern. Med. 1972, 129, 194–201.
17. Engelman, D.M. Membranes are more mosaic than fluid. Nature 2005, 438, 578–580.
18. Dupuy, A.D.; Engelman, D.M. Protein area occupancy at the center of the red blood cell membrane. Proc. Natl. Acad. Sci. USA
2008, 105, 2848–2852.
19. Domański, J.; Marrink, S.J.; Schäfer, L.V. Transmembrane helices can induce domain formation in crowded model membranes.
Biochim. Biophys. Acta 2012, 1818, 984–994.
20. Lindén, M.; Sens, P.; Phillips, R. Entropic tension in crowded membranes. PLoS Comput. Biol. 2012, 8, e1002431.
21. Goose, J.E.; Sansom, M.S.P. Reduced Lateral Mobility of Lipids and Proteins in Crowded Membranes. PLoS Comput. Biol 2013,
9, e1003033.
22. Höfling, F.; Franosch, T. Anomalous transport in the crowded world of biological cells. Rep. Prog. Phys. 2013, 76, 046642.
23. Javanainen, M.; Hammaren, H.; Monticelli, L.; Jeon, J.H.; Miettinen, M.S.; Martinez-Seara, H.; Metzler, R.; Vattulainen, I. Anom-
alous and normal diffusion of proteins and lipids in crowded lipid membranes. Faraday Discuss. 2013, 161, 397–417.
24. Guigas, G.; Weiss, M. Effects of protein crowding on membrane systems. Biochim Biophys Acta 2016, 1858, 2441–2450.
25. Jeon, J.H.; Javanainen, M.; Martinez-Seara, H.; Metzler, R.; Vattulainen, I. Protein Crowding in Lipid Bilayers Gives Rise to Non-
Gaussian Anomalous Lateral Diffusion of Phospholipids and Proteins. Phys. Rev. X 2016, 6, 021006.
26. Duncan, A.L.; Reddy, T.; Koldso, H.; Helie, J.; Fowler, P.W.; Chavent, M.; Sansom, M.S.P. Protein crowding and lipid complexity
influence the nanoscale dynamic organization of ion channels in cell membranes. Sci. Rep. 2017, 7, 16647.
27. Marrink, S.J.; Corradi, V.; Souza, P.C.T.; Ingólfsson, H.I.; Tieleman, D.P.; Sansom, M.S.P. Computational Modeling of Realistic
Cell Membranes. Chem. Rev. 2019, 119, 6184–6226.
28. Ando, T.; Uchihashi, T.; Scheuring, S. Filming biomolecular processes by high-speed atomic force microscopy. Chem. Rev. 2014,
114, 3120–3188.
29. Jain, M.K. The Bimolecular Lipid Membrane; Van Nostrand Reinhold: New York, NY, USA, 1972.
30. Tien, H.T. Bilayer Lipid Membranes (BLM): Theory and Practice; Marcel Dekker: New York, NY, USA, 1974.
31. Tien, H.T.; Ottova-Leitmannova, A. Planar Lipid Bilayers (BLMs) and Their Applications; Elsevier: New York, NY, USA, 2003.
32. Marrink, S.J.; Jähnig, F.; Berendsen, H.J.C. Proton transport across transient single-file water pores in a lipid membrane studied
by molecular dynamics simulations. Biophys. J. 1996, 71, 632–647.
33. Weaver, J.C.; Powell, K.T.; Mintzer, R.A.; Sloan, S.R.; Ling, H. The diffusive permeability of bilayer membranes: The contribu-
tion of transient aqueous pores. Bioelectrochem. Bioenerg. 1984, 12, 405–412.
34. Deamer, D.W.; Bramhall, J. Permeability of lipid bilayers to water and ionic solutes. Chem. Phys. Lipids 1986, 40, 167–188.
Molecules 2021, 26, 5629 19 of 39
35. Leontiadou, H.; Mark, A.E.; Marrink, S.J. Molecular dynamics simulations of hydrophilic pores in lipid bilayers. Biophys. J. 2004,
86, 2156–2164.
36. Loison, C.; Mareschal, M.; Schmid, F. Pores in bilayer membranes of amphiphilic molecules: Coarse-grained molecular dynam-
ics simulations compared with simple mesoscopic models. J. Chem. Phys. 2004, 121, 1890–1900.
37. Leontiadou, H.; Mark, A.E.; Marrink, S.J. Ion transport across transmembrane pores. Biophys. J. 2007, 92, 4209–4215.
38. Marrink, S.J.; de Vries, A.H.; Tieleman, D.P. Lipids on the move: Simulations of membrane pores, domains, stalks and curves.
Biochim. Biophys. Acta 2009, 1788, 149–168.
39. Gurtovenko, A.A.; Anwar, J.; Vattulainen, I. Defect-mediated trafficking across cell membranes: Insights from in silico model-
ing. Chem. Rev. 2010, 110, 6077–6103.
40. Bennett, W.F.; Tieleman, D.P. The importance of membrane defects-lessons from simulations. Acc. Chem. Res. 2014, 47, 22442251.
41. Bubnis, G.; Grubmüller, H. Sequential Water and Headgroup Merger: Membrane Poration Paths and Energetics from MD Sim-
ulations. Biophys. J. 2020, 119, 2418–2430.
42. Akimov, S.A.; Volynsky, P.E.; Galimzyanov, T.R.; Kuzmin, P.I.; Pavlov, K.V.; Batishchev, O.V. Pore formation in lipid membrane I:
Continuous reversible trajectory from intact bilayer through hydrophobic defect to transversal pore. Sci. Rep. 2017, 7, 12152.
43. Akimov, S.A.; Volynsky, P.E.; Galimzyanov, T.R.; Kuzmin, P.I.; Pavlov, K.V.; Batishchev, O.V. Pore formation in lipid mem-
brane II: Energy landscape under external stress. Sci. Rep. 2017, 7, 12509.
44. Parsegian, A. Energy of an ion crossing a low dielectric membrane: Solutions to four relevant electrostatic problems. Nature
1969, 221, 844–846.
45. Lee, A.G. Lipid-protein interactions in biological membranes: A structural perspective. Biochim. Biophys. Acta 2003, 1612, 1–40.
46. Tillman, T.S.; Cascio, M. Effects of membrane lipids on ion channel structure and function. Cell Biochem. Biophys. 2003, 38, 161–190.
47. Van Meer, G.; Voelker, D.R.; Feigenson, G.W. Membrane lipids: Where they are and how they behave. Nat. Rev. Mol. Cell Biol.
2008, 9, 112–124.
48. Niemelä, P.S.; Miettinen, M.S.; Monticelli, L.; Hammaren, H.; Bjelkmar, P.; Murtola, T.; Lindahl, E.; Vattulainen, I. Membrane
proteins diffuse as dynamic complexes with lipids. J. Am. Chem. Soc. 2010, 132, 7574–7575.
49. Hickey, K.D.; Buhr, M.M. Lipid bilayer composition affects transmembrane protein orientation and function. J. Lipids 2011, 2011,
208457.
50. Laganowsky, A.; Reading, E.; Allison, T.M.; Ulmschneider, M.B.; Degiacomi, M.T.; Baldwin, A.J.; Robinson, C.V. Membrane
proteins bind lipids selectively to modulate their structure and function. Nature 2014, 510, 172–175.
51. Aponte-Santamaría, C.; Briones, R.; Schenk, A.D.; Walz, T.; de Groot, B.L. Molecular driving forces defining lipid positions
around aquaporin-0. Proc. Natl. Acad. Sci. USA 2012, 109, 9887–9892.
52. Poveda, J.A.; Giudici, A.M.; Renart, M.L.; Molina, M.L.; Montoya, E.; Fernández-Carvajal, A.; Fernández-Ballester, G.; Encinar, J.A.;
González-Ros, J.M. Lipid modulation of ion channels through specific binding sites. Biochim. Biophys. Acta 2014, 1838, 15601567.
53. Hedger, G.; Sansom, M.S.P. Lipid interaction sites on channels, transporters and receptors: Recent insights from molecular
dynamics simulations. Biochim. Biophys. Acta 2016, 1858, 2390–2400.
54. Kalli, A.C.; Sansom, M.S.P.; Reithmeier, R.A.F. Molecular dynamics simulations of the bacterial UraA H+-uracil symporter in
lipid bilayers reveal a closed state and a selective interaction with cardiolipin. PLoS Comput. Biol. 2015, 11, e1004123.
55. Newport, T.D.; Sansom, M.S.P.; Stansfeld, P.J. The MemProtMD database: A resource for membrane-embedded protein struc-
tures and their lipid interactions. Nucleic Acids Res. 2019, 47, D390–D397.
56. Song, C.; de Groot, B.L.; Sansom, M.S.P. Lipid Bilayer Composition Influences the Activity of the Antimicrobial Peptide Derm-
cidin Channel. Biophys. J. 2019, 116, 1658–1666.
57. Gault, J.; Liko, I.; Landreh, M.; Shutin, D.; Bolla, J.R.; Jefferies, D.; Agasid, M.; Yen, H.Y.; Ladds, M.; Lane, D.P.; et al. Combining native
and ‘omics’ mass spectrometry to identify endogenous ligands bound to membrane proteins. Nat. Methods 2020, 17, 505508.
58. Harayama, T.; Riezman, H. Understanding the diversity of membrane lipid composition. Nat. Rev. Mol. Cell Biol. 2018, 19, 281296.
59. Casares, D.; Escribá, P.V.; Rosselló, C.A. Membrane Lipid Composition: Effect on Membrane and Organelle Structure, Function
and Compartmentalization and Therapeutic Avenues. Int. J. Mol. Sci. 2019, 20, 2167.
60. Kanonenberg, K.; Royes, J.; Kedrov, A.; Poschmann, G.; Angius, F.; Solgadi, A.; Spitz, O.; Kleinschrodt, D.; Stühler, K.; Miroux, B.; et
al. Shaping the lipid composition of bacterial membranes for membrane protein production. Microb. Cell Factories 2019, 18, 131.
61. Chorev, D.S.; Robinson, C.V. The importance of the membrane for biophysical measurements. Nat. Chem. Biol. 2020, 16, 12851292.
62. Thompson, M.J.; Baenziger, J.E. Ion channels as lipid sensors: From structures to mechanisms. Nat. Chem. Biol. 2020, 16, 13311342.
63. Doktorova, M.; Symons, J.L.; Levental, I. Structural and functional consequences of reversible lipid asymmetry in living mem-
branes. Nat. Chem. Biol. 2020, 16, 1321–1330.
64. Lorent, J.H.; Levental, K.R.; Ganesan, L.; Rivera-Longsworth, G.; Sezgin, E.; Doktorova, M.; Lyman, E.; Levental, I. Plasma
membranes are asymmetric in lipid unsaturation, packing and protein shape. Nat. Chem. Biol. 2020, 16, 644–652.
65. Makarova, M.; Owen, D.M. Asymmetry across the membrane. Nat. Chem. Biol. 2020, 16, 605–606.
66. Pohl, A.; Devaux, P.F.; Herrmann, A. Function of prokaryotic and eukaryotic ABC proteins in lipid transport. Biochim. Biophys.
Acta 2005, 1733, 29–52.
67. Quazi, F.; Molday, R.S. Lipid transport by mammalian ABC proteins. Essays Biochem. 2011, 50, 265–290.
68. Borst, P.; Zelcer, N.; van Helvoort, A. ABC transporters in lipid transport. Biochim. Biophys. Acta 2000, 1486, 128–44.
69. Neumann, J.; Rose-Sperling, D.; Hellmich, U.A. Diverse relations between ABC transporters and lipids: An overview. Biochim.
Biophys. Acta Biomembr. 2017, 1859, 605–618.
Molecules 2021, 26, 5629 20 of 39
70. Kell, D.B. A protet-based, protonic charge transfer model of energy coupling in oxidative and photosynthetic phosphorylation.
Adv. Microb. Physiol. 2021, 78, 1–177.
71. Kell, D.B.; Welch, G.R. Belief: The Baggage Behind our Being; OSF Preprints: 2018. Available online: pnxcs https://osf.io/pnxcs/
(accessed on 15/9/21).
72. Kahneman, D. Thinking, Fast and Slow; Penguin: London, UK, 2011.
73. Sharot, T. The Optimism Bias; Robinson: London, UK, 2012.
74. Sharot, T. The Influential Mind; Abacus: London, UK, 2017.
75. Sharot, T. To quell misinformation, use carrotsNot just sticks. Nature 2021, 591, 347.
76. Peterson, J.C.; Bourgin, D.D.; Agrawal, M.; Reichman, D.; Griffiths, T.L. Using large-scale experiments and machine learning to
discover theories of human decision-making. Science 2021, 372, 1209–1214.
77. Artursson, P.; Palm, K.; Luthman, K. Caco-2 monolayers in experimental and theoretical predictions of drug transport. Adv.
Drug Deliv. Rev 1996, 22, 6784.
78. Sun, H.; Chow, E.C.; Liu, S.; Du, Y.; Pang, K.S. The Caco-2 cell monolayer: Usefulness and limitations. Expert Opin. Drug Metab.
Toxicol. 2008, 4, 395–411.
79. Van Breemen, R.B.; Li, Y. Caco-2 cell permeability assays to measure drug absorption. Expert Opin. Drug Metab. Toxicol. 2005, 1,
175–185.
80. Press, B. Optimization of the Caco-2 permeability assay to screen drug compounds for intestinal absorption and efflux. Methods
Mol. Biol. 2011, 763, 139–154.
81. Volpe, D.A. Drug-permeability and transporter assays in Caco-2 and MDCK cell lines. Future Med. Chem. 2011, 3, 2063–2077.
82. Mukhopadhya, I.; Murray, G.I.; Berry, S.; Thomson, J.; Frank, B.; Gwozdz, G.; Ekeruche-Makinde, J.; Shattock, R.; Kelly, C.;
Iannelli, F.; et al. Drug transporter gene expression in human colorectal tissue and cell lines: Modulation with antiretrovirals
for microbicide optimization. J. Antimicrob. Chemother. 2016, 71, 372–386.
83. O’Hagan, S.; Kell, D.B. The apparent permeabilities of Caco-2 cells to marketed drugs: Magnitude, and independence from both
biophysical properties and endogenite similarities PeerJ 2015, 3, e1405.
84. Iftikhar, M.; Iftikhar, A.; Zhang, H.; Gong, L.; Wang, J. Transport, metabolism and remedial potential of functional food extracts
(FFEs) in Caco-2 cells monolayer: A review. Food Res. Int. 2020, 136, 109240.
85. Thul, P.J.; Åkesson, L.; Wiking, M.; Mahdessian, D.; Geladaki, A.; Ait Blal, H.; Alm, T.; Asplund, A.; Björk, L.; Breckels, L.M.; et
al. A subcellular map of the human proteome. Science 2017, 356, eaal3321. https://doi.org/10.1126/science.aal3321.
86. O’Hagan, S.; Wright Muelas, M.; Day, P.J.; Lundberg, E.; Kell, D.B. GeneGini: Assessment via the Gini coefficient of reference
‘‘housekeeping’’ genes and diverse human transporter expression profiles Cell Syst. 2018, 6, 230–244.
87. Sun, D.; Lennernäs, H.; Welage, L.S.; Barnett, J.L.; Landowski, C.P.; Foster, D.; Fleisher, D.; Lee, K.D.; Amidon, G.L. Comparison
of human duodenum and Caco-2 gene expression profiles for 12,000 gene sequences tags and correlation with permeability of
26 drugs. Pharm. Res 2002, 19, 1400–1416.
88. Anderle, P.; Huang, Y.; Sadée, W. Intestinal membrane transport of drugs and nutrients: Genomics of membrane transporters
using expression microarrays. Eur. J. Pharm. Sci. 2004, 21, 17–24.
89. Landowski, C.P.; Anderle, P.; Sun, D.; Sadee, W.; Amidon, G.L. Transporter and ion channel gene expression after Caco-2 cell
differentiation using 2 different microarray technologies. AAPS J. 2004, 6, e21.
90. Pshezhetsky, A.V.; Fedjaev, M.; Ashmarina, L.; Mazur, A.; Budman, L.; Sinnett, D.; Labuda, D.; Beaulieu, J.F.; Menard, D.; Ni-
fant'ev, I.; et al. Subcellular proteomics of cell differentiation: Quantitative analysis of the plasma membrane proteome of Caco-
2 cells. Proteomics 2007, 7, 2201–2215.
91. Ahlin, G.; Hilgendorf, C.; Karlsson, J.; Szigyarto, C.A.; Uhlén, M.; Artursson, P. Endogenous gene and protein expression of
drug-transporting proteins in cell lines routinely used in drug discovery programs. Drug Metab. Dispos. 2009, 37, 2275–2283.
92. Fersht, A. Enzyme Structure and Mechanism, 2nd ed.; W.H. Freeman: San Francisco, CA, USA, 1977.
93. Keleti, T. Basic Enzyme Kinetics; Akadémiai Kiadó: Budapest, Hungary, 1986.
94. Cornish-Bowden, A. Fundamentals of Enzyme Kinetics, 2nd ed.; Portland Press: London, UK, 1995.
95. Denning, E.J.; Beckstein, O. Influence of lipids on protein-mediated transmembrane transport. Chem. Phys. Lipids 2013, 169, 5771.
96. Lanthaler, K.; Bilsland, E.; Dobson, P.; Moss, H.J.; Pir, P.; Kell, D.B.; Oliver, S.G. Genome-wide assessment of the carriers in-
volved in the cellular uptake of drugs: A model system in yeast. BMC Biol. 2011, 9, 70.
97. Winter, G.E.; Radic, B.; Mayor-Ruiz, C.; Blomen, V.A.; Trefzer, C.; Kandasamy, R.K.; Huber, K.V.M.; Gridling, M.; Chen, D.;
Klampfl, T.; et al. The solute carrier SLC35F2 enables YM155-mediated DNA damage toxicity. Nat. Chem. Biol. 2014, 10, 768–773.
98. Birsoy, K.; Wang, T.; Possemato, R.; Yilmaz, O.H.; Koch, C.E.; Chen, W.W.; Hutchins, A.W.; Gultekin, Y.; Peterson, T.R.; Carette,
J.E.; et al. MCT1-mediated transport of a toxic molecule is an effective strategy for targeting glycolytic tumors. Nat. Genet. 2013,
45, 104–108.
99. Bailey, T.L.; Nieto, A.; McDonald, P.H. A Nonradioactive High-Throughput Screening-Compatible Cell-Based Assay to Identify
Inhibitors of the Monocarboxylate Transporter Protein 1. Assay Drug Dev. Technol. 2019, 17, 275–284.
100. Wright Muelas, M.; Mughal, F.; O’Hagan, S.; Day, P.J.; Kell, D.B. The role and robustness of the Gini coefficient as an unbiased
tool for the selection of Gini genes for normalising expression profiling data Sci. Rep. 2019, 9, 17960.
101. Girardi, E.; César-Razquin, A.; Lindinger, S.; Papakostas, K.; Lindinger, S.; Konecka, J.; Hemmerich, J.; Kickinger, S.; Kartnig,
F.; Gürtl, B.; et al. A widespread role for SLC transmembrane transporters in resistance to cytotoxic drugs. Nat. Chem. Biol. 2020,
16, 469–478.
Molecules 2021, 26, 5629 21 of 39
102. Kell, D.B. Hitchhiking into the cell. Nat. Chem. Biol .2020, 16, 367–368.
103. Gründemann, D.; Harlfinger, S.; Golz, S.; Geerts, A.; Lazar, A.; Berkels, R.; Jung, N.; Rubbert, A.; Schömig, E. Discovery of the
ergothioneine transporter. Proc. Natl. Acad. Sci. USA 2005, 102, 5256–5261.
104. Babcock, J.J.; Li, M. Deorphanizing the human transmembrane genome: A landscape of uncharacterized membrane proteins.
Acta Pharmacol. Sin. 2014, 35, 11–23.
105. Stieger, B.; Hagenbuch, B. Recent advances in understanding hepatic drug transport. F1000Research 2016, 5, 2465.
106. Hashimoto, M.; Girardi, E.; Eichner, R.; Superti-Furga, G. Detection of Chemical Engagement of Solute Carrier Proteins by a
Cellular Thermal Shift Assay. ACS Chem. Biol. 2018, 13, 1480–1486.
107. Clémençon, B.; Lüscher, B.P.; Hediger, M.A. Establishment of a novel microscale thermophoresis ligand-binding assay for char-
acterization of SLC solute carriers using oligopeptide transporter PepT1 (SLC15 family) as a model system. J. Pharmacol. Toxicol.
Methods 2018, 92, 67–76.
108. Girardi, E.; Agrimi, G.; Goldmann, U.; Fiume, G.; Lindinger, S.; Sedlyarov, V.; Srndic, I.; Gurtl, B.; Agerer, B.; Kartnig, F.; et al.
Epistasis-driven identification of SLC25A51 as a regulator of human mitochondrial NAD import. Nat. Commun. 2020, 11, 6145.
