ArticlePDF AvailableLiterature Review

Electrochemical Biosensors Based on S-Layer Proteins

Authors:

Abstract and Figures

Designing and development of electrochemical biosensors enable molecule sensing and quantification of biochemical compositions with multitudinous benefits such as monitoring, detection, and feedback for medical and biotechnological applications. Integrating bioinspired materials and electrochemical techniques promote specific, rapid, sensitive, and inexpensive biosensing platforms for (e.g., point-of-care testing). The selection of biomaterials to decorate a biosensor surface is a critical issue as it strongly affects selectivity and sensitivity. In this context, smart biomaterials with the intrinsic self-assemble capability like bacterial surface (S-) layer proteins are of paramount importance. Indeed, by forming a crystalline two-dimensional protein lattice on many sensors surfaces and interfaces, the S-layer lattice constitutes an immobilization matrix for small biomolecules and lipid membranes and a patterning structure with unsurpassed spatial distribution for sensing elements and bioreceptors. This review aims to highlight on exploiting S‑layer proteins in biosensor technology for various applications ranging from detection of metal ions over small organic compounds to cells. Furthermore, enzymes immobilized on the S‑layer proteins allow specific detection of several vital biomolecules. The special features of the S‑layer protein lattice as part of the sensor architecture enhances surface functionalization and thus may feature an innovative class of electrochemical biosensors.
Content may be subject to copyright.
Sensors 2020, 20, 1721; doi:10.3390/s20061721 www.mdpi.com/journal/sensors
Review
Electrochemical Biosensors Based on S-Layer Proteins
Samar Damiati 1,2,3 and Bernhard Schuster 2,*
1 Department of Biochemistry, Faculty of Science, King Abdulaziz University (KAU), Jeddah 21589,
Saudi Arabia; sdamiati@kau.edu.sa
2 Institute for Synthetic Bioarchitectures, Department of NanoBiotechnology, BOKU,
University of Natural Resources and Life Sciences, Vienna, Muthgasse 11, 1190 Vienna, Austria
3 Division of Nanobiotechnology, Department of Protein Science, Science for Life Laboratory,
School of Engineering Sciences in Chemistry, Biotechnology and Health, KTH Royal Institute
of Technology, 171 21 Solna, Stockholm, Sweden
* Correspondence: bernhard.schuster@boku.ac.at; Tel.: +43 1 47654 80436
Received: 12 February 2020; Accepted: 17 March 2020; Published: 19 March 2020
Abstract: Designing and development of electrochemical biosensors enable molecule sensing and
quantification of biochemical compositions with multitudinous benefits such as monitoring,
detection, and feedback for medical and biotechnological applications. Integrating bioinspired
materials and electrochemical techniques promote specific, rapid, sensitive, and inexpensive
biosensing platforms for (e.g., point-of-care testing). The selection of biomaterials to decorate a
biosensor surface is a critical issue as it strongly affects selectivity and sensitivity. In this context,
smart biomaterials with the intrinsic self-assemble capability like bacterial surface (S-) layer proteins
are of paramount importance. Indeed, by forming a crystalline two-dimensional protein lattice on
many sensors surfaces and interfaces, the S-layer lattice constitutes an immobilization matrix for
small biomolecules and lipid membranes and a patterning structure with unsurpassed spatial
distribution for sensing elements and bioreceptors. This review aims to highlight on exploiting
S-layer proteins in biosensor technology for various applications ranging from detection of metal
ions over small organic compounds to cells. Furthermore, enzymes immobilized on the S-layer
proteins allow specific detection of several vital biomolecules. The special features of the S-layer
protein lattice as part of the sensor architecture enhances surface functionalization and thus may
feature an innovative class of electrochemical biosensors.
Keywords: S-layer proteins; biosensor; biocompatible layer; self-assembly; bioinspired material
1. Introduction
Biosensors are analytical devices that contain an integrated set of tools for the conversion of a
biological response or an analyte concentration to specific (semi)quantitative information [1]. A
typical biosensor is composed of three main parts (Figure 1): (i) Bioreceptors that specifically
recognize an analyte and capture it; (ii) an interface matrix, which is a functional area where reactions
occur; and (iii) a physico-chemical transducer element that converts the generated event into a
(mostly electronic) quantifiable signal [1,2]. The detector unit can amplify the transducer signal before
its transmission for processing and analysis by computer software. Several transduction mechanisms
depend on the conversion of optical, acoustic, electrical, chemical, mechanical, or thermal properties,
or a combination of different mechanisms, to detect biorecognition events [1,2]. All these techniques
have led to the development of biosensors that are currently on the market and there is a great deal
of interest in developing devices applicable as diagnostic tools for point-of-care (POC) testing. The
main obstacles for the development of POC sensing devices with high sensitivity and selectivity are
the complexity of the transduction principles and high production costs. Moreover, patients cannot
use the POC devices by themselves, as skillful technicians are necessary to operate these instruments.
Sensors 2020, 20, 1721 2 of 22
Electrochemical biosensors offer many advantages, including simplicity, low cost, short
measurement time, requirement for only small volumes of analyte and reagents, and high selectivity
and sensitivity. Moreover, easy operation is feasible that enables patients to handle electrochemical
biosensors at home without the need for expensive facilities and personnel. Further, microelectronic
circuits can not only be produced at low cost but can also easily connect with common electronic
read-out and processing devices [2–4]. In contrast, electrochemical biosensors also possess some
limitations. For example, although they make it possible to detect a target analyte in a complex
biofluid sample, the pH and ionic strength can markedly interfere with the biochemical response
signal. Further, the selectivity and sensitivity depend mainly on the architecture of the sensing layers.
The design of the latter and the content of the biological recognition elements that are integrated
within or connected to a transducer dominate the response to a specific analyte and subsequently
control the biochemical event [2,5]. The biorecognition systems in biosensors should enable the
formation of complementary structures with high affinity, such as antibody–antigen, enzyme–
substrate, and receptor–ligand pairs. This is important because it is proportional to an analyte’s
concentration [3]. Hence, there is a growing need to utilize nanotechnology at electrochemical
biosensors in order to reduce the dimensions of the sensor elements and improve the signal-to-noise
ratio. Moreover, exploiting (nano)biomaterials for functionalizing the sensor surface can significantly
improve its sensitivity, especially to low-molecular-weight analytes. Several natural biomaterials
provide support to improve the functionality of the sensing system. However, these materials must
also satisfy particular criteria, such as being stable, flexible, biocompatible, and biodegradable. The
biomaterials immobilized on the surface of a biosensor must not interact with the biological
molecules to avoid any undesired physical–chemical interactions and, consequently, interference
with the measured signal [6]. Promising biomolecules are surface (S-) layer proteins, which exhibit
the natural feature to self-assemble into mono- or, if desired into guided double layers in solution
and at various interfaces and surfaces [7,8].
Figure 1. Schematic drawing (not drawn to scale) of the elements of an electrochemical biosensor.
FETs: Field-Effect Transistors.
This review highlights the general approaches for the development of electrochemical
biosensors and focuses on the exploitation of S-layer proteins to functionalize the sensing surface.
The S-layer lattice constitutes a biocompatible intermediate layer for linking biological molecules to
silicon oxide or metal surfaces and hence, decouples biological molecules from the electrode (Figure
2) [9,10]. This is of utmost importance as biological molecules may lose their structure and function
on inorganic surfaces. The pores of the S-layer lattice allow unhindered electron transfer from and to
the electrode surface and additionally provide an ion reservoir.
Sensors 2020, 20, 1721 3 of 22
Figure 2. Schematic drawing (not drawn to scale) of an electrochemical biosensor with an S-layer
lattice as intermediate layer for linking biorecognition elements to the Au (gold) electrode surfaces.
2. Bacterial S-Layer Proteins
2.1. General Features
An S-layer is per definition a “two-dimensional array of proteinaceous subunits forming the
surface layer on prokaryotic cells” (Figure 3) [7]. The S-layer covers the entire cell surface as a closed
layer [11–14], shapes the outermost structure of many bacteria, and are a nearly ubiquitous
characteristics of archaea [7,15–18]. A frequent post-translational modification in S-layer proteins is
protein glycosylation of certain amino acid residues, which is a remarkable characteristic of many
archaeal and some bacterial S-layer proteins [16,19–22]. Table 1 summarizes the most important
features with respect to its application as intermediate layer, ion reservoir and immobilization matrix
in biosensor development.
Figure 3. Transmission electron microscopy image of a freeze-etched and metal shadowed
preparation of (a) an archaeal cell (from Methanocorpusuculum sinense), and (b) a bacterial cell (from
Desulfotomaculum nigrificans). Bars, 200 nm. With permission from Sleytr et al. 2014 [7] (CC BY-NC-
ND 3.0).
Table 1. Summary of the biosensor-relevant properties of S-layer proteins and lattices Reference.
Sensors 2020, 20, 1721 4 of 22
Molecular weight of S-layer protein subunits: 40–200,000 Da 7,23,29
Reactive groups (e.g., carboxyl- and amino-residues) occur on each protomer in identical
position and orientation 7,23,24
Two-dimensional (glyco)protein crystal composed of identical subunits 7,25
Oblique (p2), square (p4) or hexagonal (p6) space group symmetry 7,16,26,27
Center-to-center spacing of unit cells (= morphological units) of crystalline lattice: 3.5–35 nm 7,27
Layer thickness: 5–10 nm 7,8,
High porosity (30%-70%) with pores of identical size (2–8 nm), morphology, and
physicochemical properties 7,24
Topography: Inner surface smooth, outer surface more corrugated 7,8,26
Anisotropic charge distribution between outer and inner face: Outer face charge neutral due to
an equal number of carboxyl- and amino groups. Inner face net negatively charged due to an
excess of carboxyl groups
7,8,24
Antifouling, non-sticky outer surface 7,11,28
Self-assembly capability in aqueous media, on the air/water interface, on lipid films, and on
solid surfaces like metals (gold, silver, platinum, stainless steel), glass, silicon, silicon oxide and
nitride, mica, polymers (e.g., polystyrene, polyester, cellulose, polydimethylsiloxane (PDMS),
indium tin oxide (ITO), highly oriented pyrolytic graphite (HOPG), and carbon nanotubes
7,9,28–30
S-layer proteins have so far suggested to mediate a broad range of specific biological functions,
including protection against (e.g., bdellovibrios, bacteriophages, and phagocytosis) promoters for cell
adhesion (e.g., to host enzymes and cells, immune-modulators, surface recognition, molecular sieve,
molecule and ion traps, antifouling coatings, and virulence factors in pathogenic organisms)
[7,23,31,32]. Moreover, the S-layer lattice is involved in the determination of cell shape and can aid in
the cell division process in archaea possessing S-layers as the exclusive envelope component external
to the cytoplasmic membrane [32–35].
The most significant characteristics of S-layer proteins is the intrinsic ability of natural and
recombinant protein subunits to self-assemble on surfaces or interfaces into crystalline arrays. These
materials comprise glass, silicon oxide and nitride, mica, noble metals like gold, titan, platinum, but
also of stainless steel or many polymers as polystyrene, polyester, and cellulose, and technically
relevant materials like highly oriented pyrolytic graphite (HOPG) or indium tin oxide (ITO) [28,29].
Initiation of crystal growth at interfaces (e.g., solid supports, air–water interface or lipid membranes)
occurs simultaneously at many randomly distributed nucleation points, and proceeds in plane until
the crystalline domains meet. This finally results in the formation of a closed, coherent mosaic of
individual, several micrometer large S-layer protein patches within 45 to 90 minutes [36–39]. The
growth of extended S-layers patches is favored at low monomer concentrations due to the
corresponding low number of nucleation sites. The individual patches are monocrystalline and
separated by grain boundaries. However, these S-layer protein patches form a coherent lattice
thereby coating areas in the cm²-range.
In general, S-layer protein lattices constitute a unique immobilization matrix for presenting
diversified biomolecules in nanometer distances [7,40,41]. Moreover, the S-layer protein lattice acts
as a versatile scaffolding for the formation of supported lipid membranes. The latter provide the
essential ambience for the insertion of membrane-active peptides and the reconstitution membrane
proteins [9,42–44].
2.2. Antifouling Properties
An important biological function of the S-layer lattice is to prevent fouling of macromolecules
on the cell surface. Electron micrographs showed no adsorbed molecules on the S-layer lattice even
for cells harvested from complex environments or growth media containing a great variety of
macromolecules [11]. In addition, S-layers from thermophilic bacteria did not adsorb charged
macromolecules on their surface or inside the pores because this would hinder the transport of
nutrients and metabolites [45–47]. Based on these results, S-layers proteins can be considered as
structures with excellent “antifouling” properties.
Sensors 2020, 20, 1721 5 of 22
The antifouling characteristics of the S-layer protein lattice has not only investigated in vitro
with S-layer ultrafiltration membranes (SUMs) [48–50], but also with coatings in electrochemical
microfluidic biochips [28]. Cyclic voltammetry (CV) measurements using ferricyanide/ferrocyanide
(FCN; [Fe(CN)
6
]
3/4
) as the redox system allowed the determination of the peak current of gold
electrodes with and without an S-layer lattice cover in the presence of human serum albumin (HSA;
30 mg/mL). Interestingly, HSA adsorption led to a much higher clogging of the plain gold electrode,
as the CV measurements resulted in a more than 96% reduced peak current for the plain electrode
(0.3 µA), as compared to the electrode with the recrystallized S-layer protein (12,8 µA) [28]. Moreover,
surface plasmon resonance (SPR) analysis showed that the adsorption of HSA (60 mg/mL) and
human blood components is much higher for plain gold surfaces (3.74 and 4.44 ng/mm
-2
for HSA and
human blood, respectively) than for S-layer protein covered gold surfaces (0.18 and 0.46 ng/mm
-2
for
HSA and human blood, respectively) [28]. Hence, these two in vitro studies demonstrate imposingly
the antifouling characteristics of S-layer protein lattices on metal electrodes.
The antifouling property of the S-layer protein lattice was also compared to other favored
immobilization supports, which are polyethylene glycol (PEG) and bovine serum albumin (BSA). As
determined by SPR, the S-layer lattice showed again the lowest amount of adsorbed HSA (0.18
ng/mm
-2
) followed by PEG (0.76 ng/mm
-2
) and BSA (1.92 ng/mm
-2
) [28].
2.3. Electrochemical Properties
As previously mentioned, the S-layer lattice is a porous structure with a porosity up to 70%
[7,24]. The latter feature additionally provides an ion reservoir. Figure 4 shows the cyclic
voltammogram of an S-layer protein recrystallized on a gold electrode in the presence of the FCN
redox system and 100 mM KCl. The bare electrode surface has the highest peak current, which
indicates a small resistance in the electron transfer. Subsequently, recrystallization of the S-layer
protein onto the gold sensor leads to a slightly reduced peak current, which attributes to a decrease
in the electron transfer resistance. Thus, the S-layer lattice reduces the electrode area to some extent,
but the porous intermediate layer allows for efficient redox-mediator diffusion and unhindered
electron transfer from and to the electrode surface, thus the sensitive detection of subsequent
biorecognition events. The resistance and capacitance of the S-layer lattice itself is negligible. Hence,
further binding events of analytes or (cancer) cells at different quantities results in measurable
changes in the electrochemical properties of the sensing layer.
Figure 4. Cyclic voltammogram of the bare and the S-layer protein-coated gold electrode in 10 mM
[Fe(CN)
6
]
3/4
containing 100 mM KCl at a scan rate of 50 mV/s.
In general, proteins are biopolymers comprised of amino acids. That is why it is not surprising
that the electrical properties of the surface onto which the S-layer protein is self-assembled has a
Sensors 2020, 20, 1721 6 of 22
pronounced effect on the lattice formation. The electrochemical behavior of the S-layer SbpA isolated
from Lysinibacillus sphaericus CCM 2177 on gold electrodes was studied, beside spectroscopical and
microscopical techniques, by in-situ electrochemical quartz crystal microbalance [51–53].
The self-assembly and bond formation of S-layer proteins on a gold surface can be affected and
controlled by direct electrochemical manipulation. The S-layer proteins self-assemble to a crystalline
lattice in the positively-charged electrochemical double layer region where solvated anions form the
electrochemical double layer. In this region, the carboxyl groups of the S-layer protein can interact
with the positively-charged gold atoms on the electrode surface. In contrast, there is no lattice
formation in the negatively-charged electrochemical double layer region [51]. In another study, the
S-layer lattice of Deinococcus radiodurans shows ion-gating properties as demonstrated by EIS [52]. Ion
transport appears to be mainly due to an electrical gradient inside the pores, presuming to originate
from the negative charges, which are present on this S-layer lattice. A detailed study on the gating
characteristics of this nanoporous structure toward various ionic species suggested, that the
immobilized S-layers experience a strong interaction with cations, particularly Ca2+-ions. The latter
interlink laterally with the S-layer proteins lattice in the positively-charged electrochemical double
layer region, thus facilitate the formation of the crystalline S-layer protein lattice [53,54].
An aluminum—S-layer protein (from Lactobacillus brevis ATCC 8287)—ITO/polyethylene
terephthalate device structure allowed the examination of the nonvolatile memory characteristics of
this S-layer protein by CV [55]. The S-layer protein is very interesting for this approach, beside other
intrinsic features, because it is redox inactive. It turned out that the nonvolatile resistive switching
characteristics of the S-layer protein can be utilized for active memory storage in a powered-off state
of electronic chips. Thus, this flexible memory device could find a way in wearable storage
applications like smart bands and sports equipment sensors [55].
3. Basic Principles of Electrochemical Biosensors Used in Combination with S-Layer Proteins
Electrochemical measurements depend on the spontaneous interaction between electrical
energy and a chemical reaction involving an oxidation–reduction reaction to generate an electrical
current or vice versa. The chemical events that occur between immobilized biomaterials and the
analytes result in the production/consumption of ions or electrons, which has an effect on the
electrical current, the electrical potential, or any other electrical property of the solution. These
reactions take place at the interface of a metal or semiconductor electrode and an electrolyte [56–58].
Thus, detection is feasible if the reactions occur in close contact with the electrode surface. Hence, the
electrodes significantly influence the performance of the electrochemical biosensor. One has to
consider several factors when choosing a proper electrode, including its material, dimension, and
possibility to carry out surface modifications. Most electrochemical cells are composed of three
electrodes (Figure 5):
A reference electrode (RE): This electrode is usually made of Ag/AgCl and stands at a distance
from the place where the reaction occurs to provide a potential that is proportional to the known and
stable solution. The RE allows normalizing of the measurements.
A counter (auxiliary) electrode (CE): This electrode is the source of the current, which is
subsequently applied to the working electrode.
A working electrode (WE) (the sensing or redox electrode): This electrode acts as the
transduction element in the biochemical reaction. CE and WE should be conductive and chemically
stable. Thus, depending on the analyte and the nature of the reaction, the mainly used electrode
materials are gold, silver, platinum, silicon, carbon, and graphene. An alternative to electrochemical
cells is to screen-print the three electrodes on an insulating substrate. These so-called screen-printed
electrodes have several advantages, including simplicity, ease of mass production, low-cost
construction, and low analyte/reagent consumption [2,59,60].
Sensors 2020, 20, 1721 7 of 22
Figure 5. The two types of electrochemical biosensors with three electrodes: reference (RE), working
(WE), and counter (CE) connected to a potentiostat.