109. Yee, S.W.; Buitrago, D.; Stecula, A.; Ngo, H.X.; Chien, H.C.; Zou, L.; Koleske, M.L.; Giacomini, K.M. Deorphaning a solute carrier
22 family member, SLC22A15, through functional genomic studies. FASEB J. 2020, 34, 15734–15752.
110. André, P.; Debray, M.; Scherrmann, J.M.; Cisternino, S. Clonidine transport at the mouse blood-brain barrier by a new H+ anti-
porter that interacts with addictive drugs. J. Cereb. Blood Flow Metab. 2009, 29, 1293–1304.
111. Auvity, S.; Chapy, H.; Goutal, S.; Caille, F.; Hosten, B.; Smirnova, M.; Decleves, X.; Tournier, N.; Cisternino, S. Diphenhydramine
as a selective probe to study H+-antiporter function at the blood-brain barrier: Application to [11C]diphenhydramine positron
emission tomography imaging. J. Cereb. Blood Flow Metab. 2017, 37, 2185–2195.
112. Chapy, H.; Smirnova, M.; Andre, P.; Schlatter, J.; Chiadmi, F.; Couraud, P.O.; Scherrmann, J.M.; Decleves, X.; Cisternino, S.
Carrier-mediated cocaine transport at the blood-brain barrier as a putative mechanism in addiction liability. Int. J. Neuropsycho-
pharmacol. 2014, 18, pyu001.
113. Chapy, H.; André, P.; Declèves, X.; Scherrmann, J.M.; Cisternino, S. A polyspecific drug/proton antiporter mediates diphenhy-
dramine and clonidine transport at the mouse blood-retinal barrier. Br. J. Pharmacol. 2015, 172, 4714–4725.
114. Okura, T.; Hattori, A.; Takano, Y.; Sato, T.; Hammarlund-Udenaes, M.; Terasaki, T.; Deguchi, Y. Involvement of the pyrilamine
transporter, a putative organic cation transporter, in blood-brain barrier transport of oxycodone. Drug Metab. Dispos. 2008, 36,
2005–2013.
115. Okura, T.; Higuchi, K.; Kitamura, A.; Deguchi, Y. Proton-coupled organic cation antiporter-mediated uptake of apomorphine
enantiomers in human brain capillary endothelial cell line hCMEC/D3. Biol Pharm. Bull. 2014, 37, 286–291.
116. Tega, Y.; Kubo, Y.; Yuzurihara, C.; Akanuma, S.; Hosoya, K. Carrier-Mediated Transport of Nicotine Across the Inner Blood-
Retinal Barrier: Involvement of a Novel Organic Cation Transporter Driven by an Outward H+ Gradient. J. Pharm. Sci. 2015, 104,
3069–3075.
117. Tega, Y.; Akanuma, S.; Kubo, Y.; Hosoya, K. Involvement of the H+/organic cation antiporter in nicotine transport in rat liver.
Drug Metab. Dispos. 2015, 43, 89–92.
118. Dickens, D.; Rädisch, S.; Chiduza, G.N.; Giannoudis, A.; Cross, M.J.; Malik, H.; Schaeffeler, E.; Sison-Young, R.L.; Wilkinson,
E.L.; Goldring, C.E.; et al. Cellular uptake of the atypical antipsychotic clozapine is a carrier-mediated process. Mol. Pharm. 2018,
15, 3557–3572.
119. Dvorak, V.; Wiedmer, T.; Ingles-Prieto, A.; Altermatt, P.; Batoulis, H.; Bärenz, F.; Bender, E.; Digles, D.; Dürrenberger, F.; Heit-
man, L.H.; et al. An overview of cell-based assay platforms for the solute carriers family of transporters. Front. Pharmacol. 2021,
12, 722889.
120. Mestres, J.; Gregori-Puigjané, E.; Valverde, S.; Solé, R.V. The topology of drug-target interaction networks: Implicit dependence
on drug properties and target families. Mol. Biosyst. 2009, 5, 1051–1057.
121. Hu, Y.; Gupta-Ostermann, D.; Bajorath, J. Exploring compound promiscuity patterns and multi-target activity spaces. Comput.
Struct. Biotechnol. J. 2014, 9, e201401003.
122. Bajorath, J. Analyzing Promiscuity at the Level of Active Compounds and Targets. Mol. Inform. 2016, 35, 583–587.
123. Gilberg, E.; Jasial, S.; Stumpfe, D.; Dimova, D.; Bajorath, J. Highly Promiscuous Small Molecules from Biological Screening
Assays Include Many Pan-Assay Interference Compounds but Also Candidates for Polypharmacology. J. Med. Chem. 2016, 59,
10285–10290.
124. Hu, Y.; Bajorath, J. Entering the 'big data' era in medicinal chemistry: Molecular promiscuity analysis revisited. Future Sci. OA
2017, 3, FSO179.
125. Bofill, A.; Jalencas, X.; Oprea, T.I.; Mestres, J. The human endogenous metabolome as a pharmacology baseline for drug discov-
ery. Drug Discov. Today 2019, 24, 1806–1820.
126. Cerisier, N.; Petitjean, M.; Regad, L.; Bayard, Q.; Réau, M.; Badel, A.; Camproux, A.C. High Impact: The Role of Promiscuous
Binding Sites in Polypharmacology. Molecules 2019, 24, 2529.
127. Feldmann, C.; Miljkovic, F.; Yonchev, D.; Bajorath, J. Identifying Promiscuous Compounds with Activity against Different Tar-
get Classes. Molecules 2019, 24, 4185.
128. Perrin, J.; Werner, T.; Kurzawa, N.; Rutkowska, A.; Childs, D.D.; Kalxdorf, M.; Poeckel, D.; Stonehouse, E.; Strohmer, K.; Heller,
B.; et al. Identifying drug targets in tissues and whole blood with thermal-shift profiling. Nat. Biotechnol. 2020, 38, 303–308.
Molecules 2021, 26, 5629 22 of 39
129. Yang, Z.Y.; He, J.H.; Lu, A.P.; Hou, T.J.; Cao, D.S. Frequent hitters: Nuisance artifacts in high-throughput screening. Drug Discov.
Today 2020, 25, 657–667.
130. Niphakis, M.J.; Lum, K.M.; Cognetta, A.B.; Correia, B.E.; Ichu, T.A.; Olucha, J.; Brown, S.J.; Kundu, S.; Piscitelli, F.; Rosen, H.; et
al. A Global Map of Lipid-Binding Proteins and Their Ligandability in Cells. Cell 2015, 161, 1668–1680.
131. Superti-Furga, G.; Lackner, D.; Wiedmer, T.; Ingles-Prieto, A.; Barbosa, B.; Girardi, E.; Ulrich Goldman; Gürtl, B.; Klavins, K.;
Klimek, C.; Lindinge, S.; et al. The RESOLUTE consortium: Unlocking SLC transporters for drug discovery. Nat. Rev. Drug
Discov. 2020, 19, 429–430.
132. Garst, A.D.; Bassalo, M.C.; Pines, G.; Lynch, S.A.; Halweg-Edwards, A.L.; Liu, R.; Liang, L.; Wang, Z.; Zeitoun, R.; Alexander,
W.G.; et al. Genome-wide mapping of mutations at single-nucleotide resolution for protein, metabolic and genome engineering.
Nat. Biotechnol. 2017, 35, 48–55.
133. McGlincy, N.J.; Meacham, Z.A.; Reynaud, K.K.; Muller, R.; Baum, R.; Ingolia, N.T. A genome-scale CRISPR interference guide
library enables comprehensive phenotypic profiling in yeast. BMC Genom. 2021, 22, 205.
134. Stovicek, V.; Holkenbrink, C.; Borodina, I. CRISPR/Cas system for yeast genome engineering: Advances and applications. FEMS
Yeast Res. 2017, 17, fox030.
135. Anzalone, A.V.; Koblan, L.W.; Liu, D.R. Genome editing with CRISPR-Cas nucleases, base editors, transposases and prime
editors. Nat. Biotechnol. 2020, 38, 824–844.
136. Hanna, R.E.; Doench, J.G. Design and analysis of CRISPR-Cas experiments. Nat. Biotechnol. 2020, 38, 813–823.
137. Lee, H.K.; Oh, Y.; Hong, J.; Lee, S.H.; Hur, J.K. Development of CRISPR technology for precise single-base genome editing: A
brief review. BMB Rep. 2021, 54, 98–105.
138. Prideaux, B.; Dartois, V.; Staab, D.; Weiner, D.M.; Goh, A.; Via, L.E.; Barry, C.E. 3rd; Stoeckli, M. High-sensitivity MALDI-MRM-
MS imaging of moxifloxacin distribution in tuberculosis-infected rabbit lungs and granulomatous lesions. Anal. Chem. 2011, 83,
2112–2118.
139. Sarathy, J.P.; Dartois, V.; Lee, E.J. The role of transport mechanisms in Mycobacterium tuberculosis drug resistance and tolerance.
Pharmaceuticals 2012, 5, 1210–1235.
140. Wang, N.; Dartois, V.; Carter, C.L. An optimized method for the detection and spatial distribution of aminoglycoside and van-
comycin antibiotics in tissue sections by mass spectrometry imaging. J. Mass Spectrom. 2021, 56, e4708.
141. Römpp, A.; Guenther, S.; Takats, Z.; Spengler, B. Mass spectrometry imaging with high resolution in mass and space (HR2 MSI)
for reliable investigation of drug compound distributions on the cellular level. Anal. Bioanal. Chem. 2011, 401, 65–73.
142. Lietz, C.B.; Gemperline, E.; Li, L. Qualitative and quantitative mass spectrometry imaging of drugs and metabolites. Adv. Drug
Deliv. Rev. 2013, 65, 1074–1085.
143. Morosi, L.; Zucchetti, M.; D'Incalci, M.; Davoli, E. Imaging mass spectrometry: Challenges in visualization of drug distribution
in solid tumors. Curr. Opin. Pharmacol. 2013, 13, 807–812.
144. Patel, K.J.; Trédan, O.; Tannock, I.F. Distribution of the anticancer drugs doxorubicin, mitoxantrone and topotecan in tumors
and normal tissues. Cancer Chemother. Pharmacol. 2013, 72, 127–138.
145. Nerini, I.F.; Morosi, L.; Zucchetti, M.; Ballerini, A.; Giavazzi, R.; D'Incalci, M. lntratumor Heterogeneity and Its Impact on Drug
Distribution and Sensitivity. Clin. Pharmacol. Ther. 2014, 96, 224–238.
146. Prideaux, B.; Stoeckli, M. Mass spectrometry imaging for drug distribution studies. J. Proteom. 2012, 75, 4999–5013.
147. Prideaux, B.; ElNaggar, M.S.; Zimmerman, M.; Wiseman, J.M.; Li, X.H.; Dartois, V. Mass spectrometry imaging of levofloxacin
distribution in TB-infected pulmonary lesions by MALDI-MSI and continuous liquid microjunction surface sampling. Int. J.
Mass Spectrom. 2015, 377, 699–708.
148. Swales, J.G.; Tucker, J.W.; Spreadborough, M.J.; Iverson, S.L.; Clench, M.R.; Webborn, P.J.; Goodwin, R.J. Mapping drug distri-
bution in brain tissue using liquid extraction surface analysis mass spectrometry imaging. Anal. Chem. 2015, 87, 10146–10152.
149. Mann, A.; Han, H.; Eyal, S. Imaging transporters: Transforming diagnostic and therapeutic development. Clin. Pharmacol. Ther.
2016, 100, 479–488.
150. Esteve, C.; Jones, E.A.; Kell, D.B.; Boutin, H.; McDonnell, L.A. Mass spectrometry imaging shows major derangements in neu-
rogranin and in purine metabolism in the triple-knockout 3xTg Alzheimer-like mouse model. Biochim. Biophys. Acta 2017, 1865,
747–754.
151. Prideaux, B.; Lenaerts, A.; Dartois, V. Imaging and spatially resolved quantification of drug distribution in tissues by mass
spectrometry. Curr. Opin. Chem. Biol. 2018, 44, 93–100.
152. Karlsson, O.; Hanrieder, J. Imaging mass spectrometry in drug development and toxicology. Arch. Toxicol. 2017, 91, 2283–2294.
153. Sørensen, I.S.; Janfelt, C.; Nielsen, M.M.B.; Mortensen, R.W.; Knudsen, N.Ø.; Eriksson, A.H.; Pedersen, A.J.; Nielsen, K.T. Com-
bination of MALDI-MSI and cassette dosing for evaluation of drug distribution in human skin explant. Anal. Bioanal. Chem.
2017, 409, 4993–5005.
154. Tournier, N.; Stieger, B.; Langer, O. Imaging techniques to study drug transporter function in vivo. Pharmacol. Ther. 2018, 189,
104–122.
155. Son, H.; Jang, K.; Lee, H.; Kim, S.E.; Kang, K.W.; Lee, H. Use of Molecular Imaging in Clinical Drug Development: A Systematic
Review. Nucl. Med. Mol. Imaging 2019, 53, 208–215.
156. Gilmore, I.S.; Heiles, S.; Pieterse, C.L. Metabolic Imaging at the Single-Cell Scale: Recent Advances in Mass Spectrometry Imag-
ing. Annu. Rev. Anal. Chem. 2019, 12, 201–224.
Molecules 2021, 26, 5629 23 of 39
157. Mokosch, A.S.; Gerbig, S.; Grevelding, C.G.; Haeberlein, S.; Spengler, B. High-resolution AP-SMALDI MSI as a tool for drug
imaging in Schistosoma mansoni. Anal. Bioanal. Chem. 2021, 413, 2755–2766.
158. Newman, C.F.; Havelund, R.; Passarelli, M.K.; Marshall, P.S.; Francis, I.; West, A.; Alexander, M.R.; Gilmore, I.S.; Dollery, C.T.
Intracellular Drug Uptake-A Comparison of Single Cell Measurements Using ToF-SIMS Imaging and Quantification from Cell
Populations with LC/MS/MS. Anal. Chem. 2017, 89, 11944–11953.
159. Fox, B.W.; Schroeder, F.C. Toward spatially resolved metabolomics. Nat. Chem. Biol. 2020, 16, 1039–1040.
160. Pardridge, W.M. Blood-brain barrier endogenous transporters as therapeutic targets: A new model for small molecule CNS
drug discovery. Expert Opin. Ther. Targets 2015, 19, 1059–1072.
161. Brzica, H.; Abdullahi, W.; Ibbotson, K.; Ronaldson, P.T. Role of Transporters in Central Nervous System Drug Delivery and
Blood-Brain Barrier Protection: Relevance to Treatment of Stroke. J. Cent. Nerv. Syst. Dis. 2017, 9, 1179573517693802.
162. Al-Majdoub, Z.M.; Al Feteisi, H.; Achour, B.; Warwood, S.; Neuhoff, S.; Rostami-Hodjegan, A.; Barber, J. Proteomic Quantifica-
tion of Human Blood-Brain Barrier SLC and ABC Transporters in Healthy Individuals and Dementia Patients. Mol. Pharm. 2019,
16, 1220–1233.
163. Gomez-Zepeda, D.; Taghi, M.; Scherrmann, J.M.; Decleves, X.; Menet, M.C. ABC Transporters at the Blood-Brain Interfaces,
Their Study Models, and Drug Delivery Implications in Gliomas. Pharmaceutics 2019, 12, 20.
164. Williams, E.I.; Betterton, R.D.; Davis, T.P.; Ronaldson, P.T. Transporter-Mediated Delivery of Small Molecule Drugs to the Brain:
A Critical Mechanism That Can Advance Therapeutic Development for Ischemic Stroke. Pharmaceutics 2020, 12, 154.
165. Uhlén, M.; Fagerberg, L.; Hallstrom, B.M.; Lindskog, C.; Oksvold, P.; Mardinoglu, A.; Sivertsson, Ǻ.; Kampf, C.; Sjöstedt, E.;
Asplund, A.; et al. Tissue-based map of the human proteome. Science 2015, 347, 1260419.
166. Kell, D.B. Control of metabolite efflux in microbial cell factories: Current advances and future prospects. OSF Preprints 2018,
xg9jh, doi:10.31219/osf.io/xg9jh.
167. Sauer, M.; Porro, D.; Mattanovich, D.; Branduardi, P. Microbial production of organic acids: Expanding the markets. Trends
Biotechnol. 2008, 26, 100–108.
168. Boyarskiy, S.; Tullman-Ercek, D. Getting pumped: Membrane efflux transporters for enhanced biomolecule production. Curr.
Opin. Chem. Biol. 2015, 28, 15–19.
169. Mukhopadhyay, A. Tolerance engineering in bacteria for the production of advanced biofuels and chemicals. Trends Microbiol.
2015, 23, 498–508.
170. Jones, C.M.; Hernández Lozada, N.J.; Pfleger, B.F. Efflux systems in bacteria and their metabolic engineering applications. Appl.
Microbiol. Biotechnol. 2015, 99, 9381–9393.
171. Van der Hoek, S.A.; Borodina, I. Transporter engineering in microbial cell factories: The ins, the outs, and the in-betweens. Curr.
Opin. Biotechnol. 2020, 66, 186–194.
172. Wang, G.; Møller-Hansen, I.; Babaei, M.; D’Ambrosio, V.; Christensen, H.B.; Darbani, B.; Jensen, M.K.; Borodina, I.
Transportome-wide engineering of Saccharomyces cerevisiae. Metab. Eng. 2021, 64, 52–63.
173. Jezierska, S.; Van Bogaert, I.N.A. Crossing boundaries: The importance of cellular membranes in industrial biotechnology. J.
Ind. Microbiol. Biotechnol. 2017, 44, 721–733.
174. Zhu, Y.; Zhou, C.; Wang, Y.; Li, C. Transporter Engineering for Microbial Manufacturing. Biotechnol. J. 2020, 15, e1900494.
175. Onyeabor, M.; Martinez, R.; Kurgan, G.; Wang, X. Engineering transport systems for microbial production. Adv. Appl. Microbiol.
2020, 111, 33–87.
176. Soares-Silva, I.; Ribas, D.; Sousa-Silva, M.; Azevedo-Silva, J.; Rendulić, T.; Casal, M. Membrane transporters in the bioproduc-
tion of organic acids: State of the art and future perspectives for industrial applications. FEMS Microbiol. Lett. 2020, 367, fnaa118.
177. Lane, T.S.; Rempe, C.S.; Davitt, J.; Staton, M.E.; Peng, Y.; Soltis, D.E.; Melkonian, M.; Deyholos, M.; Leebens-Mack, J.H.; Chase,
M.; et al. Diversity of ABC transporter genes across the plant kingdom and their potential utility in biotechnology. BMC Bio-
technol. 2016, 16, 47.
178. Nielsen, J.; Keasling, J.D. Engineering Cellular Metabolism. Cell 2016, 164, 1185–1197.
179. Fang, S.; Huang, X.; Zhang, X.; Zhang, M.; Hao, Y.; Guo, H.; Liu, L.N.; Yu, F.; Zhang, P. Molecular mechanism underlying
transport and allosteric inhibition of bicarbonate transporter SbtA. Proc. Natl. Acad. Sci. USA 2021, 118, e2101632118.
180. Zhang, C.; Chen, X.; Stephanopoulos, G.; Too, H.P. Efflux transporter engineering markedly improves amorphadiene produc-
tion in Escherichia coli. Biotechnol. Bioeng. 2016, 113, 1755–1763.
181. Steiger, M.G.; Rassinger, A.; Mattanovich, D.; Sauer, M. Engineering of the citrate exporter protein enables high citric acid pro-
duction in Aspergillus niger. Metab. Eng. 2019, 52, 224–231.
182. Kurgan, G.; Kurgan, L.; Schneider, A.; Onyeabor, M.; Rodriguez-Sanchez, Y.; Taylor, E.; Martinez, R.; Carbonell, P.; Shi, X.; Gu,
H.; et al. Identification of major malate export systems in an engineered malate-producing Escherichia coli aided by substrate
similarity search. Appl. Microbiol. Biotechnol. 2019, 103, 9001–9011.
183. Chen, X.; Wang, Y.; Dong, X.; Hu, G.; Liu, L. Engineering rTCA pathway and C4-dicarboxylate transporter for L-malic acid
production. Appl. Microbiol. Biotechnol. 2017, 101, 4041–4052.
184. Cao, W.; Yan, L.; Li, M.; Liu, X.; Xu, Y.; Xie, Z.; Liu, H. Identification and engineering a C4-dicarboxylate transporter for im-
provement of malic acid production in Aspergillus niger. Appl Microbiol. Biotechnol. 2020, 104, 9773–9783.
185. Severi, E.; Thomas, G.H. Antibiotic export: Transporters involved in the final step of natural product production. Microbiology
2019, 165, 805–818.
Molecules 2021, 26, 5629 24 of 39
186. Salvador López, J.M.; Van Bogaert, I.N.A. Microbial fatty acid transport proteins and their biotechnological potential. Biotechnol.
Bioeng. 2021, 118, 2184–2201.
187. Hu, Y.; Zhu, Z.; Nielsen, J.; Siewers, V. Heterologous transporter expression for improved fatty alcohol secretion in yeast. Metab.
Eng. 2018, 45, 51–58.
188. Mingardon, F.; Clement, C.; Hirano, K.; Nhan, M.; Luning, E.G.; Chanal, A.; Mukhopadhyay, A. Improving olefin tolerance and
production in E. coli using native and evolved AcrB. Biotechnol. Bioeng. 2015, 112, 879–888.
189. Darbani, B.; Stovicek, V.; van der Hoek, S.A.; Borodina, I. Engineering energetically efficient transport of dicarboxylic acids in
yeast Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. USA 2019, 116, 19415–19420.
190. Lennen, R.M.; Jensen, K.; Mohammed, E.T.; Malla, S.; Börner, R.A.; Chekina, K.; Özdemir, E.; Bonde, I.; Koza, A.; Maury, J.; et
al. Adaptive laboratory evolution reveals general and specific chemical tolerance mechanisms and enhances biochemical pro-
duction. bioRxiv 2019, 634105, https://doi.org/10.1101/634105.
191. Claus, S.; Jenkins Sánchez, L.; Van Bogaert, I.N.A. The role of transport proteins in the production of microbial glycolipid bio-
surfactants. Appl. Microbiol. Biotechnol. 2021, 105, 1779–1793.
192. Jezierska, S.; Claus, S.; Van Bogaert, I.N.A. Identification and importance of mitochondrial citrate carriers and ATP citrate lyase
for glycolipid production in Starmerella bombicola. Appl. Microbiol. Biotechnol. 2020, 104, 6235–6248.
193. Markakis, K.; Lowe, P.T.; Davison-Gates, L.; O’Hagan, D.; Rosser, S.J.; Elfick, A. An Engineered E. coli Strain for Direct in Vivo
Fluorination. ChemBioChem 2020, 21, 1856–1860.
194. Claus, S.; Jezierska, S.; Van Bogaert, I.N.A. Protein-facilitated transport of hydrophobic molecules across the yeast plasma mem-
brane. FEBS Lett. 2019, 593, 1508–1527.
195. Hult, K.; Berglund, P. Enzyme promiscuity: Mechanism and applications. Trends Biotechnol. 2007, 25, 231–238.
196. Babtie, A.C.; Bandyopadhyay, S.; Olguin, L.F.; Hollfelder, F. Efficient catalytic promiscuity for chemically distinct reactions.
Angew. Chem. Int. Ed. Engl. 2009, 48, 3692–3694.
197. Nobeli, I.; Favia, A.D.; Thornton, J.M. Protein promiscuity and its implications for biotechnology. Nat. Biotechnol. 2009, 27, 157167.
198. Carbonell, P.; Faulon, J.L. Molecular signatures-based prediction of enzyme promiscuity. Bioinformatics 2010, 26, 2012–2019.
199. Carbonell, P.; Lecointre, G.; Faulon, J.L. Origins of specificity and promiscuity in metabolic networks. J. Biol. Chem. 2011, 286,
43994–44004.
200. O’Hagan, S.; Kell, D.B. Consensus rank orderings of molecular fingerprints illustrate the ‘most genuine’ similarities between
marketed drugs and small endogenous human metabolites, but highlight exogenous natural products as the most important
‘natural’ drug transporter substrates. ADMET DMPK 2017, 5, 85–125.
201. Samanta, S.; O’Hagan, S.; Swainston, N.; Roberts, T.J.; Kell, D.B. VAE-Sim: A novel molecular similarity measure based on a
variational autoencoder. Molecules 2020, 25, 3446.
202. Sierzputowska, K.; Baxter, C.R.; Housden, B.E. Variable Dose Analysis: A Novel High-throughput RNAi Screening Method for
Drosophila Cells. Bio-Protocol 2018, 8, e3112.
203. Senior, E.; Bull, A.T.; Slater, J.H. Enzyme evolution in a microbial community growing on the herbicide Dalapon. Nature 1976,
263, 476–479.
204. Barrick, J.E.; Yu, D.S.; Yoon, S.H.; Jeong, H.; Oh, T.K.; Schneider, D.; Lenski, R.E.; Kim, J.F. Genome evolution and adaptation in
a long-term experiment with Escherichia coli. Nature 2009, 461, 1243–1247.
205. Good, B.H.; McDonald, M.J.; Barrick, J.E.; Lenski, R.E.; Desai, M.M. The dynamics of molecular evolution over 60,000 genera-
tions. Nature 2017, 551, 45–50.
206. Tenaillon, O.; Barrick, J.E.; Ribeck, N.; Deatherage, D.E.; Blanchard, J.L.; Dasgupta, A.; Wu, G.C.; Wielgoss, S.; Cruveiller, S.;
Médigue, C.; et al. Tempo and mode of genome evolution in a 50,000-generation experiment. Nature 2016, 536, 165–170.
207. Khan, A.I.; Dinh, D.M.; Schneider, D.; Lenski, R.E.; Cooper, T.F. Negative epistasis between beneficial mutations in an evolving
bacterial population. Science 2011, 332, 1193–1196.
208. Woods, R.J.; Barrick, J.E.; Cooper, T.F.; Shrestha, U.; Kauth, M.R.; Lenski, R.E. Second-order selection for evolvability in a large
Escherichia coli population. Science 2011, 331, 1433–1436.
209. Lenski, R.E.; Wiser, M.J.; Ribeck, N.; Blount, Z.D.; Nahum, J.R.; Morris, J.J.; Zaman, L.; Turner, C.B.; Wade, B.D.; Maddamsetti,
R.; et al. Sustained fitness gains and variability in fitness trajectories in the long-term evolution experiment with Escherichia coli.
Proc. Biol. Sci. 2015, 282, 20152292.
210. Van den Bergh, B.; Swings, T.; Fauvart, M.; Michiels, J. Experimental Design, Population Dynamics, and Diversity in Microbial
Experimental Evolution. Microbiol. Mol. Biol. Rev. 2018, 82, e00008-18.
211. Dragosits, M.; Mattanovich, D. Adaptive laboratory evolutionPrinciples and applications for biotechnology. Microb. Cell Factories
2013, 12, 64.
212. LaCroix, R.A.; Palsson, B.O.; Feist, A.M. A Model for Designing Adaptive Laboratory Evolution Experiments. Appl. Environ.
Microbiol. 2017, 83, e03115-16.
213. Mundhada, H.; Seoane, J.M.; Schneider, K.; Koza, A.; Christensen, H.B.; Klein, T.; Phaneuf, P.V.; Herrgard, M.; Feist, A.M.;
Nielsen, A.T. Increased production of L-serine in Escherichia coli through Adaptive Laboratory Evolution. Metab. Eng. 2017, 39,
141–150.
214. Pereira, R.; Wei, Y.; Mohamed, E.; Radi, M.; Malina, C.; Herrgård, M.J.; Feist, A.M.; Nielsen, J.; Chen, Y. Adaptive laboratory
evolution of tolerance to dicarboxylic acids in Saccharomyces cerevisiae. Metab. Eng. 2019, 56, 130–141.
Molecules 2021, 26, 5629 25 of 39
215. Phaneuf, P.V.; Gosting, D.; Palsson, B.O.; Feist, A.M. ALEdb 1.0: A database of mutations from adaptive laboratory evolution
experimentation. Nucleic Acids Res. 2019, 47, D1164–D1171.
216. Phaneuf, P.V.; Yurkovich, J.T.; Heckmann, D.; Wu, M.; Sandberg, T.E.; King, Z.A.; Tan, J.; Palsson, B.O.; Feist, A.M. Causal
mutations from adaptive laboratory evolution are outlined by multiple scales of genome annotations and condition-specificity.
BMC Genom. 2020, 21, 514.
217. Portnoy, V.A.; Bezdan, D.; Zengler, K. Adaptive laboratory evolutionHarnessing the power of biology for metabolic engi-
neering. Curr. Opin. Biotechnol. 2011, 22, 590–594.
218. Reyes, L.H.; Kao, K.C. Growth-Coupled Carotenoids Production Using Adaptive Laboratory Evolution. Methods Mol. Biol. 2018,
1671, 319–330.
219. Winkler, J.; Reyes, L.H.; Kao, K.C. Adaptive laboratory evolution for strain engineering. Methods Mol. Biol. 2013, 985, 211–222.
220. Godara, A.; Kao, K.C. Adaptive laboratory evolution for growth coupled microbial production. World J. Microbiol. Biotechnol.
2020, 36, 175.
221. Lee, S.; Kim, P. Current Status and Applications of Adaptive Laboratory Evolution in Industrial Microorganisms. J. Microbiol.
Biotechnol. 2020, 30, 793–803.
222. Sandberg, T.E.; Salazar, M.J.; Weng, L.L.; Palsson, B.O.; Feist, A.M. The emergence of adaptive laboratory evolution as an effi-
cient tool for biological discovery and industrial biotechnology. Metab. Eng. 2019, 56, 116.
223. Zhu, Z.; Zhang, J.; Ji, X.; Fang, Z.; Wu, Z.; Chen, J.; Du, G. Evolutionary engineering of industrial microorganisms-strategies and
applications. Appl. Microbiol. Biotechnol. 2018, 102, 4615–4627.
224. Dykhuizen, D.E.; Hartl, D.L. Selection in Chemostats. Microbiol. Rev. 1983, 47, 150–168.
225. Flegr, J. Two distinct types of natural selection in turbidostat-like and chemostat-like ecosystems. J. Theor. Biol. 1997, 188, 121126.
226. Gresham, D.; Dunham, M.J. The enduring utility of continuous culturing in experimental evolution. Genomics 2014, 104, 399405.
227. McGeachy, A.M.; Meacham, Z.A.; Ingolia, N.T. An Accessible Continuous-Culture Turbidostat for Pooled Analysis of Complex
Libraries. ACS Synth. Biol. 2019, 8, 844–856.
228. Delneri, D.; Leong, H.S.; Hayes, A.; Davey, H.M.; Kell, D.B.; Oliver, S.G. Assessing contributions to fitness of individual genes
via genome-wide competition analysis. Yeast 2003, 20, S337.
229. Delneri, D.; Hoyle, D.C.; Gkargkas, K.; Cross, E.J.; Rash, B.; Zeef, L.; Leong, H.S.; Davey, H.M.; Hayes, A.; Kell, D.B.; et al.
Identification and characterization of high-flux-control genes of yeast through competition analyses in continuous cultures. Nat.
Genet. 2008, 40, 113–117.
230. Pir, P.; Gutteridge, A.; Wu, J.; Rash, B.; Kell, D.B.; Zhang, N.; Oliver, S.G. The genetic control of growth rate: A systems biology
study in yeast. BMC Syst. Biol. 2012, 6, 4.
231. Wortel, M.T.; Bosdriesz, E.; Teusink, B.; Bruggeman, F.J. Evolutionary pressures on microbial metabolic strategies in the che-
mostat. Sci. Rep. 2016, 6, 29503.
232. Kell, D.B.; Oliver, S.G. Here is the evidence, now what is the hypothesis? The complementary roles of inductive and hypothesis-
driven science in the post-genomic era. Bioessays 2004, 26, 99–105.
233. Bennett, R.K.; Gregory, G.J.; Gonzalez, J.E.; Har, J.R.G.; Antoniewicz, M.R.; Papoutsakis, E.T. Improving the Methanol Tolerance
of an Escherichia coli Methylotroph via Adaptive Laboratory Evolution Enhances Synthetic Methanol Utilization. Front. Micro-
biol. 2021, 12, 638426.
234. Pereira, R.; Mohamed, E.T.; Radi, M.S.; Herrgård, M.J.; Feist, A.M.; Nielsen, J.; Chen, Y. Elucidating aromatic acid tolerance at
low pH in Saccharomyces cerevisiae using adaptive laboratory evolution. Proc. Natl. Acad. Sci. USA 2020, 117, 27954–27961.
235. Mohamed, E.T.; Wang, S.; Lennen, R.M.; Herrgard, M.J.; Simmons, B.A.; Singer, S.W.; Feist, A.M. Generation of a platform strain
for ionic liquid tolerance using adaptive laboratory evolution. Microb. Cell Factories 2017, 16, 204.
236. Mitchell, A.M.; Srikumar, T.; Silhavy, T.J. Cyclic Enterobacterial Common Antigen Maintains the Outer Membrane Permeability
Barrier of Escherichia coli in a Manner Controlled by YhdP. mBio 2018, 9, e01321-18.
237. Grimm, J.; Shi, H.; Wang, W.; Mitchell, A.M.; Wingreen, N.S.; Huang, K.C.; Silhavy, T.J. The inner membrane protein YhdP
modulates the rate of anterograde phospholipid flow in Escherichia coli. Proc. Natl. Acad. Sci. USA 2020, 117, 26907–26914.
238. Avrahami-Moyal, L.; Engelberg, D.; Wenger, J.W.; Sherlock, G.; Braun, S. Turbidostat culture of Saccharomyces cerevisiae W303-
1A under selective pressure elicited by ethanol selects for mutations in SSD1 and UTH1. FEMS Yeast Res. 2012, 12, 521–533.
239. Davey, H.M.; Davey, C.L.; Woodward, A.M.; Edmonds, A.N.; Lee, A.W.; Kell, D.B. Oscillatory, stochastic and chaotic growth
rate fluctuations in permittistatically-controlled yeast cultures. Biosystems 1996, 39, 43–61.
240. Hoffmann, S.A.; Wohltat, C.; Muller, K.M.; Arndt, K.M. A user-friendly, low-cost turbidostat with versatile growth rate estima-
tion based on an extended Kalman filter. PLoS ONE 2017, 12, e0181923.
241. Markx, G.H.; Davey, C.L.; Kell, D.B. The permittistat: A novel type of turbidostat. J. Gen. Microbiol. 1991, 137, 735–743.
242. Munson, R.J. Turbidostats. In Methods in Microbiology; Norris, J.R., Ribbons, D.W., Eds.; Academic Press: 1970; Volume 2, pp.
349–376.
243. Watson, T.G. The Present Status and Future Prospects of the Turbidostat. J. Appl. Chem. Biotechnol. 1972, 22, 229–243.
244. Guarino, A.; Shannon, B.; Marucci, L.; Grierson, C.; Savery, N.; Bernardo, M. A low-cost,open-sourceTurbidostat design for in-
vivo control experiments in Synthetic Biology. IFAC Pap. Online 2019, 52, 244–248.
245. Takahashi, C.N.; Miller, A.W.; Ekness, F.; Dunham, M.J.; Klavins, E. A low cost, customizable turbidostat for use in synthetic
circuit characterization. ACS Synth. Biol. 2015, 4, 32–38.
246. Harris, C.M.; Kell, D.B. The estimation of microbial biomass. Biosensors 1985, 1, 17–84.
Molecules 2021, 26, 5629 26 of 39
247. Harris, C.M.; Todd, R.W.; Bungard, S.J.; Lovitt, R.W.; Morris, J.G.; Kell, D.B. The dielectric permittivity of microbial suspensions
at radio frequencies: A novel method for the estimation of microbial biomass. Enzym. Microb. Technol. 1987, 9, 181–186.
248. Kell, D.B.; Markx, G.H.; Davey, C.L.; Todd, R.W. Real-time monitoring of cellular biomass: Methods and applications. Trends
Anal. Chem. 1990, 9, 190–194.
249. Pirt, S.J. Principles of Microbe and Cell Cultivation; Wiley: London, UK, 1975.
250. Rembeza, E.; Engqvist, M.K. Experimental investigation of enzyme functional annotations reveals extensive annotation error.
bioRxiv 2020, https://doi.org/10.1101/2020.12.18.423474.
251. Borodina, I.; Kenny, L.C.; McCarthy, C.M.; Paramasivan, K.; Pretorius, R.; Roberts, T.J.; van der Hoek, S.A.; Kell, D.B. The biol-
ogy of ergothioneine, an antioxidant nutraceutical. Nutr. Res. Rev. 2020, 33, 190–217.
252. Cheah, I.K.; Halliwell, B. Ergothioneine, recent developments. Redox Biol. 2021, 42, 101868.
253. Gründemann, D. The ergothioneine transporter controls and indicates ergothioneine activityA review. Prev. Med. 2012, 54,
S71–S74.
254. Tschirka, J.; Kreisor, M.; Betz, J.; Gründemann, D. Substrate selectivity check of the ergothioneine transporter. Drug Metab.
Dispos. 2018, 46, 779–785.
255. Broadhurst, D.; Goodacre, R.; Reinke, S.N.; Kuligowski, J.; Wilson, I.D.; Lewis, M.R.; Dunn, W.B. Guidelines and considerations
for the use of system suitability and quality control samples in mass spectrometry assays applied in untargeted clinical metab-
olomic studies. Metabolomics 2018, 14, 72.
256. Dunn, W.B.; Erban, A.; Weber, R.J.M.; Creek, D.J.; Brown, M.; Breitling, R.; Hankemeier, T.; Goodacre, R.; Neumann, S.; Kopka,
J.; et al. Mass Appeal: Metabolite identification in mass spectrometry-focused untargeted metabolomics. Metabolomics 2013, 9,
S44–S66.
257. Nash, W.J.; Dunn, W.B. From mass to metabolite in human untargeted metabolomics: Recent advances in annotation of metab-
olites applying liquid chromatography-mass spectrometry data. Trends Anal. Chem. 2019, 120, 115324.
258. Wright Muelas, M.; Roberts, I.; Mughal, F.; O’Hagan, S.; Day, P.J.; Kell, D.B. An untargeted metabolomics strategy to measure
differences in metabolite uptake and excretion by mammalian cell lines. Metabolomics 2020, 16, 107.
259. Sajid, A.; Lusvarghi, S.; Murakami, M.; Chufan, E.E.; Abel, B.; Gottesman, M.M.; Durell, S.R.; Ambudkar, S.V. Reversing the
direction of drug transport mediated by the human multidrug transporter P-glycoprotein. Proc. Natl. Acad. Sci. USA 2020, 117,
29609–29617.
260. Sun, N.; Li, D.; Fonzi, W.; Li, X.; Zhang, L.; Calderone, R. Multidrug-resistant transporter Mdr1p-mediated uptake of a novel
antifungal compound. Antimicrob. Agents Chemother. 2013, 57, 5931–5939.
261. Jindal, S.; Yang, L.; Day, P.J.; Kell, D.B. Involvement of multiple influx and efflux transporters in the accumulation of cationic
fluorescent dyes by Escherichia coli. BMC Microbiol. 2019, 19, 195; also bioRxiv 603688v1.
262. Salcedo-Sora, J.E.; Kell, D.B. A quantitative survey of bacterial persistence in the presence of antibiotics: Towards antipersister
antimicrobial discovery. Antibiotics 2020, 9, 508.
263. Salcedo-Sora, J.E.; Jindal, S.; O’Hagan, S.; Kell, D.B. A palette of fluorophores that are differentially accumulated by wild-type
and mutant strains of Escherichia coli: Surrogate ligands for bacterial membrane transporters. Microbiology 2021, 167, 001016.
264. O’Hagan, S.; Kell, D.B. Structural similarities between some common fluorophores used in biology, marketed drugs, endoge-
nous metabolites, and natural products. Mar. Drugs 2020, 18, 582.
265. Kaprelyants, A.S.; Kell, D.B. Rapid assessment of bacterial viability and vitality using rhodamine 123 and flow cytometry. J.
Appl. Bacteriol. 1992, 72, 410–422.
266. Kaprelyants, A.S.; Kell, D.B. Dormancy in stationary-phase cultures of Micrococcus luteus: Flow cytometric analysis of starvation
and resuscitation. Appl. Env. Microbiol. 1993, 59, 3187–3196.
267. Davey, H.M.; Kell, D.B. Flow cytometry and cell sorting of heterogeneous microbial populations: The importance of single-cell
analysis. Microbiol. Rev. 1996, 60, 641–696.
268. Buranda, T.; Gineste, C.; Wu, Y.; Bondu, V.; Perez, D.; Lake, K.R.; Edwards, B.S.; Sklar, L.A. A High-Throughput Flow Cytom-
etry Screen Identifies Molecules That Inhibit Hantavirus Cell Entry. SLAS Discov. 2018, 23, 634–645.
269. Edwards, B.S.; Sklar, L.A. Flow Cytometry: Impact on Early Drug Discovery. J. Biomol. Screen. 2015, 20, 689–707.
270. Strouse, J.J.; Ivnitski-Steele, I.; Waller, A.; Young, S.M.; Perez, D.; Evangelisti, A.M.; Ursu, O.; Bologa, C.G.; Carter, M.B.; Salas,
V.M.; et al. Fluorescent substrates for flow cytometric evaluation of efflux inhibition in ABCB1, ABCC1, and ABCG2 transport-
ers. Anal. Biochem. 2013, 437, 77–87.
271. Tegos, G.P.; Evangelisti, A.M.; Strouse, J.J.; Ursu, O.; Bologa, C.; Sklar, L.A. A high throughput flow cytometric assay platform
targeting transporter inhibition. Drug Disc. Today Technol. 2014, 12, e95–e103.