Classification of electrochemical biosensors occurs according to their characteristics, such as
biorecognition element, the biological mechanisms, the electrochemical detection technique, or
combinations of these approaches. In the following section, we briefly discuss classes of
electrochemical biosensors based on the mode of signal transduction. Modification on some types of
these electrochemical biosensors occurred either by the self-assembly of isolated S-layer protein
subunits on the sensor surface or by using a SUM.
3.1. Amperometric Biosensors
Amperometry means to measure the current that results from the electrochemical oxidation or
reduction of an electroactive species. Applying a constant potential at the WE or on an array of
electrodes with respect to the RE induces the generated current, which is associated with the redox
process. In amperometric devices, the term “amperometry” refers to the technique characterized by
monitoring the current at a constant potential. In turn, the term “voltammetry” refers to a technique
monitoring the current during variations in the potential over a certain potential range. In both
techniques, the measurable current correlates directly with the production/consumption rate of the
electroactive species or to the bulk analyte’s concentration [1,2,6,61]. In this amperometric approach,
also SUMs and S-layer proteins recrystallized directly on the gold electrode constitute key
components for the immobilization of enzymes (see Table 2) [28,62,63].
3.2. Potentiometric Biosensors
Potentiometric devices monitor the accumulation of a charge potential either at the WE
compared to the RE or between two REs separated by a semipermeable membrane. Monitoring the
potential occurs in an electrochemical cell under the condition that no significant current flows
between the two electrodes. The transducer is usually an ion-selective electrode (ISE) composed of
an electrochemical sensor with a recognition element based on a thin film or a selective membrane.
The conversion of the biorecognition event into a potential signal by the ISE gives rise to a measurable
signal [64].
3.3. Conductometric Biosensors
Conductometry is based on the relationship between a biorecognition reaction and conductance.
Changes in the concentration of ionic species or of ionic strength in the sample result in an alteration
in the electrical conductivity of the solution or in the current flow. Hence, conductometric devices
Sensors 2020, 20, 1721 8 of 22
measure the capability of an analyte or a medium, such as an electrolyte solution or nanowires,
respectively, to conduct an electrical current between two metal electrodes (e.g., Ag or Pt). An
ohmmeter measures the changes in conductance between these electrodes. Some studies consider
conductometric biosensors to be a subset of impedimetric biosensors. However, conductometric and
impedimetric devices constitute convenient sensors to investigate enzymatic reactions that generate
changes in the concentration of charged species in a solution [3,65,66]. Both techniques use nowadays
interdigitated microelectrodes where the enzymatic reaction results in changes in the ionic strength
and conductivity of a solution between two electrodes.
3.4. Field-Effect Transistors (FETs)
FETs are devices composed of three electrodes: source, gate, and drain. These types of transistor
use an electric field to control the conductivity of a channel where a depletion region of charge
carriers lies between the source and drain electrodes in a semiconducting material. Variations in the
electric field potential relative to the source and drain electrodes at the gate electrode contribute to
controlling the conductivity. The configuration and doping of the semiconducting material govern
the charge carrier’s behavior in the conduction channel, i.e., at the gate electrode. The presence of an
appropriate positive or negative potential may either attract or repel charge carriers, such as
electrons, in the conduction channel. This may result in the filling or emptying of the depletion region
of charge carriers and, subsequently, form or deform the effective electrical dimensions of the
conduction channel. Hence, the conductance between the source and drain electrodes can be
controlled. There are several types of FET devices for the measurement of ion concentrations;
however, the most popular are ion-sensitive field-effect transistors (ISFETs) and enzyme field-effect
transistors (EnFETs) [1,2,67].
3.5. Impedimetric Biosensors
Impedance techniques are efficient tools for measuring alterations in electrical properties that
result from biorecognition reactions at a modified electrode’s surface. Electrochemical impedance
spectroscopy (EIS) is one of the most useful tools for constructing biosensors that can monitor the
current response to an applied potential. Complex impedance is the sum of the real and imaginary
impedance components, resistance and reactance, respectively, of a system as a function of the
angular frequency (ω). Calculation of the impedance occurs by altering the excitation frequency (f) of
the applied potential over a range of frequencies. EIS can be used to investigate the properties of a
material or specific processes that can affect the capacitivity or conductivity/resistivity of an
electrochemical system [2,4,68]. EIS technique is widely used in electrochemical biosensors based on
S-layer proteins for several applications to assess membrane-active peptide insertion and membrane
protein reconstitution into artificial lipid membranes and specific cell detection [69–73].
4. Sensor Surface Modifications
The construction of an ideal biosensor involves two areas: bioactive and inert. The bioactive area
is the region where the biochemical reactions and detection events take place. The inert area is the
one, which surrounds the bioactive one. Both areas must be composed of carefully selected elements
in order to avoid interference of reactions and non-specific adsorption. The bioactive surface needs a
smart design to improve the sensing sensitivity and to minimize sample consumption. However,
functionalization of an active surface requires conquering several challenges. The molecules that are
adsorbed/attached onto the surface must be in a suitable pattern and at an appropriate density to
avoid an insufficient number of molecules or crowding on the sensing area, which may influence
analyte capture and the sensor’s functionality (Figure 6). Further, the developed biosensor must be
stable and not be affected by environmental conditions such as time frame and storage.
Sensors 2020, 20, 1721 9 of 22
Figure 6. Direct and indirect transduction (not drawn to scale). In direct approach, the electron
transfers are close to the surface, whereas in the indirect one, electron shuttles between the reaction
site and the sensor surface. In the proposed model, the S-layer protein lattice constitutes an
intermediate matrix. In the lipid-based biosensor (left), electrons transfer from the outer membrane
to the inner membrane and vice versa via a channel protein. In the detection biosensor (middle and
right), electrons transfer between the enzyme–substrate complex and cell/antibody and electrode
surface, respectively. The S-layer lattice provides an immobilization matrix and ion reservoir. The
pores of the S-layer lattice ensure no impact on the electron transfer. Fc: fragment crystallizable;
rSbpA/ZZ: recombinant S-layer protein from Lysinibacillus sphaericus CCM 2177 with fused Fc-binding
Z-domain (synthetic analog of immunoglobulin G (IgG-binding B—domain) of protein A of
Staphylococcus aureus).
Materials to prepare the sensing substrate are manifold. However, these materials must fulfil the
special requirements of electrochemical measurements, such as electrical conductivity. Most suitable
substrates are carbon, graphene, carbon nanotubes, gold, and platinum, depending on the
measurement technique and nature of the analyte itself. For example, carbon nanotubes offer many
advantages, including good conductivity, high chemical stability, a high surface area, and high
mechanical strength [74]. Further, metallic surfaces, such as a gold surface can be modified with self-
assembled monolayers (SAMs) of thiols in order to allow molecules to be adsorbed in a highly
ordered and well-organized pattern while maintaining electron diffusion. Furthermore, during
fabrication of an electrochemical biosensor, it is usually favorable to coat the WE with an extra layer
as a spacer or mediator to enhance the immobilization of biomolecules in a well-oriented manner.
The intermediate layer can provide another advantage, which is to minimize non-specific binding.
Polymers (e.g., chitosan, agarose, hydrogels, etc.) and proteins (e.g., S-layer proteins) are convenient
to functionalize sensor surfaces and exhibited a stable matrix for a lipid-based biosensor suitable for
studying and utilizing membrane protein as sensing elements [9]. In contrast, the intermediate layer
influences the sensitivity of the sensor because an increase in the distance to the surface corresponds
to a decrease in the signal (Figure 3). In electrochemical biosensors, the reaction needs to occur close
as possible to the electrode surface because the products can diffuse in many directions away from
the surface, which can contribute to a lower signal. Thus, a careful choice of the spacer or mediator is
necessary to ensure that these molecules act as a shuttle transferring electrons between reaction site
and electrode surface [2,75,76].
Sensors 2020, 20, 1721 10 of 22
The fabrication of lipid membrane-based biosensors used with electrochemical techniques
depends mainly on the composition of the lipids and proteins. Biological membranes are the most
important electrified interfaces in living systems. Due to the hydrophilic head and hydrophobic tail,
lipid molecules are amphiphilic and, hence, can be self-assembled to form a bilayer that consists of
two monolayers. In this configuration, the non-polar tails point toward the hydrophobic interior of
the bilayer and polar head is oriented toward the external aqueous surrounding. The generated lipid
matrix can further host proteins that act as receptors, pores or channels. Natural and synthetic
membranes are selective to permeable ions that cross between the inside and outside of membranes.
Further, high electrochemical potentials can be maintained across the interior and exterior
monolayers [72,73,77,78]. The chemical nature of the membrane components and the reactions that
occur at the interface of the monolayer or within the bilayer govern the permeability and structural
properties of the membrane. Many membrane proteins in lipid membranes are electrogenic and
translocate a net charge across the membrane. Thus, monitoring the functions of membrane proteins
occurs directly by measuring the flow of current along an external electrical circuit when these
protein molecules are in an activated state. Further, monitoring of deactivation of proteins occurs also
by tracking the drop in current after addition of a chemical agent, such as a drug (agonist or
antagonist) [77]. A well-known model that shows how the electrical and permeability characteristics
of a lipid membrane and its protein content are affected is the partition of anesthetics into a lipid
membrane. The response of a biomembrane to the presence of anesthetics or toxins is usually rapid,
reversible in a concentration-dependent manner, and may be voltage-dependent. The first step in
building a lipid membrane-based biosensor is the choice of the lipid composition. A natural cell
membrane is negatively charged and, hence, exploiting negatively-charged lipids is a challenging
approach to create artificial but biomimetic lipid membranes. The choice of lipid is critical as it may
further influence the protein reconstitution and membrane activity [79]. The generation of a synthetic
membrane on a negatively-charged surface, such as mica or silicon can result in repulsive electrostatic
interactions that disturb lipid bilayer formation, [80,81]. This is in particular the case if the pH and
ion strength of the surrounding solution is not adjusted appropriate [80]. However, the addition of
positively-charged molecules, such as positively-charged lipids and peptides/proteins may foster the
integrity to the membrane. Moreover, alterations in the pH affect the degree of ionization, resulting
in changes in lipid pKa values and consequently the charge of the lipid molecule [81,82]. Insertion of
membrane-active peptides is in general not difficult as these peptides are soluble (e.g., in alcoholic
solutions). The insertion of membrane proteins is a big challenge, as with few exceptions membrane
proteins have to be solubilized by means of detergents [83]. The reconstitution of membrane proteins
in a synthetic membrane, however, can nowadays also be performed either by cell-based or cell-free
protein expression systems. Each system offers advantages and disadvantages as extensively
discussed in a number of publications [84–86]. The formation of a lipid membrane-based biosensor
can occur directly on a metal electrode or indirectly by a matrix layer. The latter is favorable as it can
minimize direct contact between the reconstituted membrane protein and the substrate, which can
further minimize protein denaturation. Additionally, it represents a membrane model that is robust,
highly hydrated, and has long-term stability.