272. Jindal, S.; Thampy, H.; Day, P.J.; Kell, D.B. Very rapid flow cytometric assessment of antimicrobial susceptibility during the
apparent lag phase of bacterial (re)growth Microbiology 2019, 165, 439–454.
273. Dragan, A.I.; Pavlovic, R.; McGivney, J.B.; Casas-Finet, J.R.; Bishop, E.S.; Strouse, R.J.; Schenerman, M.A.; Geddes, C.D. SYBR
Green I: Fluorescence properties and interaction with DNA. J. Fluoresc. 2012, 22, 1189–1199.
274. Berney, M.; Vital, M.; Hülshoff, I.; Weilenmann, H.U.; Egli, T.; Hammes, F. Rapid, cultivation-independent assessment of mi-
crobial viability in drinking water. Water Res. 2008, 42, 4010–4018.
275. Hammes, F.; Berney, M.; Wang, Y.; Vital, M.; Köster, O.; Egli, T. Flow-cytometric total bacterial cell counts as a descriptive
microbiological parameter for drinking water treatment processes. Water Res. 2008, 42, 269–277.
Molecules 2021, 26, 5629 27 of 39
276. Mendes, P.; Girardi, E.; Superti-Furga, G.; Kell, D.B. Why most transporter mutations that cause antibiotic resistance are to
efflux pumps rather than to import transporters. bioRxiv 2020, https://doi.org/10.1101/2020.01.16.909507.
277. Featherstone, D.E.; Broadie, K. Wrestling with pleiotropy: Genomic and topological analysis of the yeast gene expression net-
work. Bioessays 2002, 24, 267–274.
278. Rudd, K.E. Linkage map of Escherichia coli K-12, edition 10: The physical map. Microbiol. Mol. Biol. Rev. 1998, 62, 985–1019.
279. Ghatak, S.; King, Z.A.; Sastry, A.; Palsson, B.O. The y-ome defines the 35% of Escherichia coli genes that lack experimental evi-
dence of function. Nucleic Acids Res. 2019, 47, 2446–2454.
280. Yasir, M.; Turner, A.K.; Bastkowski, S.; Baker, D.; Page, A.J.; Telatin, A.; Phan, M.-D.; Monahan, L.; Savva, G.M.; Darling, A.; et
al. TraDIS-Xpress: A high-resolution whole-genome assay identifies novel mechanisms of triclosan action and resistance. Ge-
nome Res. 2020, 30, 239–249.
281. Turner, A.K.; Yasir, M.; Bastkowski, S.; Telatin, A.; Page, A.J.; Charles, I.G.; Webber, M.A. A genome-wide analysis of Escherichia
coli responses to fosfomycin using TraDIS-Xpress reveals novel roles for phosphonate degradation and phosphate transport
systems. J. Antimicrob. Chemother. 2020, 75, 3144–3151.
282. Höglund, P.J.; Nordström, K.J.V.; Schiöth, H.B.; Fredriksson, R. The solute carrier families have a remarkably long evolutionary
history with the majority of the human families present before divergence of Bilaterian species. Mol. Biol. Evol. 2011, 28, 15311541.
283. Hediger, M.A.; Clemencon, B.; Burrier, R.E.; Bruford, E.A. The ABCs of membrane transporters in health and disease (SLC
series): Introduction. Mol. Aspects Med. 2013, 34, 95–107.
284. Rives, M.L.; Javitch, J.A.; Wickenden, A.D. Potentiating SLC transporter activity: Emerging drug discovery opportunities. Bio-
chem Pharmacol 2017, 135, 111.
285. Liu, X. SLC Family Transporters. Adv. Exp. Med. Biol. 2019, 1141, 101–202.
286. Colas, C.; Ung, P.M.U.; Schlessinger, A. SLC transporters: Structure, function, and drug discovery. MedChemComm 2016, 7, 10691081.
287. Pizzagalli, M.D.; Bensimon, A.; Superti-Furga, G. A guide to plasma membrane solute carrier proteins. FEBS J. 2021, 288, 27842835.
288. César-Razquin, A.; Snijder, B.; Frappier-Brinton, T.; Isserlin, R.; Gyimesi, G.; Bai, X.; Reithmeier, R.A.; Hepworth, D.; Hediger,
M.A.; Edwards, A.M.; et al. A call for systematic research on solute carriers. Cell 2015, 162, 478–487.
289. Kory, N.; uit de Bos, J.; van der Rijt, S.; Jankovic, N.; Gura, M.; Arp, N.; Pena, I.A.; Prakash, G.; Chan, S.H.; Kunchok, T.; et al.
MCART1/SLC25A51 is required for mitochondrial NAD transport. Sci. Adv. 2020, 6, eabe5310.
290. Nour-Eldin, H.H.; Nørholm, M.H.H.; Halkier, B.A. Screening for plant transporter function by expressing a normalized Ara-
bidopsis full-length cDNA library in Xenopus oocytes. Plant. Methods 2006, 2, 17.
291. Pfefferkorn, J.A.; Litchfield, J.; Hutchings, R.; Cheng, X.M.; Larsen, S.D.; Auerbach, B.; Bush, M.R.; Lee, C.; Erasga, N.; Bowles,
D.M.; et al. Discovery of novel hepatoselective HMG-CoA reductase inhibitors for treating hypercholesterolemia: A bench-to-
bedside case study on tissue selective drug distribution. Bioorg. Med. Chem. Lett. 2011, 21, 2725–2731.
292. Pfefferkorn, J.A.; Guzman-Perez, A.; Litchfield, J.; Aiello, R.; Treadway, J.L.; Pettersen, J.; Minich, M.L.; Filipski, K.J.; Jones, C.S.;
Tu, M.; et al. Discovery of (S)-6-(3-cyclopentyl-2-(4-(trifluoromethyl)-1H-imidazol-1-yl)propanamido)nicotini c acid as a hepa-
toselective glucokinase activator clinical candidate for treating type 2 diabetes mellitus. J. Med. Chem. 2012, 55, 1318–1333.
293. Meyer zu Schwabedissen, H.E.; Robert, B.; Hussner, J.; Juhnke, B.O.; Gliesche, D.; Boettcher, K.; Sternberg, K.; Schmitz, K.P.;
Kroemer, H.K. Cell specific expression of uptake transportersA potential approach for cardiovascular drug delivery devices.
Mol. Pharm. 2014, 11, 665–672.
294. Grixti, J.; O'Hagan, S.; Day, P.J.; Kell, D.B. Enhancing drug efficacy and therapeutic index through cheminformatics-based se-
lection of small molecule binary weapons that improve transporter-mediated targeting: A cytotoxicity system based on gem-
citabine. Front. Pharmacol. 2017, 8, 155.
295. Orozco, C.C.; Atkinson, K.; Ryu, S.; Chang, G.; Keefer, C.; Lin, J.; Riccardi, K.; Mongillo, R.K.; Tess, D.; Filipski, K.J.; et al. Struc-
tural attributes influencing unbound tissue distribution. Eur. J. Med. Chem. 2020, 185, 111813.
296. Nyquist, M.D.; Corella, A.; Burns, J.; Coleman, I.; Gao, S.; Tharakan, R.; Riggan, L.; Cai, C.; Corey, E.; Nelson, P.S.; et al. Exploit-
ing AR-Regulated Drug Transport to Induce Sensitivity to the Survivin Inhibitor YM155. Mol. Cancer Res. 2017, 15, 521–531.
297. Nyquist, M.D.; Prasad, B.; Mostaghel, E.A. Harnessing Solute Carrier Transporters for Precision Oncology. Molecules 2017, 22, 539.
298. Al-Abdulla, R.; Perez-Silva, L.; Abete, L.; Romero, M.R.; Briz, O.; Marin, J.J.G. Unraveling ‘The Cancer Genome Atlas’ infor-
mation on the role of SLC transporters in anticancer drug uptake. Expert Rev. Clin. Pharmacol. 2019, 12, 329–341.
299. Huang, K.M.; Uddin, M.E.; DiGiacomo, D.; Lustberg, M.B.; Hu, S.; Sparreboom, A. Role of SLC transporters in toxicity induced
by anticancer drugs. Expert Opin. Drug Metab. Toxicol. 2020, 16, 493–506.
300. Wu, Z.; Xu, J.; Liang, C.; Meng, Q.; Hua, J.; Wang, W.; Zhang, B.; Liu, J.; Yu, X.; Shi, S. Emerging roles of the solute carrier family
in pancreatic cancer. Clin. Transl. Med. 2021, 11, e356.
301. Ceriani, L.; Verme, P. The origins of the Gini index: Extracts from Variabilità e Mutabilità (1912) by Corrado Gini. J. Econ. Inequal.
2012, 10, 421–443.
302. Lv, X.; Zhang, G.; Ren, G. Gini index estimation for lifetime data. Lifetime Data Anal. 2017, 23, 275–304.
303. Lawal, H.O.; Krantz, D.E. SLC18: Vesicular neurotransmitter transporters for monoamines and acetylcholine. Mol. Aspects Med.
2013, 34, 360–372.
304. Wilkinson, R.; Pickett, K. The Spirit Level: Why Equality is Better for Everyone; Penguin Books: London, UK, 2009.
305. Mortazavi, A.; Williams, B.A.; McCue, K.; Schaeffer, L.; Wold, B. Mapping and quantifying mammalian transcriptomes by RNA-
Seq. Nat. Methods 2008, 5, 621–658.
Molecules 2021, 26, 5629 28 of 39
306. Wagner, G.P.; Kin, K.; Lynch, V.J. Measurement of mRNA abundance using RNA-seq data: RPKM measure is inconsistent
among samples. Theory Biosci. 2012, 131, 281–255.
307. Zhao, S.; Ye, Z.; Stanton, R. Misuse of RPKM or TPM normalization when comparing across samples and sequencing protocols.
RNA 2020, 26, 903–909.
308. Anand, B.S.; Dey, S.; Mitra, A.K. Current prodrug strategies via membrane transporters/receptors. Expert Opin. Biol. Ther. 2002,
2, 607–620.
309. Clas, S.D.; Sanchez, R.I.; Nofsinger, R. Chemistry-enabled drug delivery (prodrugs): Recent progress and challenges. Drug Discov.
Today 2014, 19, 7987.
310. Huttunen, K.M.; Raunio, H.; Rautio, J. ProdrugsFrom Serendipity to Rational Design. Pharmacol. Rev. 2011, 63, 750–771.
311. Majumdar, S.; Duvvuri, S.; Mitra, A.K. Membrane transporter/receptor-targeted prodrug design: Strategies for human and vet-
erinary drug development. Adv. Drug Deliv. Rev. 2004, 56, 1437–1452.
312. Mazzaferro, S.; Bouchemal, K.; Ponchel, G. Oral delivery of anticancer drugs II: The prodrug strategy. Drug Discov. Today 2013,
18, 93–98.
313. Sievaen, E. Exploitation of bile acid transport systems in prodrug design. Molecules 2007, 12, 1859–1889.
314. Sinokrot, H.; Smerat, T.; Najjar, A.; Karaman, R. Advanced Prodrug Strategies in Nucleoside and Non-Nucleoside Antiviral
Agents: A Review of the Recent Five Years. Molecules 2017, 22, 1736.
315. Yang, C.; Tirucherai, G.S.; Mitra, A.K. Prodrug based optimal drug delivery via membrane transporter/receptor. Expert Opin.
Biol. Ther. 2001, 1, 159–175.
316. Zhang, Y.; Sun, J.; Sun, Y.; Wang, Y.; He, Z. Prodrug Design Targeting Intestinal PepT1 for Improved Oral Absorption: Design
and Performance. Curr. Drug Metab. 2013, 14, 675–687.
317. Minhas, G.S.; Newstead, S. Structural basis for prodrug recognition by the SLC15 family of proton-coupled peptide transporters.
Proc. Natl. Acad. Sci. USA 2019, 116, 804–809.
318. Minhas, G.S.; Newstead, S. Recent advances in understanding prodrug transport through the SLC15 family of proton-coupled
transporters. Biochem. Soc. Trans. 2020, 48, 337–346.
319. Huttunen, J.; Peltokangas, S.; Gynther, M.; Natunen, T.; Hiltunen, M.; Auriola, S.; Ruponen, M.; Vellonen, K.S.; Huttunen, K.M.
L-Type Amino Acid Transporter 1 (LAT1/Lat1)-Utilizing Prodrugs Can Improve the Delivery of Drugs into Neurons, Astrocytes
and Microglia. Sci. Rep. 2019, 9, 12860.
320. Huttunen, J.; Gynther, M.; Vellonen, K.S.; Huttunen, K.M. L-Type amino acid transporter 1 (LAT1)-utilizing prodrugs are car-
rier-selective despite having low affinity for organic anion transporting polypeptides (OATPs). Int. J. Pharm. 2019, 571, 118714.
321. Montaser, A.B.; Jarvinen, J.; Löffler, S.; Huttunen, J.; Auriola, S.; Lehtonen, M.; Jalkanen, A.; Huttunen, K.M. L-Type Amino
Acid Transporter 1 Enables the Efficient Brain Delivery of Small-Sized Prodrug across the Blood-Brain Barrier and into Human
and Mouse Brain Parenchymal Cells. ACS Chem. Neurosci. 2020, 11, 4301–4315.
322. Montaser, A.; Lehtonen, M.; Gynther, M.; Huttunen, K.M. L-Type Amino Acid Transporter 1-Utilizing Prodrugs of Ketoprofen
Can Efficiently Reduce Brain Prostaglandin Levels. Pharmaceutics 2020, 12, 344.
323. Peura, L.; Malmioja, K.; Huttunen, K.; Leppanen, J.; Hämäläinen, M.; Forsberg, M.M.; Gynther, M.; Rautio, J.; Laine, K. Design,
synthesis and brain uptake of LAT1-targeted amino acid prodrugs of dopamine. Pharm. Res. 2013, 30, 2523–2537.
324. Puris, E.; Gynther, M.; Huttunen, J.; Petsalo, A.; Huttunen, K.M. L-type amino acid transporter 1 utilizing prodrugs: How to
achieve effective brain delivery and low systemic exposure of drugs. J. Control. Release 2017, 261, 93–104.
325. Pardridge, W.M. Blood-brain barrier delivery. Drug Discov. Today 2007, 12, 54–61.
326. Agbabiaka, T.B.; Savovic, J.; Ernst, E. Methods for causality assessment of adverse drug reactions: A systematic review. Drug
Saf. 2008, 31, 21–37.
327. Davies, E.C.; Green, C.F.; Mottram, D.R.; Pirmohamed, M. Adverse drug reactions in hospitals: A narrative review. Curr. Drug
Saf. 2007, 2, 79–87.
328. Hazell, L.; Shakir, S.A. Under-reporting of adverse drug reactions : A systematic review. Drug Saf. 2006, 29, 385–396.
329. Ji, Z.L.; Han, L.Y.; Yap, C.W.; Sun, L.Z.; Chen, X.; Chen, Y.Z. Drug Adverse Reaction Target Database (DART) : Proteins related
to adverse drug reactions. Drug Saf. 2003, 26, 685–690.
330. King, C.; McKenna, A.; Farzan, N.; Vijverberg, S.J.; van der Schee, M.P.; Maitland-van der Zee, A.H.; Arianto, L.; Bisgaard, H.;
BØnnelykke, K.; Berce, V.; et al. Pharmacogenomic associations of adverse drug reactions in asthma: Systematic review and
research prioritisation. Pharm. J. 2020, 20, 621–628.
331. Miguel, A.; Azevedo, L.F.; Araújo, M.; Pereira, A.C. Frequency of adverse drug reactions in hospitalized patients: A systematic
review and meta-analysis. Pharmacoepidemiol. Drug Saf. 2012, 21, 1139–1154.
332. Osanlou, O.; Pirmohamed, M.; Daly, A.K. Pharmacogenetics of Adverse Drug Reactions. Adv. Pharmacol. 2018, 83, 155–190.
333. Pirmohamed, M. Pharmacogenetics of idiosyncratic adverse drug reactions. Handb. Exp. Pharmacol. 2010, 196, 477–491.
334. Pirmohamed, M. Personalized Pharmacogenomics: Predicting Efficacy and Adverse Drug Reactions. Annu. Rev. Genom. Hum.
Genet. 2014, 15, 349–370.
335. Sakiris, M.A.; Sawan, M.; Hilmer, S.N.; Awadalla, R.; Gnjidic, D. Prevalence of adverse drug events and adverse drug reactions
in hospital among older patients with dementia: A systematic review. Br. J. Clin. Pharmacol. 2021, 87, 375385.
336. Su, S.C.; Chung, W.H.; Hung, S.I. Digging Up the Human Genome: Current Progress in Deciphering Adverse Drug Reactions.
BioMed Res. Int. 2014, 2014, 824343.
Molecules 2021, 26, 5629 29 of 39
337. Wei, C.Y.; Lee, M.T.; Chen, Y.T. Pharmacogenomics of adverse drug reactions: Implementing personalized medicine. Hum. Mol.
Genet. 2012, 21, R58–R65.
338. Wilke, R.A.; Lin, D.W.; Roden, D.M.; Watkins, P.B.; Flockhart, D.; Zineh, I.; Giacomini, K.M.; Krauss, R.M. Identifying genetic
risk factors for serious adverse drug reactions: Current progress and challenges. Nat. Rev. Drug Discov. 2007, 6, 904–916.
339. Zhang, L.L.; Yang, S.; Wei, W.; Zhang, X.J. Genetic polymorphisms affect efficacy and adverse drug reactions of DMARDs in
rheumatoid arthritis. Pharm. Genom. 2014, 24, 531–538.
340. Zolk, O.; Fromm, M.F. Transporter-mediated drug uptake and efflux: Important determinants of adverse drug reactions. Clin.
Pharmacol. Ther. 2011, 89, 798–805.
341. Hakkarainen, K.M.; Andersson Sundell, K.; Petzold, M.; Hagg, S. Prevalence and perceived preventability of self-reported ad-
verse drug events--a population-based survey of 7099 adults. PLoS ONE 2013, 8, e73166.
342. Benfenati, E. In Silico Methods for Predicting Drug Toxicity; Springer: Berlin/Heidelber, Germany, 2016.
343. Gyllensten, H.; Jönsson, A.K.; Hakkarainen, K.M.; Svensson, S.; Hägg, S.; Rehnberg, C. Comparing Methods for Estimating
Direct Costs of Adverse Drug Events. Value Health 2017, 20, 1299–1310.
344. Giblin, K.A.; Basili, D.; Afzal, A.M.; Rosenbrier-Ribeiro, L.; Greene, N.; Barrett, I.; Hughes, S.J.; Bender, A. New Associations
between Drug-Induced Adverse Events in Animal Models and Humans Reveal Novel Candidate Safety Targets. Chem. Res.
Toxicol. 2021, 34, 438–451.
345. Insani, W.N.; Whittlesea, C.; Alwafi, H.; Man, K.K.C.; Chapman, S.; Wei, L. Prevalence of adverse drug reactions in the primary
care setting: A systematic review and meta-analysis. PLoS ONE 2021, 16, e0252161.
346. Tong, H.; Phan, N.V.T.; Nguyen, T.T.; Nguyen, D.V.; Vo, N.S.; Le, L. Review on Databases and Bioinformatic Approaches on
Pharmacogenomics of Adverse Drug Reactions. Pharmgenom. Pers. Med. 2021, 14, 61–75.
347. Cook, D.; Brown, D.; Alexander, R.; March, R.; Morgan, P.; Satterthwaite, G.; Pangalos, M.N. Lessons learned from the fate of
AstraZeneca’s drug pipeline: A five-dimensional framework. Nat. Rev. Drug Discov. 2014, 13, 419–431.
348. Morgan, P.; Brown, D.G.; Lennard, S.; Anderton, M.J.; Barrett, J.C.; Eriksson, U.; Fidock, M.; Hamren, B.; Johnson, A.; March, R.E.; et
al. Impact of a five-dimensional framework on R&D productivity at AstraZeneca. Nat. Rev. Drug Discov. 2018, 17, 167181.
349. Kola, I.; Landis, J. Can the pharmaceutical industry reduce attrition rates? Nat. Rev. Drug Discov. 2004, 3, 711–715.
350. Leeson, P.D.; Empfield, J.R. Reducing the risk of drug attrition associated with physicochemical properties. Annu. Rep. Med.
Chem 2010, 45, 393–407.
351. Muthas, D.; Boyer, S.; Hasselgren, C. A critical assessment of modeling safety-related drug attrition. MedChemComm 2013, 4,
1058–1065.
352. Waring, M.J.; Arrowsmith, J.; Leach, A.R.; Leeson, P.D.; Mandrell, S.; Owen, R.M.; Pairaudeau, G.; Pennie, W.D.; Pickett, S.D.;
Wang, J.; et al. An analysis of the attrition of drug candidates from four major pharmaceutical companies. Nat. Rev. Drug Discov.
2015, 14, 475–486.
353. Liu, Z.; Delavan, B.; Roberts, R.; Tong, W. Lessons Learned from Two Decades of Anticancer Drugs. Trends Pharmacol. Sci. 2017,
38, 852–872.