Another application that exploits an electrochemical technique is detection biosensors. These
devices inherently possess biorecognition elements that specifically recognize a target analyte. The
biorecognition molecules can be immobilized directly onto the WE by physical adsorption or onto a
modified matrix based on chemical interactions. This should ensure correct orientation and to
minimize random adsorption, which can further improve the capture efficiency. Antibodies,
nanobodies, aptamers, affimers, and receptors are examples of used biorecognition elements with
high affinity and specificity for particular antigens and cells [87]. Quantitatively, the formation of a
complex between the biorecognition molecule and its target results in a measurable signal. Moreover,
biosensing devices expedite the exploitation of enzymes as electrocatalysts. Oxidation or reduction
reactions with the substrate generate products, which give rise to a detectable signal. Similar to
detection sensors with biorecognition molecules, enzymatic biosensors can be constructed either by
direct enzyme adsorption onto the sensor surface or on a mediator/intermediate matrix. The most
Sensors 2020, 20, 1721 11 of 22
common enzymatic biosensors on the market are the ones composed of glucose oxidase for the
detection of blood glucose levels in diabetic patients [33,88–90].
5. Application of S-Layer Proteins in Biosensors
5.1. S-Layer Protein on Gold Surfaces
The utilization of S-layer proteins as the sensing element for detecting metal ions using EIS is a
straightforward approach. The immobilization of the S-layer protein from L. sphaericus JG-A12 on a
gold surface occurred by covalent binding of intrinsic cysteine residues of the S-layer protein to a
SAM presenting terminal maleimide groups (Table 2). Another possibility is to link the biotinylated
S-layer protein via neutravidin to a mixed SAM that presents biotin molecules as well [91]. The single
intrinsic S-layer proteins immobilized on the electrode surface acted as specific binding sites for
aqueous uranyl ions (UO22+). In this approach, the S-layer proteins have not assembled to a crystalline
lattice. Both, phosphate and carboxyl groups of the S-layer protein are involved in the recognition to
UO22+. The limit of detection for UO22+ ions was determined to 10-12 M. Moreover, minor interference
from Ni2+, Cs+, Cd2+, and Co2+ ions are detectable. [91]. A straightforward approach might be to collect
S-layer carrying organisms from metal polluted environments in order to use their isolated S-layer
proteins as sensing elements for the detection of heavy metal ions or radioactive ions. This is in
particular a field of application for monitoring environmental pollution.
Table 2. Summary of electrochemical biosensors based on S-layer proteins.
Electrode
Architecture
Immobilization
Method
Biorecognition
Element
Detected
Species
Electro-
Chemical
Method
Tested
Detection
Range
Linear
Range Stability Remark Reference
Au-SAM- SLP Maleimide-Cys;
biotin-avidin SLP UO22+ EIS 10-5–10-12 M 10-5–10-8
M N.D. LOD 10-12 M 91
Au/SLP-GOx Chemical (EDC) GOx + FCN glucose Amperometric 0.5–50 mM
0.5–50
mM 2.2 h blood, HSA,
plasma 28
Au/Pt-GOx-
SUM Chemical (EDC) GOx glucose Amperometric 2–20 mM Up to
12 mM 48 h Response
10–30 s 62
Au-AlcOx-
SUM Chemical (EDC) alcohol oxidase ethanol Amperometric - Up to 7
mM N.D.
Signal: 2.5
µA cm-²mM-
1
63
Au-XanOx-
SUM Chemical (EDC) xanthine oxidase xanthine Amperometric - Up to
0.6 mM N.D.
Signal: 30
µA cm-²mM-
1
64
Au-Maltase/
GOx-SUM Chemical (EDC) maltase + GOx maltose Amperometric - Up to
1.5 mM N.D.
Signal: 1.5
µA cm-²mM-
1
64
Au-Inv/Mut
/GOx-SUM Chemical (EDC)
invertase +
mutarotase +
GOx
sucrose Amperometric 1–35 mM Up to
30 mM 36 h Response
300 s 92
C-ChOx/ SLP Mixed multi-
layers ChOx cholesterol CV 3.1 mM N.D. N.D. Langmuir/
Blodgett 93
Au-SLP/ZZ-
anti-Ab
ZZ-domain + anti-
CD133
anti-CD133
antibody
Liver cancer
cells (HepG2) CV 1 x 105–6 x
106 cells
Up to
6x106
cells
N.D.
S-layer
fusion
protein
72
Au-SLP-folate Chemical (EDC) folate Breast cancer
cells (MFC-7) SWV 1 x 104–5 x
105 cells N.D. N.D. LOD 1 x 105
cells/mL 73
Ab—antibody; Au—gold; Cys—cysteine; EDC-1-ethyl-3—(3-dimethylaminopropyl)carbodiimide; FCN—ferrocyanide/ferricyanide; HAS—human serum albumin; LOD—limit
of detection; N.D. —not detected; SAM—self-assembled monolayer; SLP—S-layer protein; SUM—S-layer ultrafiltration membrane; ZZ—Fc-binding Z-domain, a synthetic
analog of immunoglobulin G (IgG-binding B—domain) of protein A of Staphylococcus aureus.
5.2. S-Layer Protein and Enzymes
The intrinsic permeability characteristics of self-assembled S-layer protein lattices of diverse
Bacillaceae is useful to produce SUMs, which serve as isoporous molecular sieves with a pore
dimension of 4 to 5 nm [45,46,48–50]. SUMs are not only proper structures for filtration applications
where a sharp cut-off is needed, but also suitable as immobilization matrix for functional molecules
(Figure 7). Beside enzymes (glucose oxidase (GOx), β-glucosidase, glucuronidase, cholesterol esterase
and oxidase, mutarotase, invertase, maltase, xanthine oxidase, alcohol oxidase, naringinase,
peroxidase, laccase, etc.), also ligands (protein A, streptavidin) or mono- and polyclonal antibodies
have been immobilized as densely-packed sensing layers on SUMs [7,47,48,63,94,95].
Sensors 2020, 20, 1721 12 of 22
Figure 7. Generalized scheme of the construction of an S-layer ultrafiltration membrane (SUM)-based
biosensor (not drawn to scale). S-layer carrying fragments or self-assembly products are deposited on
a commercially available microfilter in a pressure-dependent process. The enzyme is deposited on the
S-layer surface and chemically linked to the S-layer protein. The immobilized enzyme is contacted by
a thin metal layer (PLD: pulse-laser-deposition). Finally, this composite structure is deposited with
the metal layer side on the working electrode (WE) and mounted in a flow cell (RE: reference
electrode, CE: counter electrode). The analyte is pumped in the flow cell. After passage across the
modified SUM, the analyte reacts with the enzyme, which is detected by amperometry.
The first amperometric sensor based on an S-layer protein lattice, developed 27 years ago, was
comprised of an SUM with covalently bound GOx [62]. Table 2 shows a summary of the performance
of SUM-based biosensors. The immobilization of the GOx on the SUM caused an activity reduction
to approximately 40% [62]. Hence, improvements were necessary. In a first attempt, the transducer
was modified to an oxygen optode with optical read-out system [96]. The recrystallization of an S-
layer protein improved the fiber optic probe as the GOx molecules could be bound in a very close
proximity and its tightest packing. The next improvement was to replace the argon sputtering
method for metallization of the composite S-layer protein/GOx layer by the pulse-laser-deposition
technique (Figure 7) [97]. Indeed, this technique achieved to retain the enzyme activity to up to 80%,
which represents a two-fold GOx activity with reference to the first biosensor utilizing the S-layer
protein lattice as binding matrix for the GOx molecules (Table 2) [62].
The continuous, stable, and accurate determination of glucose in biological fluids is a
challenging approach particularly in blood because many constituent parts interfere the
measurement and monitoring. S-layer protein lattices provide highly hydrated, nm-thick, biological
antifouling materials [98]. Moreover, SbpA has the ability to form monomolecular lattice structures
at various microchip surfaces (e.g., glass, silicone oxide, polydimethylsiloxane, platinum, and gold)
within one hour. To utilize these features, the idea came up to develop a microfluidic device
consisting of four individually addressable amperometric biosensor arrays, each coated by an S-layer
lattice [28]. In this arrangement, the S-layer lattice has multiple tasks, which are to offer a reduced
unspecific binding of blood components, an anchoring structure for the enzyme in a way that the
catalytic sites of GOx are accessible, and to act as an efficient molecular sieve. The flow cell
configuration made it possible to measure simultaneously the concentration of the glucose in blood,
perform auto-calibration routines, detect mediator-interferences, and subtract the background. The
lab-on-a-chip device precisely measured the glucose concentration in blood (Table 2). The latter is in
particular important during hemodialysis because it allows individual balancing of the glucose level.
The highly porous S-layer protein coating eliminated unspecific adsorption events in the presence of
human serum albumin, human plasma and freshly drawn blood samples. Most important, the
undisturbed diffusion of the mediator to the electrode surface enabled bioelectrochemical
measurements of glucose in the concentration range of 500 µM to 50 mM [28].
Sensors 2020, 20, 1721 13 of 22
To generate a biosensor for sucrose (Table 2), immobilization of the three enzymes GOx,
mutarotase, and invertase on S-layer protein self-assemblies from Clostridium thermohydrosulfuricum
L111-69 occurred using spacer molecules comprising aspartic acid [94]. The S-layer protein self-
assemblies carrying the different enzymes were mixed and deposited on the microfilter. This resulted
in a final structure shown in Figure 7c, but with three immobilized enzymes. Hence, this is a further
example demonstrating that sandwich structures comprising recrystallized S-layer
protein/enzyme/metal layer constitute seminal biosensor systems with high efficiency.
In order to measure cholesterol in biological fluids, biosensors often take advantage of the
enzyme cholesterol oxidase (ChOx) [63,99]. A coating method that has proven itself as very useful
and straightforward is based on the production of monomolecular films at the air–water interface. It
was possible to generate films out of ChOx and mixtures of ChOx with S-layer proteins at the water–
air interface [100]. The mixed films showed the advantage of being stable over a longer period
compared to the pure ChOx films. The Langmuir–Blodgett transfer of mixed ChOx/S-layer protein
films on screen-printed carbon electrodes enabled to study their possible application in biosensor
devices (Table 2) [95]. Atomic force microscopy demonstrated the successful transfer of films, which
caused a reduction of the roughness of the electrode surface as well. CV measurements revealed
increased oxidation peak intensity and reduced oxidation potential for the mixed films compared
with the pure enzyme films. Moreover, the presence of S-layer proteins reduced the potential where
oxidation occurs, making the sensor more selective, since a higher potential will oxidize non-wanted
species, which can interfere with cholesterol detection. Hence, screen-printed carbon electrodes
combined with a monomolecular ChOx/S-layer protein film constitute a promising biosensor for
measuring cholesterol [95].
5.3. Biosensor for Sensing Cells
The lack of effective diagnostic tools contributes directly to the high mortality rate of cancer.
Early detection of cancer can significantly contribute to an effective treatment and increase the
survival rate of patients. The currently used diagnostic techniques include polymerase chain reaction,
flow cytometry, immunohistochemistry, immunofluorescence, and mass spectrometry. These assays
are usually costly, time-consuming, labor-intensive, require several steps and multiple reagents, and
suffer from some limitations to their selectivity and sensitivity. In contrast, electrochemical
biosensors have attracted attention as a promising alternative to the conventional methods. An
additional advantage that electrochemical technology provides is the ability to detect a whole cell or
an intact cell by identifying the expression of a surface marker on the cell membrane, which excludes
the requirement for pre-processing steps. Moreover, besides their ability to detect cancer cells, cell-
based biosensors provide an efficient and simple approach for the investigation of cell viability, cell
proliferation, and the effect of chemotherapeutic drugs on cancer cell viability [101,102]. Several
successful studies describe the capture/recognition of cancer cells at the electrode surface using
various electrochemical methods, including CV [103], square wave voltammetry (SWV) [104],
differential pulse voltammetry (DPV) [105], and EIS [106].
The careful design of the interface architecture of an electrochemical biosensor is very important
to ensure high performance, sensitivity, and selectivity. The main factor that restricts the biosensor
efficiency is the orientation and density of the affinity biomolecules (recognition elements (e.g.,
antibody, nanobody, aptamer) and antigens (e.g., cell)) on the sensing area [107,108]. For example,
immobilization of antibodies onto WEs is possible by approaches that involve physical adsorption
and chemical interactions. Despite the extensive use of these methods, they still suffer from serious
drawbacks. Physical adsorption can result in the immobilization of antibodies in a random
orientation, while covalent coupling based on chemicals (e.g., glutaraldehyde, carbodiimide) or a self-
assembled monolayer (e.g., nanoparticles, protein A, protein G) may alter the biological activity of
the antigen-binding sites of the antibody [8,109,110].
To overcome these limitations, the recombinant S-layer fusion protein rSbpA/ZZ has been
exploited to develop a detection biosensor for circulating tumor cells (CTCs) that overexpress CD133.
CD133 is a five-transmembrane, single-chain glycoprotein expressed in various human epithelial
Sensors 2020, 20, 1721 14 of 22
carcinomas, including lung, colon, liver, and gastric carcinoma [111]. Functionalization of the
biosensing area of the fabricated sensor occurred by immobilization of rSbpA/ZZ and anti-CD133
antibodies. rSbpA/ZZ has two copies of the 58-amino-acid-long Fc-binding Z-domain, a synthetic
analog of the immunoglobulin G (IgG)-binding B-domain of protein A of Staphylococcus aureus
[112,113]. The protein A binding site of rSbpA/ZZ helps to immobilize the antibodies (IgG) in the
correct orientation [113]. Thus, rSbpA/ZZ, when used to generate a uniform matrix on a gold
electrode, enables the immobilization of anti-CD133 antibodies in the correct orientation to enhance
the recognition and capture of the surface marker CD133, which is expressed on membranes of
hepatic cancer stem cells [72]. The CV technique allowed evaluation of the functionality of the
developed sensor. Voltammetric measurements converted the interactions between the anti-CD133
antibody and the hepatic cells into an electronic signal. The peak current decreased consistently in
response to an increasing number of captured cells on the electrode surface (Table 2). The importance
of the development of biosensors for the detection of CTCs derives from their roles in cancer
diagnosis, as they can provide an evaluation of a treatment strategy, its effectiveness, and the
possibility of cancer recurrence after treatment. The main obstacle for the capture of CTCs is their
limited number in the blood (1–100 cells/mL blood) [114].
In a recent study, modification of the S-layer protein SbpA occurred by chemical binding of folic
acid. This SbpA-folate construct self-assembled to form a monomolecular S-layer lattice on a gold
surface [73]. The S-layer protein lattice offered an even distribution of the folate residues on the
surface of acoustic and electrochemical biosensors. This arrangement ensured the specific capture of
human breast adenocarcinoma cells (MCF-7), which expose folate receptors on their cell surface.
Indeed, clear differentiation between human liver hepatocellular carcinoma (HepG2) cells, which do
not express folate receptors and MCF-7 cells, was feasible (Table 2). The thinness and antifouling
characteristics of the SbpA lattice caused an increased efficiency for cell capturing and no need to
block the surface to prevent unspecific binding, respectively. Beyond, there is no necessity for
immobilizing antibodies because the folate residues constitute an alternative to antibodies for
capturing target cells. Over all, evaluation of the developed biosensors by different techniques
provides more information about the efficiency of the system. QCM-D measurements tracked the
formation of SbpA-folate modified sensor and capturing of cancer cells efficiently in real-time and
under controlled conditions. Although the QCM-D technique shows a limited detection range, it
allows tracking the cell viability [73]. Hence, it is worth to perform further QCM-D studies to
elucidate the cellular response to chemotherapeutic agents. In addition, electrochemical recordings
substantiate the selectivity and specificity of cell binding by this engineered biosensor. SWV
measurements demonstrated a direct relationship between the peak current and the number of cells
captured on the sensor surface. The peak current increased to a more negative value as the cell density
increased, indicating a rising coverage of the electrode’s surface by the captured cells [73].
Interestingly, the same density of HepG2 cells, which lack the folate receptor, did not show a
significant change in the peak current. Thus, modifying biomaterials that enable the presentation of
ligand molecules in an optimal density and orientation can result in efficient biosensors with high
selectivity.
Continuing research to identify efficient recognition elements and careful design of surface
architectures may help us to make further progress in the fabrication of biosensors. This, in turn, may
provide us with functional platforms to detect cancer and assist with early diagnosis. Moreover, this
biosensor offers a straightforward, cost-efficient and disposable check-up for the early detection of
cancer.
5.4. S-Layer Protein and Functionalized Lipid Membrane
Archaea, which constitute one of the three evolutionary domains that make up all life on our
planet are small single-celled organisms. They thrive in extremely harsh environments on Earth, such
as sulfuric hot springs or deep-ocean hydrothermal vents that reach 386 K under colossal pressure
[79,115–117]. The cell envelope structure of some archaeal species (e.g., Sulfolobus spp.) constitute
simply of a cytoplasm membrane comprising of ether lipids and a membrane-anchored S-layer lattice
Sensors 2020, 20, 1721 15 of 22
[17,79]. The latter may provide mechanical and osmotic cell stabilization [118,119]. In a biomimetic
approach, the cell envelope structure of archaea constitutes the construction manual for S-layer-
supported lipid membranes (SsLMs) [17,32,79]. Due to the unique interaction of archaeal S-layer
proteins with the cytoplasm membrane, no suitable methods for disintegration of archaeal S-layer
protein lattices and the reassembly into monomolecular arrays on lipid films are available. Thus, the
building blocks for the generation of SsLMs are isolated or chemically-synthesized phospholipids
and isolated or genetically-produced S-layer proteins from Gram-positive bacteria (Figure 6, indirect
transduction, outer left part) [9,42–44,120,121].
With respect to biosensor development, a stable lipid membrane is a prerequisite to make use of
membrane-active peptides and membrane proteins as sensing elements. Indeed, many studies have
proven that the attached S-layer lattice increases the mechanical stability of the lipid membrane
significantly without impairing the fluidity of the membrane [9,42,43]. Membrane-active peptides as
well as membrane proteins need the hydrophobic matrix to be able to adopt their native structure
and hence, their functionality. The porous S-layer lattice functions as (1) stabilizing and anchoring
scaffold for the lipid membrane; (2) biocompatible layer linking biological molecules with silicon
oxide or metals; (3) intermediate protein layer, which decouples the lipid membrane from the solid
support; (4) ion reservoir, which permit long-time electrochemical measurements; and (5) spacer,
which allows the incorporation of bulky membrane proteins [9].