354. Williams, R.J. Biochemical Individuality; John Wiley: New York, NY, USA, 1956.
355. Kell, D.B.; Ryder, H.M.; Kaprelyants, A.S.; Westerhoff, H.V. Quantifying heterogeneity: Flow cytometry of bacterial cultures.
Antonie van Leeuwenhoek 1991, 60, 145–158.
356. Amur, S.; Zineh, I.; Abernethy, D.R.; Huang, S.M.; Lesko, L.J. Pharmacogenomics and adverse drug reactions. Pers. Med. 2010,
7, 633–642.
357. Clarke, J.D.; Cherrington, N.J. Genetics or environment in drug transport: The case of organic anion transporting polypeptides
and adverse drug reactions. Expert Opin. Drug Metab. Toxicol. 2012, 8, 349–360.
358. Daly, A.K. Using Genome-Wide Association Studies to Identify Genes Important in Serious Adverse Drug Reactions. Annu.
Rev. Pharmacol. Toxicol. 2012, 52, 21–35.
359. Giorgi, M.A.; Caroli, C.; Arazi, H.C.; Di Girolamo, G. Pharmacogenomics and adverse drug reactions: The case of statins. Expert
Opin. Pharmacother. 2011, 12, 1499–1509.
360. Kato, M.; Fukuda, T.; Wakeno, M.; Fukuda, K.; Okugawa, G.; Ikenaga, Y.; Yamashita, M.; Takekita, Y.; Nobuhara, K.; Azuma,
J.; et al. Effects of the serotonin type 2A, 3A and 3B receptor and the serotonin transporter genes on paroxetine and fluvoxamine
efficacy and adverse drug reactions in depressed Japanese patients. Neuropsychobiology 2006, 53, 186–195.
361. Mealey, K.L. Adverse drug reactions in veterinary patients associated with drug transporters. Vet. Clin. N. Am. Small Anim.
Pract. 2013, 43, 1067–1078.
362. Meyer, U.A. Pharmacogenetics and adverse drug reactions. Lancet 2000, 356, 1667–1671.
363. Rollason, V.; Lloret-Linares, C.; Lorenzini, K.I.; Daali, Y.; Gex-Fabry, M.; Piguet, V.; Besson, M.; Samer, C.; Desmeules, J. Evalu-
ation of Phenotypic and Genotypic Variations of Drug Metabolising Enzymes and Transporters in Chronic Pain Patients Facing
Adverse Drug Reactions or Non-Response to Analgesics: A Retrospective Study. J. Pers. Med. 2020, 10, 198.
364. Zhou, Y.; Zhang, G.Q.; Wei, Y.H.; Zhang, J.P.; Zhang, G.R.; Ren, J.X.; Duan, H.G.; Rao, Z.; Wu, X.A. The impact of drug trans-
porters on adverse drug reaction. Eur. J. Drug Metab. Pharm. 2013, 38, 7785.
365. Bodo, A.; Bakos, E.; Szeri, F.; Varadi, A.; Sarkadi, B. The role of multidrug transporters in drug availability, metabolism and
toxicity. Toxicol. Lett. 2003, 140–141, 133–143.
366. Ciarimboli, G. Role of organic cation transporters in drug-induced toxicity. Expert Opin. Drug Metab. Toxicol. 2011, 7, 159–174.
Molecules 2021, 26, 5629 30 of 39
367. Ciarimboli, G.; Holle, S.K.; Vollenbrocker, B.; Hagos, Y.; Reuter, S.; Burckhardt, G.; Bierer, S.; Herrmann, E.; Pavenstadt, H.;
Rossi, R.; et al. New Clues for Nephrotoxicity Induced by Ifosfamide: Preferential Renal Uptake via the Human Organic Cation
Transporter 2. Mol. Pharm. 2011, 8, 270–279.
368. Damaraju, V.L.; Mowles, D.; Yao, S.; Ng, A.; Young, J.D.; Cass, C.E.; Tong, Z. Role of human nucleoside transporters in the
uptake and cytotoxicity of azacitidine and decitabine. Nucleosides Nucleotides Nucleic Acids 2012, 31, 236–255.
369. Damaraju, V.L.; Scriver, T.; Mowles, D.; Kuzma, M.; Ryan, A.J.; Cass, C.E.; Sawyer, M.B. Erlotinib, Gefitinib, and Vandetanib
Inhibit Human Nucleoside Transporters and Protect Cancer Cells from Gemcitabine Cytotoxicity. Clin. Cancer Res. 2014, 20,
176–186.
370. DeGorter, M.K.; Xia, C.Q.; Yang, J.J.; Kim, R.B. Drug Transporters in Drug Efficacy and Toxicity. Annu. Rev. Pharmacol. Toxicol.
2012, 52, 249–273.
371. Elwi, A.N.; Damaraju, V.L.; Kuzma, M.L.; Baldwin, S.A.; Young, J.D.; Sawyer, M.B.; Cass, C.E. Human concentrative nucleoside
transporter 3 is a determinant of fludarabine transportability and cytotoxicity in human renal proximal tubule cell cultures.
Cancer Chemother. Pharmacol. 2009, 63, 289–301.
372. Feng, B.; El-Kattan, A.F.; Radi, Z.A. Renal transporters in drug disposition, drug-drug interactions, and nephrotoxicity. Curr.
Protoc. Toxicol. 2012, 53, 23.3.1–23.3.15.
373. Fischer, A.; Hoeger, S.J.; Stemmer, K.; Feurstein, D.J.; Knobeloch, D.; Nussler, A.; Dietrich, D.R. The role of organic anion trans-
porting polypeptides (OATPs/SLCOs) in the toxicity of different microcystin congeners in vitro: A comparison of primary hu-
man hepatocytes and OATP-transfected HEK293 cells. Toxicol. Appl. Pharmacol. 2010, 245, 920.
374. Fisel, P.; Renner, O.; Nies, A.T.; Schwab, M.; Schaeffeler, E. Solute carrier transporter and drug-related nephrotoxicity: The
impact of proximal tubule cell models for preclinical research. Expert Opin. Drug Metab. Toxicol. 2014, 10, 395–408.
375. Huang, K.M.; Hu, S.; Sparreboom, A. Drug transporters and anthracycline-induced cardiotoxicity. Pharmacogenomics 2018, 19,
883–888.
376. Jabir, R.S.; Naidu, R.; Annuar, M.A.; Ho, G.F.; Munisamy, M.; Stanslas, J. Pharmacogenetics of taxanes: Impact of gene poly-
morphisms of drug transporters on pharmacokinetics and toxicity. Pharmacogenomics 2012, 13, 1979–1988.
377. Kamal, M.A.; Keep, R.F.; Smith, D.E. Role and relevance of PEPT2 in drug disposition, dynamics, and toxicity. Drug Metab.
Pharm. 2008, 23, 236–242.
378. Krajcsi, P.; Vereczkey, L. Transporter-drug interactions and transporter-mediated toxicity in the liver/hepatocyte. Preface. Drug
Metab. Rev. 2010, 42, 379.
379. Li, S.; Chen, Y.; Zhang, S.; More, S.S.; Huang, X.; Giacomini, K.M. Role of organic cation transporter 1, OCT1 in the pharmaco-
kinetics and toxicity of cis-diammine(pyridine)chloroplatinum(II) and oxaliplatin in mice. Pharm. Res. 2011, 28, 610–625.
380. Mor, A.L.; Kaminski, T.W.; Karbowska, M.; Pawlak, D. New Insight into Organic Anion Transporters from the Perspective of
Potentially Important Interactions and Drugs Toxicity. J. Physiol. Pharmacol. 2018, 69, 307–324.
381. More, S.S.; Li, S.; Yee, S.W.; Chen, L.; Xu, Z.; Jablons, D.M.; Giacomini, K.M. Organic cation transporters modulate the uptake
and cytotoxicity of picoplatin, a third-generation platinum analogue. Mol. Cancer Ther. 2010, 9, 1058–1069.
382. Nakamura, T.; Yonezawa, A.; Hashimoto, S.; Katsura, T.; Inui, K. Disruption of multidrug and toxin extrusion MATE1 potenti-
ates cisplatin-induced nephrotoxicity. Biochem. Pharmacol. 2010, 80, 1762–1767.
383. Niemi, M. Transporter pharmacogenetics and statin toxicity. Clin. Pharmacol. Ther. 2010, 87, 130–133.
384. Parmar, S.; Seeringer, A.; Denich, D.; Gärtner, F.; Pitterle, K.; Syrovets, T.; Ohmle, B.; Stingl, J.C. Variability in transport and
biotransformation of cytarabine is associated with its toxicity in peripheral blood mononuclear cells. Pharmacogenomics 2011, 12,
503–514.
385. Schuetz, J.D.; Swaan, P.W.; Tweedie, D.J. The role of transporters in toxicity and disease. Drug Metab. Dispos. 2014, 42, 541–545.
386. Sprowl, J.A.; Ciarimboli, G.; Lancaster, C.S.; Giovinazzo, H.; Gibson, A.A.; Du, G.Q.; Janke, L.J.; Cavaletti, G.; Shields, A.F.;
Sparreboom, A. Oxaliplatin-induced neurotoxicity is dependent on the organic cation transporter OCT2. Proc. Natl. Acad. Sci.
USA 2013, 110, 11199–11204.
387. Szakács, G.; Váradi, A.; Özvegy-Laczka, C.; Sarkadi, B. The role of ABC transporters in drug absorption, distribution, metabo-
lism, excretion and toxicity (ADME-Tox). Drug Discov. Today 2008, 13, 379–393.
388. Visscher, H.; Rassekh, S.R.; Sandor, G.S.; Caron, H.N.; van Dalen, E.C.; Kremer, L.C.; van der Pal, H.J.; Rogers, P.C.; Rieder, M.J.;
Carleton, B.C.; et al. Genetic variants in SLC22A17 and SLC22A7 are associated with anthracycline-induced cardiotoxicity in
children. Pharmacogenomics 2015, 16, 1065–1076.
389. Wang, L.; Sweet, D.H. Renal organic anion transporters (SLC22 family): Expression, regulation, roles in toxicity, and impact on
injury and disease. AAPS J. 2013, 15, 53–69.
390. Zhang, S.; Lovejoy, K.S.; Shima, J.E.; Lagpacan, L.L.; Shu, Y.; Lapuk, A.; Chen, Y.; Komori, T.; Gray, J.W.; Chen, X.; et al. Organic
cation transporters are determinants of oxaliplatin cytotoxicity. Cancer Res. 2006, 66, 8847–8857.
391. Evers, R.; Piquette-Miller, M.; Polli, J.W.; Russel, F.G.M.; Sprowl, J.A.; Tohyama, K.; Ware, J.A.; de Wildt, S.N.; Xie, W.; Brouwer,
K.L.R.; International Transporter Consortium. Disease-Associated Changes in Drug Transporters May Impact the Pharmacoki-
netics and/or Toxicity of Drugs: A White Paper From the International Transporter Consortium. Clin. Pharmacol. Ther. 2018, 104,
900–915.
392. Hu, S.; Sprowl, J.A. Strategies to Reduce Solute Carrier-Mediated Toxicity. Clin. Pharmacol. Ther. 2018, 104, 799–802.
Molecules 2021, 26, 5629 31 of 39
393. Chu, X.; Liao, M.; Shen, H.; Yoshida, K.; Zur, A.A.; Arya, V.; Galetin, A.; Giacomini, K.M.; Hanna, I.; Kusuhara, H.; et al. Inter-
national Transporter Consortium, Clinical Probes and Endogenous Biomarkers as Substrates for Transporter Drug-Drug Inter-
action Evaluation: Perspectives From the International Transporter Consortium. Clin. Pharmacol. Ther. 2018, 104, 836–864.
394. Anderson, J.T.; Huang, K.M.; Lustberg, M.B.; Sparreboom, A.; Hu, S. Solute carrier transportome in chemotherapy-induced
adverse drug reactions. In Reviews of Physiology Biochemistry and Pharmacology; Springer: Berlin/Heidelberg, Germany, 2020.
https://doi.org/10.1007/112_2020_30.
395. O’Neill, J. Vaccines and Alternative Approaches: Reducing Our Dependence on Antimicrobials. The Review on Antimicrobial Resistance;
The Wellcome Trust and HM Government: London, UK, 2016.
396. O’Neill, J. Tackling Drug-Resistant Infections Globally: An. Overview of Our Work. The Review on Antimicrobial Resistance; The Well-
come Trust and HM Government: London, UK, 2016.
397. Piddock, L.; Garneau-Tsodikova, S.; Garner, C. Ask the experts: How to curb antibiotic resistance and plug the antibiotics gap?
Future Med. Chem. 2016, 8, 1027–1032.
398. Crofts, T.S.; Gasparrini, A.J.; Dantas, G. Next-generation approaches to understand and combat the antibiotic resistome. Nat.
Rev. Microbiol. 2017, 15, 422–434.
399. Wells, V.; Piddock, L.J.V. Addressing antimicrobial resistance in the UK and Europe. Lancet Infect. Dis. 2017, 17, 1230–1231.
400. Baker, S.; Thomson, N.; Weill, F.X.; Holt, K.E. Genomic insights into the emergence and spread of antimicrobial-resistant bacte-
rial pathogens. Science 2018, 360, 733–738.
401. Annunziato, G. Strategies to Overcome Antimicrobial Resistance (AMR) Making Use of Non-Essential Target Inhibitors: A
Review. Int. J. Mol. Sci. 2019, 20, 5844.
402. Dougan, G.; Dowson, C.; Overington, J.; Next Generation Antibiotic Discovery Symposium, P. Meeting the discovery challenge
of drug-resistant infections: Progress and focusing resources. Drug Discov. Today 2019, 24, 452–461.
403. Roope, L.S.J.; Smith, R.D.; Pouwels, K.B.; Buchanan, J.; Abel, L.; Eibich, P.; Butler, C.C.; Tan, P.S.; Walker, A.S.; Robotham, J.V.; et al.
The challenge of antimicrobial resistance: What economics can contribute. Science 2019, 364, eaau4679. https://doi.org/10.1126/sci-
ence.aau4679.
404. Turner, N.A.; Sharma-Kuinkel, B.K.; Maskarinec, S.A.; Eichenberger, E.M.; Shah, P.P.; Carugati, M.; Holland, T.L.; Fowler, V.G.
Methicillin-resistant Staphylococcus aureus: An overview of basic and clinical research. Nat. Rev. Microbiol. 2019, 17, 203–218.
405. Antwi, A.N.; Stewart, A.; Crosbie, M. Fighting antibiotic resistance: A narrative review of public knowledge, attitudes, and
perceptions of antibiotics use. Perspect. Public Health 2020, 140, 338–350.
406. Diallo, O.O.; Baron, S.A.; Abat, C.; Colson, P.; Chaudet, H.; Rolain, J.M. Antibiotic resistance surveillance systems: A review. J.
Glob. Antimicrob. Resist. 2020, 23, 430–438.
407. Kavvas, E.S.; Yang, L.; Monk, J.M.; Heckmann, D.; Palsson, B.O. A biochemically-interpretable machine learning classifier for
microbial GWAS. Nat. Commun. 2020, 11, 2580.
408. Jit, M.; Ng, D.H.L.; Luangasanatip, N.; Sandmann, F.; Atkins, K.E.; Robotham, J.V.; Pouwels, K.B. Quantifying the economic
cost of antibiotic resistance and the impact of related interventions: Rapid methodological review, conceptual framework and
recommendations for future studies. BMC Med. 2020, 18, 38.
409. Khan, J.; Tarar, S.M.; Gul, I.; Nawaz, U.; Arshad, M. Challenges of antibiotic resistance biofilms and potential combating strat-
egies: A review. 3 Biotech 2021, 11, 169.
410. De Kraker, M.E.A.; Stewardson, A.J.; Harbarth, S. Will 10 Million People Die a Year due to Antimicrobial Resistance by 2050?
PLoS Med. 2016, 13, e1002184.
411. Hoffman, P.S. Antibacterial Discovery: 21st Century Challenges. Antibiotics 2020, 9, 213.
412. Lu, J.; Sheldenkar, A.; Lwin, M.O. A decade of antimicrobial resistance research in social science fields: A scientometric review.
Antimicrob. Resist. Infect. Control. 2020, 9, 178.
413. De Kraker, M.E.A.; Lipsitch, M. Burden of Antimicrobial Resistance: Compared to What? Epidemiol. Rev. 2021,
https://doi.org/10.1093/epirev/mxab001.
414. Schuts, E.C.; Boyd, A.; Muller, A.E.; Mouton, J.W.; Prins, J.M. The Effect of Antibiotic Restriction Programs on Prevalence of
Antimicrobial Resistance: A Systematic Review and Meta-Analysis. Open Forum Infect. Dis. 2021, 8, ofab070.
415. Ali, J.; Rafiq, Q.A.; Ratcliffe, E. Antimicrobial resistance mechanisms and potential synthetic treatments. Future Sci. OA 2018, 4,
FSO290.
416. Blair, J.M.A.; Webber, M.A.; Baylay, A.J.; Ogbolu, D.O.; Piddock, L.J.V. Molecular mechanisms of antibiotic resistance. Nat. Rev.
Microbiol. 2015, 13, 42–51.
417. Abouelhassan, Y.; Garrison, A.T.; Yang, H.; Chavez-Riveros, A.; Burch, G.M.; Huigens, R.W., 3rd Recent Progress in Natural-
Product-Inspired Programs Aimed To Address Antibiotic Resistance and Tolerance. J. Med. Chem. 2019, 62, 7618–7642.
418. Silver, L.L. Challenges of antibacterial discovery. Clin. Microbiol. Rev. 2011, 24, 71–109.
419. White, T.A.; Kell, D.B. Comparative genomic assessment of novel broad-spectrum targets for antibacterial drugs. Comp. Funct.
Genomics 2004, 5, 304–327.
420. Payne, D.J.; Gwynn, M.N.; Holmes, D.J.; Pompliano, D.L. Drugs for bad bugs: Confronting the challenges of antibacterial dis-
covery. Nat. Rev. Drug Discov. 2007, 6, 29–40.
421. Yılmaz, Ç.; Özcengiz, G. Antibiotics: Pharmacokinetics, toxicity, resistance and multidrug efflux pumps. Biochem. Pharmacol.
2017, 133, 43–62.
Molecules 2021, 26, 5629 32 of 39
422. Ahmad, I.; Nawaz, N.; Dermani, F.K.; Kohlan, A.K.; Saidijam, M.; Patching, S.G. Bacterial Multidrug Efflux Proteins: A Major
Mechanism of Antimicrobial Resistance. Curr. Drug Targets 2018, 19, 113.
423. Chopra, I.; Roberts, M. Tetracycline antibiotics: Mode of action, applications, molecular biology, and epidemiology of bacterial
resistance. Microbiol Mol. Biol Rev. 2001, 65, 232–260.
424. Phillips-Jones, M.K.; Harding, S.E. Antimicrobial resistance (AMR) nanomachines-mechanisms for fluoroquinolone and glyco-
peptide recognition, efflux and/or deactivation. Biophys. Rev. 2018, 10, 347–362.
425. Putman, M.; van Veen, H.W.; Konings, W.N. Molecular properties of bacterial multidrug transporters. Microbiol. Mol. Biol. Rev.
2000, 64, 672–693.
426. Piddock, L.J. Multidrug-resistance efflux pumpsNot just for resistance. Nat. Rev. Microbiol 2006, 4, 629–636.
427. Piddock, L.J.V. Clinically relevant chromosomally encoded multidrug resistance efflux pumps in bacteria. Clin. Microbiol. Rev.
2006, 19, 382–402.
428. Piddock, L.J.V. The 2019 Garrod Lecture: MDR efflux in Gram-negative bacteria-how understanding resistance led to a new
tool for drug discovery. J. Antimicrob. Chemother. 2019, 74, 3128–3134.
429. Nikaido, H. Multidrug resistance in bacteria. Annu. Rev. Biochem. 2009, 78, 119–146.
430. Li, X.Z.; Plésiat, P.; Nikaido, H. The challenge of efflux-mediated antibiotic resistance in Gram-negative bacteria. Clin. Microbiol.
Rev. 2015, 28, 337–418.
431. Brown, E.D.; Wright, G.D. Antibacterial drug discovery in the resistance era. Nature 2016, 529, 336–343.
432. Du, D.; Wang-Kan, X.; Neuberger, A.; van Veen, H.W.; Pos, K.M.; Piddock, L.J.V.; Luisi, B.F. Multidrug efflux pumps: Structure,
function and regulation. Nat. Rev. Microbiol. 2018, 16, 523–539.
433. Zwama, M.; Yamaguchi, A. Molecular mechanisms of AcrB-mediated multidrug export. Res. Microbiol. 2018, 169, 372–383.
434. Zgurskaya, H.I.; Rybenkov, V.V.; Krishnamoorthy, G.; Leus, I.V. Trans-envelope multidrug efflux pumps of Gram-negative
bacteria and their synergism with the outer membrane barrier. Res. Microbiol. 2018, 169, 351–356.
435. Zgurskaya, H.I.; Walker, J.K.; Parks, J.M.; Rybenkov, V.V. Multidrug Efflux Pumps and the Two-Faced Janus of Substrates and
Inhibitors. Acc. Chem. Res. 2021, 54, 930–939.
436. De Oliveira, D.M.P.; Forde, B.M.; Kidd, T.J.; Harris, P.N.A.; Schembri, M.A.; Beatson, S.A.; Paterson, D.L.; Walker, M.J. Antimi-
crobial Resistance in ESKAPE Pathogens. Clin. Microbiol. Rev. 2020, 33, e00181-19.
437. Zhao, S.; Adamiak, J.W.; Bonifay, V.; Mehla, J.; Zgurskaya, H.I.; Tan, D.S. Defining new chemical space for drug penetration
into Gram-negative bacteria. Nat. Chem. Biol. 2020, 16, 1293–1302.
438. Rahman, T.; Yarnall, B.; Doyle, D.A. Efflux drug transporters at the forefront of antimicrobial resistance. Eur. Biophys. J. 2017,
46, 647–653.
439. Perlin, M.H.; Andrews, J.; Toh, S.S. Essential Letters in the Fungal Alphabet: ABC and MFS Transporters and Their Roles in
Survival and Pathogenicity. Adv. Genet. 2014, 85, 201–253.
440. Capela, R.; Moreira, R.; Lopes, F. An Overview of Drug Resistance in Protozoal Diseases. Int. J. Mol. Sci. 2019, 20, 5748.
441. De Koning, H.P. The Drugs of Sleeping Sickness: Their Mechanisms of Action and Resistance, and a Brief History. Trop Med.
Infect. Dis 2020, 5, 14.
442. Spengler, G.; Kincses, A.; Gajdács, M.; Amaral, L. New roads leading to old destinations: Efflux pumps as targets to reverse
multidrug resistance in bacteria. Molecules 2017, 22, 468.
443. Krishnamoorthy, G.; Leus, I.V.; Weeks, J.W.; Wolloscheck, D.; Rybenkov, V.V.; Zgurskaya, H.I. Synergy between Active Efflux
and Outer Membrane Diffusion Defines Rules of Antibiotic Permeation into Gram-Negative Bacteria. mBio 2017, 8, e01172-17.
444. Rybenkov, V.V.; Zgurskaya, H.I.; Ganguly, C.; Leus, I.V.; Zhang, Z.; Moniruzzaman, M. The Whole Is Bigger than the Sum of
Its Parts: Drug Transport in the Context of Two Membranes with Active Efflux. Chem. Rev. 2021, 121, 5597–5631.
445. Nikaido, H.; Pagès, J.M. Broad-specificity efflux pumps and their role in multidrug resistance of Gram-negative bacteria. FEMS
Microbiol. Rev. 2012, 36, 340–363.
446. Bolla, J.M.; Alibert-Franco, S.; Handzlik, J.; Chevalier, J.; Mahamoud, A.; Boyer, G.; Kieć-Kononowicz, K.; Pagès, J.M. Strategies
for bypassing the membrane barrier in multidrug resistant Gram-negative bacteria. FEBS Lett. 2011, 585, 1682–1690.
447. Otrębska-Machaj, E.; Chevalier, J.; Handzlik, J.; Szymańska, E.; Schabikowski, J.; Boyer, G.; Bolla, J.M.; Kieć-Kononowicz, K.;
Pagès, J.M.; Alibert, S. Efflux Pump Blockers in Gram-Negative Bacteria: The New Generation of Hydantoin Based-Modulators
to Improve Antibiotic Activity. Front. Microbiol. 2016, 7, 622.
448. Vila, J.; Marti, S.; Sánchez-spedes, J. Porins, efflux pumps and multidrug resistance in Acinetobacter baumannii. J. Antimicrob.
Chemother. 2007, 59, 1210–1215.
449. Schweizer, H.P. Understanding efflux in Gram-negative bacteria: Opportunities for drug discovery. Expert Opin. Drug Discov.
2012, 7, 633–642.
450. Kourtesi, C.; Ball, A.R.; Huang, Y.Y.; Jachak, S.M.; Vera, D.M.; Khondkar, P.; Gibbons, S.; Hamblin, M.R.; Tegos, G.P. Microbial
efflux systems and inhibitors: Approaches to drug discovery and the challenge of clinical implementation. Open Microbiol. J.
2013, 7, 34–52.
451. Venter, H.; Mowla, R.; Ohene-Agyei, T.; Ma, S.T. RND-type drug efflux pumps from Gram-negative bacteria: Molecular mech-
anism and inhibition. Front. Microbiol. 2015, 6, 377.
452. Alibert, S.; N'gompaza Diarra, J.; Hernandez, J.; Stutzmann, A.; Fouad, M.; Boyer, G.; Pagès, J.M. Multidrug efflux pumps and their
role in antibiotic and antiseptic resistance: A pharmacodynamic perspective. Expert Opin. Drug Met. Toxicol. 2017, 13, 301309.
453. Blair, J.M.A.; Piddock, L.J.V. How to measure export via bacterial multidrug resistance efflux pumps. MBio 2016, 7, e00840-16.
Molecules 2021, 26, 5629 33 of 39
454. Colclough, A.L.; Alav, I.; Whittle, E.E.; Pugh, H.L.; Darby, E.M.; Legood, S.W.; McNeil, H.E.; Blair, J.M. RND efflux pumps in
Gram-negative bacteria; regulation, structure and role in antibiotic resistance. Future Microbiol. 2020, 15, 143–157.
455. Seukep, A.J.; Kuete, V.; Nahar, L.; Sarker, S.D.; Guo, M. Plant-derived secondary metabolites as the main source of efflux pump
inhibitors and methods for identification. J. Pharm. Anal. 2020, 10, 277–290.
456. Slipski, C.J.; Zhanel, G.G.; Bay, D.C. Biocide Selective TolC-Independent Efflux Pumps in Enterobacteriaceae. J. Membr. Biol.
2018, 251, 15–33.
457. Altinöz, E.; Altuner, E.M. Antibiotic Resistance and Efflux Pumps. Int. J. Innov. Res. Rev. 2019, 3, 19.
458. Zhou, Y.; Joubran, C.; Miller-Vedam, L.; Isabella, V.; Nayar, A.; Tentarelli, S.; Miller, A. Thinking outside the “bug”: A unique
assay to measure intracellular drug penetration in gram-negative bacteria. Anal. Chem. 2015, 87, 3579–3584.
459. Krishnamoorthy, G.; Wolloscheck, D.; Weeks, J.W.; Croft, C.; Rybenkov, V.V.; Zgurskaya, H.I. Breaking the Permeability Barrier
of Escherichia coli by Controlled Hyperporination of the Outer Membrane. Antimicrob. Agents Chemother. 2016, 60, 7372–7381.
460. Coldham, N.G.; Webber, M.; Woodward, M.J.; Piddock, L.J.V. A 96-well plate fluorescence assay for assessment of cellular
permeability and active efflux in Salmonella enterica serovar Typhimurium and Escherichia coli. J. Antimicrob. Chemother. 2010, 65,
1655–1663.
461. Six, D.A.; Krucker, T.; Leeds, J.A. Advances and challenges in bacterial compound accumulation assays for drug discovery.
Curr. Opin. Chem. Biol. 2018, 44, 915.
462. Widya, M.; Pasutti, W.D.; Sachdeva, M.; Simmons, R.L.; Tamrakar, P.; Krucker, T.; Six, D.A. Development and Optimization of
a Higher-Throughput Bacterial Compound Accumulation Assay. ACS Infect. Dis. 2019, 5, 394–405.
463. Alav, I.; Kobylka, J.; Kuth, M.S.; Pos, K.M.; Picard, M.; Blair, J.M.A.; Bavro, V.N. Structure, Assembly, and Function of Tripartite
Efflux and Type 1 Secretion Systems in Gram-Negative Bacteria. Chem. Rev. 2021, 121, 5479–5596.
464. Vergalli, J.; Bodrenko, I.V.; Masi, M.; Moynie, L.; Acosta-Gutiérrez, S.; Naismith, J.H.; Davin-Regli, A.; Ceccarelli, M.; van den
Berg, B.; Winterhalter, M.; et al. Porins and small-molecule translocation across the outer membrane of Gram-negative bacteria.
Nat. Rev. Microbiol. 2020, 18, 164–176.
465. Masi, M.; Winterhalter, M.; Pages, J.M. Outer Membrane Porins. Subcell. Biochem. 2019, 92, 79–123.
466. Masi, M.; Réfrégiers, M.; Pos, K.M.; Pagès, J.M. Mechanisms of envelope permeability and antibiotic influx and efflux in Gram-
negative bacteria. Nat. Microbiol. 2017, 2, 17001.
467. Galocha, M.; Costa, I.V.; Teixeira, M.C. Carrier-Mediated Drug Uptake in Fungal Pathogens. Genes 2020, 11, 1324.
468. Jantsch, J.; Chikkaballi, D.; Hensel, M. Cellular aspects of immunity to intracellular Salmonella enterica. Immunol. Rev. 2011, 240,
185–195.
469. Price, C.T.D.; Al-Quadan, T.; Santic, M.; Rosenshine, I.; Abu Kwaik, Y. Host proteasomal degradation generates amino acids
essential for intracellular bacterial growth. Science 2011, 334, 1553–1557.
470. Bravo-Santano, N.; Ellis, J.K.; Mateos, L.M.; Calle, Y.; Keun, H.C.; Behrends, V.; Letek, M. Intracellular Staphylococcus aureus
Modulates Host Central Carbon Metabolism To Activate Autophagy. mSphere 2018, 3, e00374-18.
471. Thwaites, G.E.; Gant, V. Are bloodstream leukocytes Trojan Horses for the metastasis of Staphylococcus aureus? Nat. Rev. Micro-
biol. 2011, 9, 215–222.
472. Kell, D.B.; Potgieter, M.; Pretorius, E. Individuality, phenotypic differentiation, dormancy and ‘persistence’ in culturable bacte-
rial systems: Commonalities shared by environmental, laboratory, and clinical microbiology. F1000Research 2015, 4, 179.
473. Kell, D.B.; Pretorius, E. On the translocation of bacteria and their lipopolysaccharides between blood and peripheral locations
in chronic, inflammatory diseases: The central roles of LPS and LPS-induced cell death Integr. Biol. 2015, 7, 1339–1377.
474. Casadevall, A. Evolution of intracellular pathogens. Annu. Rev. Microbiol. 2008, 62, 19–33.
475. Silva, M.T.; Pestana, N.T. The in vivo extracellular life of facultative intracellular bacterial parasites: Role in pathogenesis. Im-
munobiology 2013, 218, 325–337.
476. Leon-Sicairos, N.; Reyes-Cortes, R.; Guadrón-Llanos, A.M.; Madueña-Molina, J.; Leon-Sicairos, C.; Canizalez-Román, A. Strat-
egies of Intracellular Pathogens for Obtaining Iron from the Environment. Biomed. Res. Int 2015, 2015, 476534.
477. von Bargen, K.; Gorvel, J.P.; Salcedo, S.P. Internal affairs: Investigating the Brucella intracellular lifestyle. FEMS Microbiol. Rev.
2012, 36, 533–562.
478. McClure, E.E.; Chavez, A.S.O.; Shaw, D.K.; Carlyon, J.A.; Ganta, R.R.; Noh, S.M.; Wood, D.O.; Bavoil, P.M.; Brayton, K.A.;
Martinez, J.J.; et al. Engineering of obligate intracellular bacteria: Progress, challenges and paradigms. Nat. Rev. Microbiol. 2017,
15, 544–558.
479. Garzoni, C.; Kelley, W.L. Staphylococcus aureus: New evidence for intracellular persistence. Trends Microbiol. 2009, 17, 59–65.
480. Garzoni, C.; Kelley, W.L. Return of the Trojan horse: Intracellular phenotype switching and immune evasion by Staphylococcus
aureus. EMBO Mol. Med. 2011, 3, 115–117.
481. Takeuchi, H.; Furuta, N.; Morisaki, I.; Amano, A. Exit of intracellular Porphyromonas gingivalis from gingival epithelial cells is
mediated by endocytic recycling pathway. Cell Microbiol. 2011, 13, 677–691.
482. Proal, A.D.; Marshall, T.G. Re-framing the theory of autoimmunity in the era of the microbiome: Persistent pathogens, autoan-
tibodies, and molecular mimicry. Discov. Med. 2018, 140, 299–308.
483. Proal, A.D.; VanElzakker, M.B. Pathogens Hijack Host Cell Metabolism: Intracellular Infection as a Driver of the Warburg Effect
in Cancer and Other Chronic Inflammatory Conditions. Immunometabolism 2020, 3, e210003.
484. Hunstad, D.A.; Justice, S.S. Intracellular lifestyles and immune evasion strategies of uropathogenic Escherichia coli. Annu. Rev.
Microbiol. 2010, 64, 203–221.
Molecules 2021, 26, 5629 34 of 39
485. Potgieter, M.; Bester, J.; Kell, D.B.; Pretorius, E. The dormant blood microbiome in chronic, inflammatory diseases. FEMS Mi-
crobiol. Rev. 2015, 39, 567–591.
486. Fullam, E.; Young, R.J. Physicochemical properties and Mycobacterium tuberculosis transporters: Keys to efficacious antituber-
cular drugs? RSC Med. Chem. 2021, 12, 43.
487. Prideaux, B.; Via, L.E.; Zimmerman, M.D.; Eum, S.; Sarathy, J.; O’Brien, P.; Chen, C.; Kaya, F.; Weiner, D.M.; Chen, P.Y.; et al.
The association between sterilizing activity and drug distribution into tuberculosis lesions. Nat. Med. 2015, 21, 1223–1227.
488. Santucci, P.; Greenwood, D.J.; Fearns, A.; Chen, K.; Jiang, H.; Gutierrez, M.G. Intracellular localisation of Mycobacterium tuber-
culosis affects efficacy of the antibiotic pyrazinamide. Nat. Commun. 2021, 12, 3816.
489. Smith, D.E.; Clémençon, B.; Hediger, M.A. Proton-coupled oligopeptide transporter family SLC15: Physiological, pharmaco-
logical and pathological implications. Mol. Aspects Med. 2013, 34, 323–336.
490. Samsudin, F.; Parker, J.L.; Sansom, M.S.P.; Newstead, S.; Fowler, P.W. Accurate Prediction of Ligand Affinities for a Proton-
Dependent Oligopeptide Transporter. Cell Chem. Biol. 2016, 23, 299–309.
491. Richter, M.F.; Hergenrother, P.J. The challenge of converting Gram-positive-only compounds into broad-spectrum antibiotics.
Ann. N. Y. Acad. Sci. 2019, 1435, 18–38.
492. Parker, E.N.; Drown, B.S.; Geddes, E.J.; Lee, H.Y.; Ismail, N.; Lau, G.W.; Hergenrother, P.J. Implementation of permeation rules
leads to a FabI inhibitor with activity against Gram-negative pathogens. Nat. Microbiol. 2020, 5, 67–75.
493. Perlmutter, S.J.; Geddes, E.J.; Drown, B.S.; Motika, S.E.; Lee, M.R.; Hergenrother, P.J. Compound Uptake into E. coli Can Be
Facilitated by N-Alkyl Guanidiniums and Pyridiniums. ACS Infect. Dis. 2021, 7, 162–173.
494. Muñoz, K.A.; Hergenrother, P.J. Facilitating Compound Entry as a Means to Discover Antibiotics for Gram-Negative Bacteria.
Acc. Chem. Res. 2021, 54, 1322–1333.
495. Davis, T.D.; Gerry, C.J.; Tan, D.S. General platform for systematic quantitative evaluation of small-molecule permeability in
bacteria. ACS Chem. Biol. 2014, 9, 2535–2544.
496. Aires, J.R.; Nikaido, H. Aminoglycosides are captured from both periplasm and cytoplasm by the AcrD multidrug efflux trans-
porter of Escherichia coli. J. Bacteriol. 2005, 187, 1923–1929.
497. Prabhala, B.K.; Aduri, N.G.; Sharma, N.; Shaheen, A.; Sharma, A.; Iqbal, M.; Hansen, P.R.; Brasen, C.; Gajhede, M.; Rahman, M.;
et al. The prototypical proton-coupled oligopeptide transporter YdgR from Escherichia coli facilitates chloramphenicol uptake
into bacterial cells. J. Biol. Chem. 2018, 293, 1007–1017.
498. Chen, J.M.; Uplekar, S.; Gordon, S.V.; Cole, S.T. A point mutation in cycA partially contributes to the D-cycloserine resistance
trait of Mycobacterium bovis BCG vaccine strains. PLoS ONE 2012, 7, e43467.
499. Chapeland-Leclerc, F.; Bouchoux, J.; Goumar, A.; Chastin, C.; Villard, J.; Noël, T. Inactivation of the FCY2 gene encoding purine-
cytosine permease promotes cross-resistance to flucytosine and fluconazole in Candida lusitaniae. Antimicrob. Agents Chemother.
2005, 49, 3101–3108.
500. Chen, Y.N.; Lo, H.J.; Wu, C.C.; Ko, H.C.; Chang, T.P.; Yang, Y.L. Loss of heterozygosity of FCY2 leading to the development of
flucytosine resistance in Candida tropicalis. Antimicrob. Agents Chemother. 2011, 55, 2506–2514.
501. Takahata, S.; Ida, T.; Hiraishi, T.; Sakakibara, S.; Maebashi, K.; Terada, S.; Muratani, T.; Matsumoto, T.; Nakahama, C.; Tomono, K.
Molecular mechanisms of fosfomycin resistance in clinical isolates of Escherichia coli. Int. J. Antimicrob. Agents 2010, 35, 333337.
502. Ballestero-Téllez, M.; Docobo-Pérez, F.; Portillo-Calderón, I.; Rodríguez-Martínez, J.M.; Racero, L.; Ramos-Guelfo, M.S.; Bláz-
quez, J.; Rodríguez-Baño, J.; Pascual, A. Molecular insights into fosfomycin resistance in Escherichia coli. J. Antimicrob. Chemother.
2017, 72, 1303–1309.
503. Mistry, A.; Warren, M.S.; Cusick, J.K.; Karkhoff-Schweizer, R.R.; Lomovskaya, O.; Schweizer, H.P. High-level pacidamycin re-
sistance in Pseudomonas aeruginosa is mediated by an opp oligopeptide permease encoded by the opp-fabI operon. Antimicrob.
Agents Chemother. 2013, 57, 5565–5571.
504. Pletzer, D.; Braun, Y.; Dubiley, S.; Lafon, C.; Kohler, T.; Page, M.G.; Mourez, M.; Severinov, K.; Weingart, H. The Pseudomonas
aeruginosa PA14 ABC Transporter NppA1A2BCD Is Required for Uptake of Peptidyl Nucleoside Antibiotics. J. Bacteriol. 2015,
197, 2217–2228.
505. De Koning, H.P. Uptake of pentamidine in Trypanosoma brucei brucei is mediated by three distinct transporters: Implications for
cross-resistance with arsenicals. Mol. Pharmacol. 2001, 59, 586–592.
506. Tindall, S.M.; Vallières, C.; Lakhani, D.H.; Islahudin, F.; Ting, K.N.; Avery, S.V. Heterologous Expression of a Novel Drug
Transporter from the Malaria Parasite Alters Resistance to Quinoline Antimalarials. Sci. Rep. 2018, 8, 2464.
507. Chopra, I. Molecular mechanisms involved in the transport of antibiotics into bacteria. Parasitology 1988, 96, S25–S44.
508. Chopra, I. Penetration of antibiotics to their target sites. J. Antimicrob. Chemother. 1990, 26, 607–609.
509. Delespaux, V.; de Koning, H.P. Transporters in antiparasitic drug development and resistance. In Antiparasitic and Antibacterial
Drug Discovery: Trypanosomatidae; Flohe, L., Koch, O., Jäger, T., Eds.; Wiley-Blackwell: London, UK, 2013, 335-349.
510. McMurry, L.; Levy, S.B. Two transport systems for tetracycline in sensitive Escherichia coli: Critical role for an initial rapid uptake
system insensitive to energy inhibitors. Antimicrob. Agents Chemother. 1978, 14, 201–209.
511. Smith, M.C.; Chopra, I. Energetics of tetracycline transport into Escherichia coli. Antimicrob. Agents Chemother. 1984, 25, 446–449.
512. Hutson, M. The language machines. Nature 2021, 591, 22–25.
513. Shrivastava, A.D.; Swainston, N.; Samanta, S.; Roberts, I.; Wright Muelas, M.; Kell, D.B. MassGenie: A transformer-based deep
learning method for identifying small molecules from their mass spectra. bioRxiv 2021,
https://doi.org/10.1101/2021.06.25.449969.
Molecules 2021, 26, 5629 35 of 39
514. Vermaas, J.V.; Trebesch, N.; Mayne, C.G.; Thangapandian, S.; Shekhar, M.; Mahinthichaichan, P.; Baylon, J.L.; Jiang, T.; Wang,
Y.; Muller, M.P.; et al. Microscopic Characterization of Membrane Transporter Function by In Silico Modeling and Simulation.