Up to now, most studies with SsLMs deal with the functional characterization of reconstituted
membrane-active peptides and membrane proteins. In turn, the latter may act as sensing element if
a ligand or other molecule interacts specifically with them and cause modification of the ion flow
through the respective membrane-active peptide and membrane protein.
The membrane-active peptide alamethicin forms channels in lipid membranes as well as in
SsLMs on solid supports. Amiloride, an acid-sensing ion channel blocker inhibits these channels in a
concentration dependent manner [122]. Increasing amounts of the ion channel blocker gave rise to a
significant increase in membrane resistance as determined by EIS [71]. In turn, alamethicin channels
reconstituted in SsLMs constitute sensing elements to determine the amiloride concentration in the
surrounding aqueous milieu. Thus, this example shows proof of concept for the applicability of these
composite S-layer/lipid structures for biosensing purposes.
SsLMs, generated by the europium-induced vesicle fusion technique [123] provided a proper
matrix for the insertion of gramicidin [69] and the voltage-dependent anion channel (VDAC) [70]. EIS
studies probed the function of both, the membrane-active peptide gramicidin and the membrane
protein VDAC. The incorporated VDAC caused a decrease in membrane resistance but the
membrane capacitance did not vary significantly [70]. Moreover, the membrane resistance dropped
down with increasing VDAC concentration indicating that an increasing number of functional
channels spontaneously reconstituted into the SsLM [70]. VDAC is a voltage-gated channel, which is
in the open state at a low membrane potential (˂ 10 mV) and switches to the closed state at high
membrane potentials [124–127]. Indeed, the VDAC channels incorporated in SsLMs clearly showed
this gating behavior. Furthermore, the membrane resistance decreased again after reducing the
voltage from 10 mV back to zero. Moreover, the nucleotides nicotinamide adenine dinucleotide
hydride (NADH) and nicotinamide adenine dinucleotide phosphate hydrogen (NADPH) are noted
VDAC channel blockers [128–131]. Indeed, NADH caused a significant increase in SsLM resistance
because of blocking the reconstituted VDAC channels [70]. Again, the VDAC channels reconstituted
in SsLMs comprise a platform to determine the NADH and NADPH concentration in the
surrounding aqueous milieu. Thus, the two before mentioned examples constitute proof-of-principle
studies for the feasibility to utilize membrane-active peptides and membrane proteins reconstituted
in SsLMs as electrochemical biosensor. The ability to reconstitute integral membrane proteins in
defined structures like pure lipid bilayers [9,132–135], block copolymer bilayers [136–137], and
recently lipid bilayers blended with block copolymers [138,139] on electrode and sensor surfaces is
one of the most important concerns in designing biomimetic sensing devices in future [9,10,140–142].
A vigorous demand on membrane platforms like SsLMs is present because membrane proteins
represent currently more than 50% of all drug targets [143,144]. Drug and membrane protein
Sensors 2020, 20, 1721 16 of 22
screening is; therefore, a vital evolving field in biosensor development because it bears the advantage
of an amplification of the readout signal of membrane functions without need for labelling [105,145].
6. Conclusions and Outlook
Exploiting of bioinspired materials provides great contributions to biosensing applications and
nanotechnology in general [146]. In the recent decades, nanotechnology-based sensing techniques for
either analyte detection or biomimetic approaches have achieved noticeable success. Electrochemical
biosensors are simple, accurate, and inexpensive tools. Moreover, decoration with bioinspired
biomaterials to generate biosensing platforms with good selectivity and sensitivity for, for example,
POC technology can easily done in particular if the biomolecules show the ability to self-assemble.
This review focuses on the exploitation of S-layer proteins in the development of electrochemical
biosensors. The S-layer protein lattice constitutes a very promising intermediate layer because of its
intrinsic chemical and physical characteristics (Table 1). Examples where S-layer protein-coated
sensor surfaces provide a key function range from the detection of metal ions over the immobilization
of enzymes to determine their function themselves or the conversion of a substrate by the enzyme to
measuring specific receptor–ligand interactions of cells (Table 2). Moreover, it is even possible to
discriminate between dead and alive cells. S-layer protein lattices provide also a unique scaffolding
for supported lipid membranes. The retained fluidity of the latter allows the incorporation of
membrane-active peptides and the reconstitution of membrane proteins like (G-protein coupled)
receptors, ion channels, membrane-associated enzymes, and solute carries and transporters, which
constitute nowadays the most important drug targets [83]. All these membrane-active biomolecules
constitute highly sensitive and selective sensing elements for biosensor development and high-
throughput screening devices in the field of diagnostic and drug discovery.
A challenge in the future is the integration of relevant key enabling technologies on an industrial
scale (microfluidics with microelectronics, highly sensitive detection methods and low-cost materials
for easy-to-use tools) [147–150] with biological components like S-layer proteins [7,10]. In general,
biomolecules with self-assembly capability will play a pivotal role. This is not only because of
miniaturization based on this bottom-up approach is rather easy to achieve, but also because of their
molecular dimensions and repetitive features down to the nanoscale. In future, the S-layer lattice may
constitute the key component for the design of smart biosensors relaying on two or more read-out
techniques (e.g., electrochemical, microscopic, spectroscopic, acoustic, etc.), which will enable the
determination of multiple parameters simultaneously.
The manufacturers of smart biosensors, lab-on-a-chip, and microfluidic devices will have to
consider biomaterials as self-assembly components to realize important objectives in biosensor
development like simple but reliable read-out system, volume reduction, high efficiency, low-cost,
and easy-to-use. Finally, the integration in daily life and convenient handling for customers and
patients will be the key parameter for the success of any diagnosis, POC diagnostics or other
biosensor device. In this context, wearable biosensors are gaining vital interest due to their potential
to provide continuous, real-time physiological information via noninvasive measurements of
biochemical markers in biofluids, such as sweat, tears, saliva, and interstitial fluid [151].
Author Contributions: Conceptualization, writing—review and editing, S.D. and B.S. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was funded by the Austrian Science Fund (FWF), project P 29399-B22 (to B.S.).
Acknowledgments: Open Access Funding by the Austrian Science Fund (FWF).
Conflicts of Interest: The authors declare no conflicts of interest.
References
1. Thévenot, D.R.; Toth, K.; Durst, R.A.; Wilson, G.S. Electrochemical biosensors: recommended definitions
and classification. Biosens. Bioelectron. 2001, 16, 121–131.
2. Grieshaber, D.; MacKenzie, R.; Vörös, J.; Reimhult, E. Electrochemical Biosensors—Sensor Principles and
Architectures. Sensors 2008, 8, 1400.
Sensors 2020, 20, 1721 17 of 22
3. D’Orazio, P. Biosensors in clinical chemistry. Clin. Chim. Acta 2003, 334, 41–69.
4. Bahadır, E.B.; Sezgintürk, M.K. Electrochemical biosensors for hormone analyses. Biosens. Bioelectron. 2015,
68, 62–71.
5. Zhang, X.; Guo, Q.; Cui, D. Recent advances in nanotechnology applied to biosensors. Sensors 2009, 9, 1033–1053.
6. Eggins, B. Chemical sensors and biosensors. In Analytical Techniques in the Sciences; John Wiley and Sons:
Hoboken, NJ, USA, 2002.
7. Sleytr, U.B.; Schuster, B.; Egelseer, E.M.; Pum, D. S-layers: Principles and applications. FEMS Microbio. Rev.
2014, 38, 823–864.
8. Pum, D.; Toca-Herrera, J.L.; Sleytr, U.B. S-Layer Protein Self-Assembly. Int. J. Mol. Sci. 2013, 14, 2484-2501.
9. Schuster, B.; Sleytr, U.B. Biomimetic interfaces based on S-layer proteins, lipid membranes and functional
biomolecules. J. R. Soc. Interface 2014, 11, 20140232.
10. Schuster, B. S-layer protein-based biosensors. Biosensors 2018, 8, 40.
11. Sleytr, U.B. Regular arrays of macromolecules on bacterial cell walls: structure, chemistry, assembly, and
function. Int. Rev. Cytol. 1978, 53, 1–62.
12. Sleytr, U.B.; Egelseer, E.M.; Ilk, N.; Messner, P.; Schäffer, C.; Pum, D.; Schuster, B. Nanobiotechnological
applications of S-layers. In Prokaryotic Cell Wall Compounds—Structure and Biochemistry; König, H., Claus,
H., Varma, A., Eds.; Springer: Berlin/Heidelberg, Germany, 2010.
13. Sleytr, U.B.; Schuster, B.; Egelseer, E.M.; Pum, D.; Horejs, C.M.; Tscheliessnig, R.; Ilk, N. Nanobiotechnology
with S-Layer Proteins as Building Blocks In Progress in Molecular Biology and Translational Science, Molecular
Assembly in Natural and Engineered Systems; Howorka, S., Ed.; Elsevier: Amsterdam, The Netherlands, 2011.
14. Sleytr, U.B.; Messner, P. Self-assembly of crystalline bacterial cell surface layers (S-layers). In Electron
Microscopy of Subcellular Dynamics; Plattner, H., Ed.; CRC Press: Boca Raton, FL, USA, 1989.
15. Sleytr, U.B.; Egelseer, E.M.; Ilk, N.; Pum, D.; Schuster, B. S-layers as a basic building block in a molecular
construction kit. FEBS J. 2007, 274, 323–334.
16. Albers, S.V.; Meyer, B.H. The archaeal cell envelope. Nat. Rev. Microbiol. 2011, 9, 414–426.
17. Rodrigues-Oliveira, T.; Belmok, A.; Vasconcellos, D.; Schuster, B.; Kyaw, C.M. Archaeal S-Layers:
Overview and Current State of the Art. Front. Microbiol. 2017, 8, 2597.
18. Rodrigues-Oliveira, T.; Souza, A.A.; Kruger, R.; Schuster, B.; Maria de Freitas, S.; Kyaw, C.M.
Environmental factors influence the Haloferax volcanii S-layer protein structure. PLoS ONE 2019, 14,
e0216863.
19. Mescher, M.F.; Strominger, J.L. Purification and characterization of a prokaryotic glucoprotein from the
cell envelope of Halobacterium salinarium. J. Biol. Chem. 1976, 251, 2005–2014.
20. Sleytr, U.B.; Thorne, K.J. Chemical characterization of the regularly arranged surface layers of Clostridium
thermosaccharolyticum and Clostridium thermohydrosulfuricum. J. Bacteriol. 1976, 126, 377–383.
21. Messner, P.; Steiner, K.; Zarschler, K.; Schäffer, C. S-layer nanoglycobiology of bacteria. Carbohydr. Res.
2008, 343, 1934–1951.
22. Eichler, J.; Maupin-Furlow, J. Post-translation modification in Archaea: lessons from Haloferax volcanii and
other haloarchaea. FEMS Microbiol. Rev. 2013, 37, 583–606
23. Sára, M.; Sleytr, U.B. S-Layer proteins. J. Bacteriol. 2000, 182, 859–868.
24. Sára, M.; Pum, D.; Sleytr, U.B. Permeability and charge-dependent adsorption properties of the S-layer
lattice from Bacillus coagulans E38–66. J. Bacteriol. 1992, 174, 3487–3493.
25. Sleytr, U.B.; Messner, P. Crystalline bacterial cell surface layers (S-layers). In Encyclopedia of Microbiology,
3rd Ed.; Schaechter, M., Ed.; Academic Press: Cambridge, MA, USA, 2009.
26. Pavkov-Keller, T.; Howorka, S.; Keller, W. The structure of bacterial S-layer proteins. In Molecular Assembly
in Natural and Engineered Systems; Howorka, S., Ed.; Elsevier: Amsterdam, The Netherlands, 2011.
27. Sleytr, U.B.; Beveridge, T.J. Bacterial S-layers. Trends Microbiol. 1999, 7, 253–260.
28. Picher, M.M.; Küpcü, S.; Huang, C.J.; Dostalek, J.; Pum, D.; Sleytr, U.B.; Ertl, P. Nanobiotechnology
advanced antifouling surfaces for the continuous electrochemical monitoring of glucose in whole blood
using a lab-on-a-chip. Lab Chip 2013, 13, 1780–1789.
29. Sleytr, U.B.; Györvary, E.; Pum, D. Crystallization of S-layer protein lattices on surfaces and interfaces. Prog.
Organ. Coat. 2003, 47, 279–287.
30. Breitwieser, A.; Siedlaczek, P.; Lichtenegger, H.; Sleytr, U.B.; Pum, D. S-Layer Protein Coated Carbon
Nanotubes. Coatings 2019, 9, 492.
Sensors 2020, 20, 1721 18 of 22
31. Sleytr, U.B.; Egelseer, E.M.; Pum, D.; Schuster, B. S-Layers. In NanoBiotechnology: Concepts, Methods and
Perspectives; Niemeyer, C.M., Mirkin, C.A., Eds.; Wiley-VCH: Weinheim, Germany, 2004.
32. Fagan, R.P.; Fairweather, N.F. Biogenesis and functions of bacterial S-layers. Nat. Rev. Microbiol. 2014, 12,
211–222.
33. Pum, D.; Messner, P.; Sleytr, U.B. Role of the S layer in morphogenesis and cell division of the
archaebacterium Methanocorpusculum sinense. J. Bacteriol. 1991, 173, 6865–6873.
34. Engelhardt, H.; Peters, J. Structural research on surface layers: A focus on stability, surface layer homology
domains, and surface layer–cell wall interactions. J. Struct. Biol. 1998, 124, 276–302.
35. Zink, I.A.; Pfeifer, K.; Wimmer, E.; Sleytr, U.B.; Schuster, B.; Schleper, C. CRISPR-mediated gene silencing
reveals involvement of the archaeal S-layer in cell division and virus infection. Nat. Commun. 2019, 10, 4797.
36. Györvary, E.; O’Riordan, A.; Quinn, A.J.; Redmond, G.; Pum, D.; Sleytr, U.B. Biomimetic nanostructure
fabrication: Nonlithographic lateral patterning and self-assembly of functional bacterial S-layers at silicon
supports. Nano Lett. 2003, 3, 315–319.
37. Pum, D.; Sleytr, U.B. Anisotropic crystal growth of the S-layer of Bacillus sphaericus CCM 2177 at the
air/water interface. Colloids Surface A Physicochem. Eng. Asp. 1995, 102, 99–104.
38. Pum, D.; Sleytr, U.B. Monomolecular reassembly of a crystalline bacterial cell surface layer (S-layer) on
untreated and modified silicon surfaces. Supramol. Sci. 1995, 2, 193–197.
39. Sleytr, U.B.; Sára, M.; Pum, D.; Schuster, B. Crystalline bacterial cell surface layers (S-layers): A versatile
self-assembly system. In Supramolecular Polymers, 2nd ed.; Ciferri, A., Ed.; Taylor and Francis Group: Milton
Park, UK, 2005.
40. Schuster, B.; Sleytr, U.B. Nanotechnology with S-layer proteins. In Methods in Molecular Biology; Gerrard, J.A.,
Domigan, L.J. Eds.; Humana Press, Springer Science+Business Media, , New York, NY, USA, 2020.
41. Györvary, E.; Stein, O.; Pum, D.; Sleytr, U.B. Self-assembly and recrystallization of bacterial S-layer proteins
at silicon supports imaged in real time by atomic force microscopy. J. Microsc. 2003, 212, 300–306.
42. Schuster, B.; Sleytr, U.B. Biomimetic S-layer supported lipid membranes. Curr. Nanosci. 2006, 2, 143–152.
43. Schuster, B. Biomimetic design of nanopatterned membranes. Nanobiotechnology 2005, 1, 153–164.
44. Schuster, B.; Pum, D.; Sleytr, U.B. S-layer stabilized lipid membranes. Biointerphases 2008, 3, FA3–FA11.
45. Sára, M.; Manigley, C.; Wolf, G.; Sleytr, U.B. Isoporous ultrafiltration membranes from bacterial cell
envelope layers. J. Membr. Sci. 1988, 36, 179–186.
46. Sára, M.; Sleytr, U.B. Production and characteristics of ultrafiltration membranes with uniform pores from
two-dimensional arrays of proteins. J. Membr. Sci. 1987, 33, 27–49.
47. Pum, D.; Sára, M.; Sleytr, U.B. Two-dimensional (glyco) protein crystals as patterning elements and
immobilisation matrices for the development of biosensors. In Immobilised Macromolecules: Application Potentials;
Sleytr, U.B., Messner, P., Pum, D., Sára, M., Eds.; Springer-Verlag: Berlin/Heidelberg, Germany, 1993.
48. Sleytr, U.B.; Bayley, H.; Sára, M.; Breitwieser, A.; Küpcü, S.; Mader, C.; Weigert, S.; Unger, F.M.; Messner,
P.; Jahn-Schmid, B.; et al. Applications of S-layers. FEMS Microbiol. Rev. 1997, 20, 151–175.
49. Sleytr, U.B.; Sara, M. Ultrafiltration membranes with uniform pores from crystalline bacterial cell envelope
layers. Appl. Microbiol. Biot. 1986, 25, 83–90.
50. Sára, M.; Küpcü, S.; Weiner, C.; Weigert, S.; Sleytr, U.B. Crystalline protein layers as isoporous molecular
sieves and immobilisation and affinity matrices. In Immobilised Macromolecules: Application Potentials; Sleytr,
U.B., Messner, P., Pum, D., Sára, M., Eds.; Springer-Verlag: Berlin/Heidelberg, Germany, 1993.
51. Handrea, M.; Sahre, M.; Neubauer, A.; Sleytr, U.B.; Kautek, W. Electrochemistry of nano-scale bacterial
surface protein layers on gold. Bioelectrochemistry 2003, 61, 1–8.
52. Sotiropoulou, S.; Mark, S.S.; Angert, E.R.; Batt, C.A. Nanoporous S-layer protein lattices. A biological ion
gate with calcium selectivity. J. Phys. Chem. C 2007, 111, 13232–13237.
53. Zafiu, C.; Trettenhahn, G.; Pum, D.; Sleytr, U.B.; Kautek, W. Electrochemical control of adsorption
dynamics of surface layer proteins on gold. Phys. Chem. Chem. Phys. 2011, 13, 3478–3483.
54. Zafiu, C.; Trettenhahn, G.; Pum, D.; Sleytr, U.B.; Kautek, W. Structural control of surface layer proteins at
electrified interfaces investigated by in situ Fourier transform infrared spectroscopy. Phys. Chem. Chem.
Phys. 2011, 13, 13232–13237.
55. Moudgil, A.; Kalyani, N.; Sinsinbar, G.; Das, S.; Mishra, P. S-Layer Protein for Resistive Switching and
Flexible Nonvolatile Memory Device. ACS Appl. Mater. Inter. 2018, 10, 4866–4873.
56. Bard, A.J.; Faulkner, L.R. Electrochemical Methods, Fundamentals and Applications, 1st ed.; Wiley: Hoboken,
NJ, USA, 2001.
Sensors 2020, 20, 1721 19 of 22
57. Juan Colás, J. Fabrication and Experimental Techniques. In Dual-Mode Electro-photonic Silicon Biosensors;
Springer: Berlin/Heidelberg, Germany, 2017.
58. Monosik, R.; Streďanský, M.; Šturdík, E. Biosensors—Classification, characterization and new trends. Acta
Chim. Slovaca 2012, 1, 109–120.
59. Ives, D.J.G.; Janz, G.J. Reference Electrodes, Theory and Practice, 1st ed.; Academic Press: Cambridge, MA,
USA, 1961.