Methods Enzymol. 2016, 578, 373–428.
515. Jia, R.; Martens, C.; Shekhar, M.; Pant, S.; Pellowe, G.A.; Lau, A.M.; Findlay, H.E.; Harris, N.J.; Tajkhorshid, E.; Booth, P.J.; et al.
Hydrogen-deuterium exchange mass spectrometry captures distinct dynamics upon substrate and inhibitor binding to a trans-
porter. Nat. Commun. 2020, 11, 6162.
516. Vermaas, J.V.; Rempe, S.B.; Tajkhorshid, E. Electrostatic lock in the transport cycle of the multidrug resistance transporter EmrE.
Proc. Natl. Acad. Sci. USA 2018, 115, E7502–E7511.
517. Padariya, M.; Kalathiya, U.; Baginski, M. Structural and dynamic changes adopted by EmrE, multidrug transporter protein--
Studies by molecular dynamics simulation. Biochim. Biophys. Acta 2015, 1848, 2065–2074.
518. Padariya, M.; Kalathiya, U.; Baginski, M. Structural and dynamic insights on the EmrE protein with TPP+ and related substrates
through molecular dynamics simulations. Chem. Phys. Lipids 2018, 212, 111.
519. Li, J.; Zhao, Z.; Tajkhorshid, E. Locking Two Rigid-body Bundles in an Outward-Facing Conformation: The Ion-coupling Mech-
anism in a LeuT-fold Transporter. Sci. Rep. 2019, 9, 19479.
520. Zuo, Z.; Weng, J.; Wang, W. Insights into the Inhibitory Mechanism of D13-9001 to the Multidrug Transporter AcrB through
Molecular Dynamics Simulations. J. Phys. Chem. B 2016, 120, 2145–2154.
521. Jamshidi, S.; Sutton, J.M.; Rahman, K.M. Mapping the Dynamic Functions and Structural Features of AcrB Efflux Pump Trans-
porter Using Accelerated Molecular Dynamics Simulations. Sci. Rep. 2018, 8, 10470.
522. Johnson, R.M.; Fais, C.; Parmar, M.; Cheruvara, H.; Marshall, R.L.; Hesketh, S.J.; Feasey, M.C.; Ruggerone, P.; Vargiu, A.V.;
Postis, V.L.G.; et al. Cryo-EM Structure and Molecular Dynamics Analysis of the Fluoroquinolone Resistant Mutant of the AcrB
Transporter from Salmonella. Microorganisms 2020, 8, 943.
523. Pan, C.; Weng, J.; Wang, W. Conformational Dynamics and Protein-Substrate Interaction of ABC Transporter BtuCD at the
Occluded State Revealed by Molecular Dynamics Simulations. Biochemistry 2016, 55, 6897–6907.
524. Hsu, W.L.; Furuta, T.; Sakurai, M. Analysis of the Free Energy Landscapes for the Opening-Closing Dynamics of the Maltose
Transporter ATPase MalK2 Using Enhanced-Sampling Molecular Dynamics Simulation. J. Phys. Chem. B 2015, 119, 9717–9725.
525. Gu, R.X.; Corradi, V.; Singh, G.; Choudhury, H.G.; Beis, K.; Tieleman, D.P. Conformational Changes of the Antibacterial Peptide
ATP Binding Cassette Transporter McjD Revealed by Molecular Dynamics Simulations. Biochemistry 2015, 54, 5989–5998.
526. Immadisetty, K.; Hettige, J.; Moradi, M. What Can and Cannot Be Learned from Molecular Dynamics Simulations of Bacterial
Proton-Coupled Oligopeptide Transporter GkPOT? J. Phys. Chem. B 2017, 121, 3644–3656.
527. Cáceres-Delpiano, J.; Teneb, J.; Mansilla, R.; Garcia, A.; Salas-Burgos, A. Variations in periplasmic loop interactions determine
the pH-dependent activity of the hexameric urea transporter UreI from Helicobacter pylori: A molecular dynamics study. BMC
Struct. Biol. 2015, 15, 11.
528. Heinzelmann, G.; Kuyucak, S. Molecular dynamics simulations elucidate the mechanism of proton transport in the glutamate
transporter EAAT3. Biophys. J. 2014, 106, 2675–2683.
529. Park, M.S. Molecular Dynamics Simulations of the Human Glucose Transporter GLUT1. PLoS ONE 2015, 10, e0125361.
530. Reithmeier, R.A.F.; Casey, J.R.; Kalli, A.C.; Sansom, M.S.P.; Alguel, Y.; Iwata, S. Band 3, the human red cell chloride/bicarbonate
anion exchanger (AE1, SLC4A1), in a structural context. Biochim. Biophys. Acta 2016, 1858, 1507–1532.
531. Zhang, Y.; Zheng, G.; Fu, T.; Hong, J.; Li, F.; Yao, X.; Xue, W.; Zhu, F. The binding mode of vilazodone in the human serotonin
transporter elucidated by ligand docking and molecular dynamics simulations. Phys. Chem. Chem. Phys. 2020, 22, 5132–5144.
532. Mikou, A.; Cabaye, A.; Goupil, A.; Bertrand, H.O.; Mothet, J.P.; Acher, F.C. Asc-1 Transporter (SLC7A10): Homology Models
And Molecular Dynamics Insights Into The First Steps Of The Transport Mechanism. Sci. Rep. 2020, 10, 3731.
533. Briones, R.; Aponte-Santamaria, C.; de Groot, B.L. Localization and Ordering of Lipids Around Aquaporin-0: Protein and Lipid
Mobility Effects. Front. Physiol. 2017, 8, 124.
534. Saboe, P.O.; Rapisarda, C.; Kaptan, S.; Hsiao, Y.S.; Summers, S.R.; De Zorzi, R.; Dukovski, D.; Yu, J.; de Groot, B.L.; Kumar, M.;
et al. Role of Pore-Lining Residues in Defining the Rate of Water Conduction by Aquaporin-0. Biophys. J. 2017, 112, 953–965.
535. De Maré, S.W.; Venskutonyte, R.; Eltschkner, S.; de Groot, B.L.; Lindkvist-Petersson, K. Structural Basis for Glycerol Efflux and
Selectivity of Human Aquaporin 7. Structure 2020, 28, 215–222.e3.
536. Moss, F.J.; Mahinthichaichan, P.; Lodowski, D.T.; Kowatz, T.; Tajkhorshid, E.; Engel, A.; Boron, W.F.; Vahedi-Faridi, A. Aqua-
porin-7: A Dynamic Aquaglyceroporin With Greater Water and Glycerol Permeability Than Its Bacterial Homolog GlpF. Front.
Physiol. 2020, 11, 728.
537. Casiraghi, A.; Bensimon, A.; Superti-Furga, G. Recent developments in ligands and chemical probes targeting solute carrier
transporters. Curr. Opin. Chem. Biol. 2021, 62, 53–63.
538. Betters, J.L.; Yu, L. Transporters as drug targets: Discovery and development of NPC1L1 inhibitors. Clin. Pharmacol. Ther. 2010,
87, 117–121.
539. Ecker, G.; Chiba, P. Transporters as Drug Carriers: Structure, Function, Substrates; Wiley/VCH: Weinheim, Germany, 2009.
540. Ecker, G.F.; Clausen, R.P.; Sitte, H.H. Transporters as Drug Targets; Wiley: New York, NY, USA, 2017.
541. Lin, L.; Yee, S.W.; Kim, R.B.; Giacomini, K.M. SLC transporters as therapeutic targets: Emerging opportunities. Nat. Rev. Drug
Discov. 2015, 14, 543–560.
542. Qosa, H.; Mohamed, L.A.; Alqahtani, S.; Abuasal, B.S.; Hill, R.A.; Kaddoumi, A. Transporters as Drug Targets in Neurological
Diseases. Clin. Pharmacol. Ther. 2016, 100, 441–453.
Molecules 2021, 26, 5629 36 of 39
543. César-Razquin, A.; Girardi, E.; Yang, M.; Brehme, M.; Saez-Rodriguez, J.; Superti-Furga, G. In silico prioritization of transporter-
drug relationships from drug sensitivity screens. Front. Pharmacol. 2018, 9, 1011.
544. Garibsingh, R.A.; Schlessinger, A. Advances and Challenges in Rational Drug Design for SLCs. Trends Pharmacol. Sci. 2019, 40,
790–800.
545. Scalise, M.; Console, L.; Galluccio, M.; Pochini, L.; Indiveri, C. Chemical Targeting of Membrane Transporters: Insights into
Structure/Function Relationships. ACS Omega 2020, 5, 2069–2080.
546. Nakanishi, T.; Tamai, I. Solute carrier transporters as targets for drug delivery and pharmacological intervention for chemo-
therapy. J. Pharm. Sci. 2011, 100, 3731–3750.
547. Schumann, T.; Konig, J.; Henke, C.; Willmes, D.M.; Bornstein, S.R.; Jordan, J.; Fromm, M.F.; Birkenfeld, A.L. Solute Carrier
Transporters as Potential Targets for the Treatment of Metabolic Disease. Pharmacol. Rev. 2020, 72, 343–379.
548. Zhang, Y.; Zhang, Y.; Sun, K.; Meng, Z.; Chen, L. The SLC transporter in nutrient and metabolic sensing, regulation, and drug
development. J. Mol. Cell Biol. 2019, 11, 113.
549. Li, M.; Zhang, S.; Yang, B. Urea Transporters Identified as Novel Diuretic Drug Targets. Curr. Drug Targets 2020, 21, 279–287.
550. Zeden, M.S.; Burke, Ó.; Vallely, M.; Fingleton, C.; O'Gara, J.P. Exploring amino acid and peptide transporters as therapeutic
targets to attenuate virulence and antibiotic resistance in Staphylococcus aureus. PLoS Pathog. 2021, 17, e1009093.
551. Rochette, L.; Meloux, A.; Zeller, M.; Malka, G.; Cottin, Y.; Vergely, C. Mitochondrial SLC25 Carriers: Novel Targets for Cancer
Therapy. Molecules 2020, 25, 2417.
552. Wang, W.W.; Gallo, L.; Jadhav, A.; Hawkins, R.; Parker, C.G. The Druggability of Solute Carriers. J. Med. Chem. 2020, 63, 38343867.
553. Ceska, T.; Chung, C.W.; Cooke, R.; Phillips, C.; Williams, P.A. Cryo-EM in drug discovery. Biochem. Soc. Trans. 2019, 47, 281293.
554. Bajorath, J. Molecular Similarity Concepts for Informatics Applications. Methods Mol. Biol. 2017, 1526, 231–245.
555. Bender, A.; Glen, R.C. Molecular similarity: A key technique in molecular informatics. Org. Biomol. Chem. 2004, 2, 3204–3218.
556. Bender, A.; Jenkins, J.L.; Li, Q.L.; Adams, S.E.; Cannon, E.O.; Glen, R.C. Molecular Similarity: Advances in Methods, Applica-
tions and Validations in Virtual Screening and QSAR. Annu. Rep. Comput. Chem. 2006, 2, 141–168.
557. Eckert, H.; Bajorath, J. Molecular similarity analysis in virtual screening: Foundations, limitations and novel approaches. Drug
Discov. Today 2007, 12, 225–233.
558. Floris, M.; Olla, S. Molecular Similarity in Computational Toxicology. Methods Mol. Biol. 2018, 1800, 171–179.
559. Ginn, C.M.R.; Willett, P.; Bradshaw, J. Combination of molecular similarity measures using data fusion. In Virtual Screening: An
Alternative or Complement to High Throughput Screening?; Springer: Dordrecht, The Netherlands, 2000; Volume 20, pp. 1–16.
560. Johnson, M.A.; Maggiora, G.M. Concepts and Applications of Molecular Similarity; Wiley: New York, NY, USA, 1990.
561. Maggiora, G.M.; Shanmugasundaram, V. Molecular Similarity Measures. Methods Mol. Biol. 2011, 672, 39–100.
562. Medina-Franco, J.L.; Maggiora, G.M. Molecular similarity analysis. In Chemoinformatics for Drug Discovery; Bajorath, J., Ed.
Wiley: Hoboken, NJ, USA, 2014; pp. 343–399.
563. Zahoránszky-Kőhalmi, G.; Bologa, C.G.; Oprea, T.I. Impact of similarity threshold on the topology of molecular similarity net-
works and clustering outcomes. J. Cheminfor. 2016, 8, 16.
564. Gasteiger, J. Handbook of Chemoinformatics: From Data to Knowledge; Wiley/VCH: Weinheim, Germany, 2003.
565. O’Hagan, S.; Swainston, N.; Handl, J.; Kell, D.B. A ‘rule of 0.5’ for the metabolite-likeness of approved pharmaceutical drugs.
Metabolomics 2015, 11, 323–339.
566. O’Hagan, S.; Kell, D.B. Understanding the foundations of the structural similarities between marketed drugs and endogenous
human metabolites. Front. Pharmacol. 2015, 6, 105.
567. O'Hagan, S.; Kell, D.B. MetMaxStruct: A Tversky-similarity-based strategy for analysing the (sub)structural similarities of drugs
and endogenous metabolites. Front. Pharmacol. 2016, 7, 266.
568. O'Hagan, S.; Kell, D.B. Analysis of drug-endogenous human metabolite similarities in terms of their maximum common sub-
structures. J. Cheminformatics 2017, 9, 18.
569. Thiele, I.; Swainston, N.; Fleming, R.M.T.; Hoppe, A.; Sahoo, S.; Aurich, M.K.; Haraldsdottír, H.; Mo, M.L.; Rolfsson, O.; Stobbe,
M.D.; et al. A community-driven global reconstruction of human metabolism. Nat. Biotechnol. 2013, 31, 419–425.
570. Wishart, D.S.; Feunang, Y.D.; Marcu, A.; Guo, A.C.; Liang, K.; Vazquez-Fresno, R.; Sajed, T.; Johnson, D.; Li, C.; Karu, N.; et al.
HMDB 4.0: The human metabolome database for 2018. Nucleic Acids Res. 2018, 46, D608–D617.
571. Rosen, J.; Gottfries, J.; Muresan, S.; Backlund, A.; Oprea, T.I. Novel chemical space exploration via natural products. J. Med.
Chem. 2009, 52, 1953–1962.
572. Butler, M.S.; Robertson, A.A.B.; Cooper, M.A. Natural product and natural product derived drugs in clinical trials. Nat. Prod.
Rep. 2014, 31, 1612–1661.
573. Doak, B.C.; Over, B.; Giordanetto, F.; Kihlberg, J. Oral druggable space beyond the rule of 5: Insights from drugs and clinical
candidates. Chem. Biol. 2014, 21, 1115–1142.
574. Doak, B.C.; Kihlberg, J. Drug discovery beyond the rule of 5Opportunities and challenges. Expert Opin. Drug Discov. 2017, 12,
115–119.
575. Harvey, A.L. Natural products in drug discovery. Drug Discov. Today 2008, 13, 894–901.
576. Ganesan, A. The impact of natural products upon modern drug discovery. Curr. Opin. Chem. Biol. 2008, 12, 306–317.
577. Cragg, G.M.; Newman, D.J. Natural products: A continuing source of novel drug leads. Biochim. Biophys. Acta 2013, 1830, 36703695.
578. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs over the Nearly Four Decades from 01/1981 to 09/2019.
J. Nat. Prod. 2020, 83, 770–803.
Molecules 2021, 26, 5629 37 of 39
579. Dias, D.A.; Urban, S.; Roessner, U. A historical overview of natural products in drug discovery. Metabolites 2012, 2, 303–336.
580. Ji, H.F.; Li, X.J.; Zhang, H.Y. Natural products and drug discovery Can thousands of years of ancient medical knowledge lead
us to new and powerful drug combinations in the fight against cancer and dementia? EMBO Rep. 2009, 10, 194–200.
581. Lahlou, M. The Success of Natural Products in Drug Discovery. Pharmacol. Pharm. 2013, 4, 17–31.
582. Luo, F.; Gu, J.; Chen, L.; Xu, X. Systems pharmacology strategies for anticancer drug discovery based on natural products. Mol.
Biosyst. 2014, 10, 1912–1917.
583. Silva, T.; Reis, J.; Teixeira, J.; Borges, F. Alzheimer’s disease, enzyme targets and drug discovery struggles: From natural prod-
ucts to drug prototypes. Ageing Res. Rev. 2014, 15, 116–145.
584. Wright, G.D. Something old, something new: Revisiting natural products in antibiotic drug discovery. Can. J. Microbiol. 2014,
60, 147–154.
585. Wright, G.D. Unlocking the potential of natural products in drug discovery. Microb. Biotechnol. 2019, 12, 55–57.
586. Zheng, C.L.; Wang, J.A.; Liu, J.L.; Pei, M.J.; Huang, C.; Wang, Y.H. System-level multi-target drug discovery from natural prod-
ucts with applications to cardiovascular diseases. Mol. Divers. 2014, 18, 621–635.
587. Camp, D.; Garavelas, A.; Campitelli, M. Analysis of Physicochemical Properties for Drugs of Natural Origin. J. Nat. Prod. 2015,
78, 1370–1382.
588. Lipinski, C.A. Rule of five in 2015 and beyond: Target and ligand structural limitations, ligand chemistry structure and drug
discovery project decisions. Adv. Drug Deliv. Rev. 2016, 101, 34–41.
589. Walters, W.P. Going further than Lipinski's rule in drug design. Exp. Opin. Drug Discov. 2012, 7, 99–107.
590. Zhang, M.Q.; Wilkinson, B. Drug discovery beyond the ‘rule-of-five’. Curr. Opin. Biotechnol. 2007, 18, 478–488.
591. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility
and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 1997, 23, 325.
592. Giacomini, K.M.; Huang, S.M.; Tweedie, D.J.; Benet, L.Z.; Brouwer, K.L.; Chu, X.; Dahlin, A.; Evers, R.; Fischer, V.; Hillgren,
K.M.; et al. Membrane transporters in drug development. Nat. Rev. Drug Discov. 2010, 9, 215–236.
593. Giacomini, K.M.; Galetin, A.; Huang, S.M. The International Transporter Consortium: Summarizing Advances in the Role of
Transporters in Drug Development. Clin. Pharmacol. Ther. 2018, 104, 766–771.
594. Müller, J.; Keiser, M.; Drozdzik, M.; Oswald, S. Expression, regulation and function of intestinal drug transporters: An update.
Biol. Chem. 2017, 398, 175–192.
595. Sai, Y.; Tsuji, A. Transporter-mediated drug delivery: Recent progress and experimental approaches. Drug Discov. Today 2004,
9, 712–720.
596. Liu, X. Transporter-Mediated Drug-Drug Interactions and Their Significance. Adv. Exp. Med. Biol. 2019, 1141, 241–291.
597. Saunders, N.R.; Dziegielewska, K.M.; Møllgård, K.; Habgood, M.D. Recent Developments in Understanding Barrier Mecha-
nisms in the Developing Brain: Drugs and Drug Transporters in Pregnancy, Susceptibility or Protection in the Fetal Brain?
Annu. Rev. Pharmacol. Toxicol. 2019, 59, 487–505.
598. Darbani, B.; Kell, D.B.; Borodina, I. Energetic evolution of cellular transportomes BMC Genom. 2018, 19, 418.
599. König, J.; Müller, F.; Fromm, M.F. Transporters and drug-drug interactions: Important determinants of drug disposition and
effects. Pharmacol. Rev. 2013, 65, 944–966.
600. Franke, R.M.; Gardner, E.R.; Sparreboom, A. Pharmacogenetics of drug transporters. Curr. Pharm. Des. 2010, 16, 220–230.
601. Harwood, M.D.; Neuhoff, S.; Carlson, G.L.; Warhurst, G.; Rostami-Hodjegan, A. Absolute abundance and function of intestinal
drug transporters: A prerequisite for fully mechanistic in vitro-in vivo extrapolation of oral drug absorption. Biopharm. Drug
Dispos. 2013, 34, 228.
602. Ishikawa, T.; Tsuji, A.; Inui, K.; Sai, Y.; Anzai, N.; Wada, M.; Endou, H.; Sumino, Y. The genetic polymorphism of drug trans-
porters: Functional analysis approaches. Pharmacogenomics 2004, 5, 67–99.
603. Ivanyuk, A.; Livio, F.; Biollaz, J.; Buclin, T. Renal Drug Transporters and Drug Interactions. Clin. Pharmacokinet. 2017, 56, 825892.
604. Lai, Y.; Sampson, K.E.; Stevens, J.C. Evaluation of drug transporter interactions in drug discovery and development. Comb.
Chem. High Throughput Screen. 2010, 13, 112–134.
605. Lai, Y.; Hsiao, P. Beyond the ITC White Paper: Emerging sciences in drug transporters and opportunities for drug development.
Curr. Pharm. Des. 2014, 20, 1577–1594.
606. Lee, W.; Ha, J.M.; Sugiyama, Y. Post-translational regulation of the major drug transporters in the families of organic anion
transporters and organic anion-transporting polypeptides. J. Biol. Chem. 2020, 295, 17349–17364.
607. Neul, C.; Schaeffeler, E.; Sparreboom, A.; Laufer, S.; Schwab, M.; Nies, A.T. Impact of Membrane Drug Transporters on Re-
sistance to Small-Molecule Tyrosine Kinase Inhibitors. Trends Pharmacol. Sci. 2016, 37, 904–932.
608. Nigam, S.K. What do drug transporters really do? Nat. Rev. Drug Discov. 2015, 14, 29–44.
609. Petzinger, E.; Geyer, J. Drug transporters in pharmacokinetics. Naunyn Schmiedebergs Arch. Pharmacol. 2006, 372, 465–475.
610. Rodrigues, A.D.; Taskar, K.S.; Kusuhara, H.; Sugiyama, Y. Endogenous Probes for Drug Transporters: Balancing Vision With
Reality. Clin. Pharmacol. Ther. 2018, 103, 434–448.
611. Thwaites, D.T.; Anderson, C.M. H+-coupled nutrient, micronutrient and drug transporters in the mammalian small intestine.
Exp. Physiol. 2007, 92, 603–619.
612. Unadkat, J.D.; Dahlin, A.; Vijay, S. Placental drug transporters. Curr. Drug Metab. 2004, 5, 125–131.
613. Zhang, L.; Strong, J.M.; Qiu, W.; Lesko, L.J.; Huang, S.M. Scientific perspectives on drug transporters and their role in drug
interactionst. Mol. Pharm. 2006, 3, 62–69.
Molecules 2021, 26, 5629 38 of 39
614. Vora, B.; Green, E.A.E.; Khuri, N.; Ballgren, F.; Sirota, M.; Giacomini, K.M. Drug-nutrient interactions: Discovering prescription
drug inhibitors of the thiamine transporter ThTR-2 (SLC19A3). Am. J. Clin. Nutr. 2020, 111, 110–121.
615. Terada, T.; Inui, K. Gene expression and regulation of drug transporters in the intestine and kidney. Biochem. Pharmacol. 2007,
73, 440–449.
616. Sugiyama, Y.; Steffansen, B. Transporters in Drug Development: Discovery, Optimization, Clinical Study and Regulation;
AAPS/Springer: New York, NY, USA, 2013.
617. Koepsell, H. Organic Cation Transporters in Health and Disease. Pharmacol. Rev. 2020, 72, 253–319.
618. Schlessinger, A.; Welch, M.A.; van Vlijmen, H.; Korzekwa, K.; Swaan, P.W.; Matsson, P. Molecular Modeling of Drug-Trans-
porter Interactions-An International Transporter Consortium Perspective. Clin. Pharmacol. Ther. 2018, 104, 818–835.
619. Zamek-Gliszczynski, M.J.; Giacomini, K.M.; Zhang, L. Emerging Clinical Importance of Hepatic Organic Cation Transporter 1
(OCT1) in Drug Pharmacokinetics, Dynamics, Pharmacogenetic Variability, and Drug Interactions. Clin. Pharmacol. Ther. 2018,
103, 758–760.
620. Zamek-Gliszczynski, M.J.; Taub, M.E.; Chothe, P.P.; Chu, X.; Giacomini, K.M.; Kim, R.B.; Ray, A.S.; Stocker, S.L.; Unadkat, J.D.;
Wittwer, M.B.; et al. International Transporter, C., Transporters in Drug Development: 2018 ITC Recommendations for Trans-
porters of Emerging Clinical Importance. Clin. Pharmacol. Ther. 2018, 104, 890–899.
621. Julsing, M.K.; Schrewe, M.; Cornelissen, S.; Hermann, I.; Schmid, A.; Bühler, B. Outer membrane protein AlkL boosts biocata-
lytic oxyfunctionalization of hydrophobic substrates in Escherichia coli. Appl. Environ. Microbiol. 2012, 78, 5724–5733.
622. Call, T.P.; Akhtar, M.K.; Baganz, F.; Grant, C. Modulating the import of medium-chain alkanes in E. coli through tuned expres-
sion of FadL. J. Biol. Eng. 2016, 10, 5.
623. Cornelissen, S.; Julsing, M.K.; Volmer, J.; Riechert, O.; Schmid, A.; Bühler, B. Whole-cell-based CYP153A6-catalyzed (S)-limo-
nene hydroxylation efficiency depends on host background and profits from monoterpene uptake via AlkL. Biotechnol. Bioeng.
2013, 110, 1282–1292.
624. Ciarimboli, G.; Gautron, S.; Schlatter, E. Organic Cation Transporters: Integration of Physiology, Pathology and Pharmacology;
Springer: Berilin/Heidelberg, Germany, 2016.
625. Beck, J.G.; Chatterjee, J.; Laufer, B.; Kiran, M.U.; Frank, A.O.; Neubauer, S.; Ovadia, O.; Greenberg, S.; Gilon, C.; Hoffman, A.; et al.
Intestinal Permeability of Cyclic Peptides: Common Key Backbone Motifs Identified. J. Am. Chem. Soc. 2012, 134, 1212512133.
626. Corti, G.; Maestrelli, F.; Cirri, M.; Zerrouk, N.; Mura, P. Development and evaluation of an in vitro method for prediction of
human drug absorptionII. Demonstration of the method suitability. Eur. J. Pharm. Sci. 2006, 27, 354–362.
627. Skolnik, S.; Lin, X.; Wang, J.; Chen, X.H.; He, T.; Zhang, B. Towards prediction of in vivo intestinal absorption using a 96-well
Caco-2 assay. J. Pharm. Sci. 2010, 99, 3246–3265.
628. Kell, D.B.; Samanta, S.; Swainston, N. Deep learning and generative methods in cheminformatics and chemical biology: Navi-
gating small molecule space intelligently Biochem. J. 2020, 477, 4559–4580.
629. LeCun, Y.; Bengio, Y.; Hinton, G. Deep learning. Nature 2015, 521, 436–444.
630. Schmidhuber, J. Deep learning in neural networks: An overview. Neural Netw. 2015, 61, 85–117.
631. Paliwal, K.; Lyons, J.; Heffernan, R. A Short Review of Deep Learning Neural Networks in Protein Structure Prediction Prob-
lems. Adv. Tech. Biol. Med. 2015, 3, 3.
632. Torrisi, M.; Pollastri, G.; Le, Q. Deep learning methods in protein structure prediction. Comput. Struct. Biotechnol. J. 2020, 18,
1301–1310.
633. Torrisi, M.; Pollastri, G. Brewery: Deep learning and deeper profiles for the prediction of 1D protein structure annotations.
Bioinformatics 2020, 36, 3897–3898.
634. Wang, J.; Cao, H.; Zhang, J.Z.H.; Qi, Y. Computational Protein Design with Deep Learning Neural Networks. Sci. Rep. 2018, 8, 6349.
635. Xu, J. Distance-based protein folding powered by deep learning. Proc. Natl. Acad. Sci. USA 2019, 116, 16856–16865.
636. Drori, I.; Thaker, D.; Srivatsa, A.; Jeong, D.; Wang, Y.; Nan, L.; Wu, F.; Leggas, D.; Lei, J.; Lu, W.; et al. Accurate Protein Structure
Prediction by Embeddings and Deep Learning Representations. arXiv 2019, arXiv:1911.05531v1. Available online:
https://arxiv.org/abs/1911.05531v1 (accessed on 15/9/21).
637. Senior, A.W.; Evans, R.; Jumper, J.; Kirkpatrick, J.; Sifre, L.; Green, T.; Qin, C.; Zidek, A.; Nelson, A.W.R.; Bridgland, A.; et al.
Protein structure prediction using multiple deep neural networks in the 13th Critical Assessment of Protein Structure Prediction
(CASP13). Proteins 2019, 87, 1141–1148.
638. Senior, A.W.; Evans, R.; Jumper, J.; Kirkpatrick, J.; Sifre, L.; Green, T.; Qin, C.; Zidek, A.; Nelson, A.W.R.; Bridgland, A.; et al.
Improved protein structure prediction using potentials from deep learning. Nature 2020, 577, 706–710.
639. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Zidek, A.;
Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589.
640. Tunyasuvunakool, K.; Adler, J.; Wu, Z.; Green, T.; Zielinski, M.; Zidek, A.; Bridgland, A.; Cowie, A.; Meyer, C.; Laydon, A.; et
al. Highly accurate protein structure prediction for the human proteome. Nature 2021, 596, 590–596.
641. Pereira, J.; Simpkin, A.J.; Hartmann, M.D.; Rigden, D.J.; Keegan, R.M.; Lupas, A.N. High-accuracy protein structure prediction
in CASP14. Proteins 2021, https://doi.org/10.1002/prot.26171.
642. Simpkin, A.J.; Sánchez Rodríguez, F.; Mesdaghi, S.; Kryshtafovych, A.; Rigden, D.J. Evaluation of model refinement in CASP14.
Proteins 2021, https://doi.org/10.1002/prot.26185.
Molecules 2021, 26, 5629 39 of 39
643. Baek, M.; DiMaio, F.; Anishchenko, I.; Dauparas, J.; Ovchinnikov, S.; Lee, G.R.; Wang, J.; Cong, Q.; Kinch, L.N.; Schaeffer, R.D.;
et al. Accurate prediction of protein structures and interactions using a 3-track network. bioRxiv 2021,
https://doi.org/10.1101/2021.06.14.448402.
644. Bouatta, N.; Sorger, P.; AlQuraishi, M. Protein structure prediction by AlphaFold2: Are attention and symmetries all you need?
Acta Crystallogr. D Struct. Biol. 2021, 77, 982–991.
645. Callaway, E. DeepMind’s AI predicts structures for a vast trove of proteins. Nature 2021, 595, 635.
646. Stepniewska-Dziubinska, M.M.; Zielenkiewicz, P.; Siedlecki, P. Development and evaluation of a deep learning model for pro-
tein-ligand binding affinity prediction. Bioinformatics 2018, 34, 3666–3674.
647. Tian, K.; Shao, M.; Wang, Y.; Guan, J.; Zhou, S. Boosting compound-protein interaction prediction by deep learning. Methods
2016, 110, 64–72.
648. Verma, N.; Qu, X.; Trozzi, F.; Elsaied, M.; Karki, N.; Tao, Y.; Zoltowski, B.; Larson, E.C.; Kraka, E. SSnet: A Deep Learning
Approach for Protein-Ligand Interaction Prediction. Int J. Mol. Sci 2021, 22, 1392.
649. Jiménez-Luna, J.; Grisoni, F.; Weskamp, N.; Schneider, G. Artificial intelligence in drug discovery: Recent advances and future
perspectives. Expert Opin. Drug Discov. 2021, 16, 949–959, https://doi.org/10.1080/17460441.2021.1909567.
650. Laine, E.; Eismann, S.; Elofsson, A.; Grudinin, S. Protein sequence-to-structure learning: Is this the end(-to-end revolution)?
arXiv 2021, arXiv:2105.07407. Available online: https://arxiv.org/abs/2105.07407 (accessed on 15/9/21).
651. Arús-Pous, J.; Probst, D.; Reymond, J.L. Deep Learning Invades Drug Design and Synthesis. Chimia 2018, 72, 70–71.
652. Arús-Pous, J.; Blaschke, T.; Ulander, S.; Reymond, J.L.; Chen, H.; Engkvist, O. Exploring the GDB-13 chemical space using deep
generative models. J. Cheminfor. 2019, 11, 20.
653. Jurtz, V.I.; Johansen, A.R.; Nielsen, M.; Almagro Armenteros, J.J.; Nielsen, H.; Sønderby, C.K.; Winther, O.; Sønderby, S.K. An
introduction to deep learning on biological sequence data: Examples and solutions. Bioinformatics 2017, 33, 3685–3690.
654. Gómez-Bombarelli, R.; Wei, J.N.; Duvenaud, D.; Hernández-Lobato, J.M.; Sánchez-Lengeling, B.; Sheberla, D.; Aguilera-Iparra-
guirre, J.; Hirzel, T.D.; Adams, R.P.; Aspuru-Guzik, A. Automatic Chemical Design Using a Data-Driven Continuous Represen-
tation of Molecules. ACS Cent. Sci. 2018, 4, 268–276.
655. Khemchandani, Y.; O’Hagan, S.; Samanta, S.; Swainston, N.; Roberts, T.J.; Bollegala, D.; Kell, D.B. DeepGraphMolGen, a multi-
objective, computational strategy for generating molecules with desirable properties: A graph convolution and reinforcement
learning approach. J. Cheminfor. 2020, 12, 53.
656. Shrivastava, A.D.; Kell, D.B. FragNet, a contrastive learning-based transformer model for clustering, interpreting, visualising
and navigating chemical space. Molecules 2021, 26, 2065.
657. Karimi, M.; Wu, D.; Wang, Z.; Shen, Y. DeepAffinity: Interpretable deep learning of compound-protein affinity through unified
recurrent and convolutional neural networks. Bioinformatics 2019, 35, 3329–3338.
658. Biswas, S.; Khimulya, G.; Alley, E.C.; Esvelt, K.M.; Church, G.M. Low-N protein engineering with data-efficient deep learning.
Nat. Methods 2021, 18, 389–396.
659. Wu, Z.; Yang, K.K.; Liszka, M.; Lee, A.; Batzilla, A.; Wernick, D.; Weiner, D.P.; Arnold, F.H. Signal peptides generated by atten-
tion-based neural networks. ACS Synth. Biol. 2020, 9, 2154–2161.
660. Elnaggar, A.; Heinzinger, M.; Dallago, C.; Rihawi, G.; Wang, Y.; Jones, L.; Gibbs, T.; Feher, T.; Angerer, C.; Steinegger, M.; et al.
ProtTrans: Towards Cracking the Language of Life’s Code Through Self-Supervised Deep Learning and High Performance
Computing. arXiv 2020, arXiv:2007.06225. Available online: https://arxiv.org/abs/2007.06225 (accessed on 15/921).
661. Riesselman, A.J.; Ingraham, J.B.; Marks, D.S. Deep generative models of genetic variation capture the effects of mutations. Nat.
Methods 2018, 15, 816–822.
662. Volk, M.J.; Lourentzou, I.; Mishra, S.; Vo, L.T.; Zhai, C.; Zhao, H. Biosystems Design by Machine Learning. ACS Synth. Biol.
2020, 9, 1514–1533.
663. Mulligan, V.K. Current directions in combining simulation-based macromolecular modeling approaches with deep learning.
Expert Opin. Drug Discov. 2021, 16, 1025–1044, https://doi.org/10.1080/17460441.2021.1918097.
664. Meinen, B.A.; Bahl, C.D. Breakthroughs in computational design methods open up new frontiers for de novo protein engineer-
ing. Protein Eng. Des. Sel. 2021, 34, gzab007, https://doi.org/10.1093/protein/gzab007.
665. Kreutter, D.; Schwaller, P.; Reymond, J.-L. Predicting enzymatic reactions with a molecular transformer. Chem. Sci. 2021, 12,
8648–8659.
... 6 Owing to continuous advances in microbiology and biotechnology, the idea that microbial cell membranes are naturally permeable towards most compounds, which passively diffuse through with little physiological control, has been largely superseded. 7 As discussed by scientists specializing in cell membranes, evolution has heavily selected against passively allowing compounds to enter microbial organisms and potentially cause significant damage to the cell, as well as against persistently leaking valuable chemical intermediates. 7 For this reason, even the smallest polar molecules are generally unable to enter or exit the cell at physiologically significant rates without the presence of specific transport proteins. ...
... 7 As discussed by scientists specializing in cell membranes, evolution has heavily selected against passively allowing compounds to enter microbial organisms and potentially cause significant damage to the cell, as well as against persistently leaking valuable chemical intermediates. 7 For this reason, even the smallest polar molecules are generally unable to enter or exit the cell at physiologically significant rates without the presence of specific transport proteins. 7,8 Indeed, the plasma membrane has a remarkably high protein content, and many of these proteins actively regulate the translocation of compounds across the lipid bilayer. ...
... 7 For this reason, even the smallest polar molecules are generally unable to enter or exit the cell at physiologically significant rates without the presence of specific transport proteins. 7,8 Indeed, the plasma membrane has a remarkably high protein content, and many of these proteins actively regulate the translocation of compounds across the lipid bilayer. 8 These transporters thus indirectly control the concentration of biomolecules inside the cytoplasm and cellular compartments, affecting enzymatic conversions and the carbon flux through different metabolic pathways. ...
Article
Full-text available
As microbial membranes are naturally impermeable to even the smallest biomolecules, transporter proteins are physiologically essential for normal cell functioning. This makes transporters a key target area for engineering enhanced cell factories. As part of the wider cellular transportome, aquaporins (AQPs) are responsible for transporting small polar solutes, encompassing many compounds which are of great interest for industrial biotechnology, including cell feedstocks, numerous commercially relevant polyols and even weak organic acids. In this review, examples of cell factory engineering by targeting AQPs are presented. These AQP modifications aid in redirecting carbon fluxes and boosting bioconversions either by enhanced feedstock uptake, improved intermediate retention, increasing product export into the media or superior cell viability against stressors with applications in both bacterial and yeast production platforms. Additionally, the future potential for AQP deployment and targeting is discussed, showcasing hurdles and considerations of this strategy as well as recent advances and future directions in the field. By leveraging the natural diversity of AQPs and breakthroughs in channel protein engineering, these transporters are poised to be promising tools capable of enhancing a wide variety of biotechnological processes.
... In many cases, the genetic basis of tolerance involves mutations in transporter-encoding genes responsible for exchanging the compound across cellular membranes (Kell, 2021;Lennen et al., 2023;Pereira et al., 2019Pereira et al., , 2020Radi, Munro, et al., 2022;Radi, SalcedoSora, et al., 2022). Beyond improving product tolerance, transporter engineering can offer multiple advantages, including alleviating feedback inhibition, reducing downstream processing costs and preventing the leakage of pathway intermediates, all of which can improve the economics of bioprocesses (Kell, 2018(Kell, , 2021Kell et al., 2015;Munro & Kell, 2021;van der Hoek & Borodina, 2020;Zhu et al., 2020). ...
... In many cases, the genetic basis of tolerance involves mutations in transporter-encoding genes responsible for exchanging the compound across cellular membranes (Kell, 2021;Lennen et al., 2023;Pereira et al., 2019Pereira et al., , 2020Radi, Munro, et al., 2022;Radi, SalcedoSora, et al., 2022). Beyond improving product tolerance, transporter engineering can offer multiple advantages, including alleviating feedback inhibition, reducing downstream processing costs and preventing the leakage of pathway intermediates, all of which can improve the economics of bioprocesses (Kell, 2018(Kell, , 2021Kell et al., 2015;Munro & Kell, 2021;van der Hoek & Borodina, 2020;Zhu et al., 2020). ...
Article
Full-text available
Aromatic compounds are used in pharmaceutical, food, textile and other industries. Increased demand has sparked interest in exploring biotechnological approaches for their sustainable production as an alternative to chemical synthesis from petrochemicals or plant extraction. These aromatic products may be toxic to microorganisms, which complicates their production in cell factories. In this study, we analysed the toxicity of multiple aromatic compounds in common production hosts. Next, we screened a subset of toxic aromatics, namely 2‐phenylethanol, 4‐tyrosol, benzyl alcohol, berberine and vanillin, against transporter deletion libraries in Escherichia coli and Saccharomyces cerevisiae. We identified multiple transporter deletions that modulate the tolerance of the cells towards these compounds. Lastly, we engineered transporters responsible for 2‐phenylethanol tolerance in yeast and showed improved 2‐phenylethanol bioconversion from L‐phenylalanine, with deletions of YIA6, PTR2 or MCH4 genes improving titre by 8–12% and specific yield by 38–57%. Our findings provide insights into transporters as targets for improving the production of aromatic compounds in microbial cell factories.
... Carrier-mediated transport can be classied into two main types: facilitated diffusion and active transport. 21 Facilitated diffusion involves the passive movement of molecules along their concentration gradient. In this process, carrier proteins facilitate the transport of molecules that would otherwise have difficulty crossing the membrane due to their size, charge, or hydrophobicity. ...
Article
Full-text available
Monitoring physiological changes within cells is crucial for understanding their biological aspects and pathological activities. Fluorescent probes serve as powerful tools for this purpose, offering advantageous characteristics over genetically encoded probes. While numerous organelle-selective probes have been developed in the past decades, several challenges persist. This review explores the strategies and key factors contributing to the successful rationale design of these probes. We systematically discuss the typical mode of cellular uptake generally adopted by fluorescent probes and provide a detailed examination of the key factors to consider in design rationale from two perspectives: the properties of the target organelle and the physicochemical properties of the probe itself. Additionally, recent examples of organelle-targeted probes are presented, along with a discussion of the current challenges faced by fluorescent probes in the field.
... The resulting acetate can be upcycled into higher-value products through chemical or biological processes [35,37]. The acetate uptake by organisms is either transporter-mediated by facilitated diffusion or actively through transporters under energy consumption [35,41]. Once taken up, acetate is activated to form acetyl-CoA, serving as a precursor for various metabolic pathways, including de novo fatty acid synthesis, essential for glycolipid synthesis [35,42]. ...