60. Narayan, R.J. Medical Biosensors For point of Care (POC) Applications; Woodhead Publishing: NC, USA, 2017.
61. Bettazzi, F.; Marrazza, G.; Minunni, M.; Palchetti, I.; Scarano, S. Biosensors and Related Bioanalytical Tools.
Compr. Anal. Chem. 2017, 77, 1–33
62. Neubauer, A.; Pum, D.; Sleytr, U.B. An amperometric glucose sensor based on isoporous crystalline protein
membranes as immobilization matrix. Anal. Lett. 1993, 26, 1347–1360.
63. Pum, D.; Neubauer, A.; Sleytr, U.B.; Pentzien, S.; Reetz, S.; Kautek, W. Physico-chemical properties of
crystalline nanoscale enzyme-protein-metal layer composites in biosensors. Ber. Bunsen. Phys. Chem. 1997,
101, 1686–1689.
64. Buck, R.B.; Lindner, E. Recommendations for nomenclature of ionselective electrodes. Pure Appl. Chem.
1994, 66, 2527.
65. Cullen, D.C.; Sethi, R.S.; Lowe, C.R. Multi-analyte miniature conductance biosensor. Anal. Chim. Acta 1990,
231, 33.
66. Conneely, G.; Aherne, M.; Lu, H.; Guilbault, G.G. Electrochemical immunosensors for the detection of 19-
nortestosterone and methyltestosterone in bovine urine. Sens. Actuators B 2007, 121, 103–112.
67. Luppa, P.B.; Sokoll, L.J.; Chan, D.W. Immunosensors—Principles and applications to clinical chemistry.
Clin. Chim. Acta 2001, 314, 1–26.
68. Willner, I., Katz, E. Bioelectronics: from theory to applications.: Wiley-VCH Verlag GmbH & Co. KGaA;
Weinheim, Germany, 2005.
69. Damiati, S.; Schrems, A.; Sinner, E.K.; Sleytr, U.B.; Schuster, B. Probing peptide and protein insertion in a
biomimetic S-layer supported lipid membrane platform. Int. J. Mol. Sci. 2015, 16, 2824–2838.
70. Damiati, S.; Zayni, S.; Schrems, A.; Kiene, E.; Sleytr, U.B.; Chopineau, J.; Schuster, B.; Sinner, E.K. Inspired
and stabilized by nature: ribosomal synthesis of the human voltage gated ion channel (VDAC) into 2D-
protein-tethered lipid interfaces. Biomater. Sci. 2015, 3, 1406–1413
71. Gufler, P.C.; Pum, D.; Sleytr, U.B.; Schuster, B. Highly robust lipid membranes on crystalline S-layer supports
investigated by electrochemical impedance spectroscopy. Biochim. Biophys. Acta 2004, 1661, 154–165
72. Damiati, S.; Küpcü, S.; Peacock, M.; Eilenberger, C.; Zamzami, M.; Qadri, I.; Choudhry, H.; Sleytr, U.B.;
Schuster, B. Acoustic and hybrid 3D-printed electrochemical biosensors for the real-time immunodetection
of liver cancer cells (HepG2). Biosens. Bioelectron. 2017, 94, 500–506
73. Damiati, S.; Peacock, M.; Sleytr, U.; Schuster, B. Bioinspired Diagnostic Sensor Based on Functional
Nanostructures of S-Proteins to Target the Folate Receptors in Breast Cancer Cells. Sens. Actuators B 2018,
267, 224–230.
74. Sotiropoulou, S.; Gavalas, V.; Vamvakaki, V.; Chaniotakis, N.A. Novel carbon materials in biosensor
systems. Biosens. Bioelectron. 2003, 18, 211–215
75. Kuznetsov, B.A.; Shumakovich, G.P.; Koroleva, O.V.; Yaropolov, A.I. On applicability of laccase as label in
the mediated and mediatorless electroimmunoassay: effect of distance on the direct electron transfer
between laccase and electrode. Biosens. Bioelectron. 2001, 16, 73–84.
76. Reimhult, E.; Höök, F. Design of Surface Modifications for Nanoscale Sensor Applications. Sensors 2015, 15,
1635.
77. Guidelli, R.; Becucci, L. Electrochemistry of Biomimetic Membranes. In Applications of Electrochemistry and
Nanotechnology in Biology and Medicine II. Modern Aspects of Electrochemistry; Eliaz, N., Ed.; Springer:
Berlin/Heidelberg, Germany, 2012.
78. Damiati, S. Can We Rebuild the Cell Membrane? In Biological, Physical and Technical Basics of Cell
Engineering; Artmann, G., Artmann, A., Zhubanova, A., Digel, I., Eds.; Springer: Berlin/Heidelberg,
Germany, 2018.
79. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Molecular Biology of the Cell, 4th ed.;
Garland Science: New York, NY, USA, 2002.
80. Cremer, P.S.; Boxer, S.G. Formation and spreading of lipid bilayers on planar glass supports. J. Phys. Chem.
B 1999, 103, 2554–2559.
Sensors 2020, 20, 1721 20 of 22
81. Hardy, G.J.; Nayak, R.; Zauscher, S. Model cell membranes: Techniques to form complex biomimetic
supported lipid bilayers via vesicle fusion. Curr. Opin. Colloid Interface Sci. 2013, 18, 448–458.
82. Cho, N.J.; Frank, C.W.; Kasemo, B.; Hook, F. Quartz crystal microbalance with dissipation monitoring of
supported lipid bilayers on various substrates. Nat. Protoc. 2010, 5, 1096–1106.
83. Tiefenauer, L.; Demarche, S. Challenges in the development of functional assays of membrane proteins.
Materials 2012, 5, 2205.
84. Jackson, A.; Boutell, J.; Cooley, N.; He, M. Cell-free protein synthesis for proteomics. Brief Funct. Genom.
Proteom. 2004, 2, 308–319.
85. Endo, Y.; Sawasaki, T. Cell-free expression systems for eukaryotic protein production. Curr. Opin.
Biotechnol. 2006, 17, 373–380.
86. Damiati, S.; Mhanna, R.; Kodzius, R.; Ehmoser, E.K. Cell-Free Approaches in Synthetic Biology Utilizing
Microfluidics. Genes 2018, 9, 144.
87. Kalia, J.; Raines, R.T. Advances in Bioconjugation. Curr. Org. Chem. 2010, 14, 138–147.
88. Eggins, B.R. Biosensors: An Introduction; Springer-Verlag: Berlin/Heidelberg, Germany, 2013.
89. Fang, Y.; Ramasamy, R. Current and prospective methods for plant disease detection. Biosensors 2015, 5,
537–561.
90. Perez, J.A.C.; Sosa-Hernandez, J.E.; Hussain, S.M.; Bilal, M.; Parra-Saldivar, R.; Iqbal, H.M.N. Bioinspired
biomaterials and enzyme-based biosensors for point-of-care applications with reference to cancer and bio-
imaging. Biocat. Agricult. Biotech. 2019, 17, 168–176.
91. Conroy, D.J.R.; Millner, P.A.; Stewart, D.I.; Pollmann, K. Biosensing for the environment and defence:
Aqueous uranyl detection using bacterial surface layer proteins. Sensors 2010, 10, 4739.
92. Neubauer, A.; Hödl, C.; Pum, D.; Sleytr, U.B. A multistep enzyme sensor for sucrose based on S-layer
microparticles as immobilization matrix. Anal. Lett. 1994, 27, 849–865.
93. Guimarães, J.A.; Ferraz, H.C.; Alves, T.L.M. Langmuir-Blodgett films of cholesterol oxidase and S-layer
proteins onto screen-printed electrodes. Appl. Surf. Sci. 2014, 298, 68–74.
94. Park, T.J.; Lee, S.J.; Park, J.P.; Yang, M.H.; Choi, J.H.; Lee, S.Y. Characterization of a bacterial self-assembly surface
layer protein and its application as an electrical nanobiosensor. J. Nanosci. Nanotechno. 2011, 11, 402–407
95. Sára, M.; Küpcü, S.; Weiner, C.; Weigert, S.; Sleytr, U.B. S-layers as immobilization and affinity matrices. In
Advances in Bacterial Paracrystalline Surface Layers, Beveridge, T.J., Koval, S.F., Eds.; Plenum Press: New
York, NY, USA, 1993.
96. Neubauer, A.; Pum, D.; Sleytr, U.B. Fibre-optic glucose biosensor using enzyme membranes with 2-D
crystalline structure. Biosens. Bioelectron. 1996, 11, 317–325.
97. Neubauer, A.; Pentzien, S.; Reetz, S.; Kautek, W.; Pum, D.; Sleytr, U.B. Pulsed-laser metal contacting of
biosensors on the basis of crystalline enzyme-protein layer composites. Sens. Actuators B Chem. 1997, 40,
231–236
98. Schuster, B.; Sleytr, U.B. Relevance of glycosylation of S-layer proteins for cell surface properties. Acta
Biomater. 2015, 19, 149–157.
99. Wu, S.; Hao, J.; Yang, S.; Sun, Y.; Wang, Y.; Zhang, W.; Mao, H.; Song, X.-M. Layer-by-layer self-assembly
film of PEI-reduced graphene oxide composites and cholesterol oxidase for ultrasensitive cholesterol
biosensing. Sens. Actuators B Chem. 2019, 298, 126856.
100. Ferraz, H.C.; Guimarães, J.A.; Alves, T.L.M.; Constantino, C.J.L. Monomolecular films of cholesterol
oxidase and S-Layer proteins. Appl. Surf. Sci. 2011, 257, 6535–6539.
101. El-Said, W.A.; Yea, C.H.; Kim, H.; Oh, B.K.; Choi, J.W. Cell-based chip for the detection of anticancer effect
on HeLa cells using cyclic voltammetry. Biosens. Bioelectron. 2009, 24, 1259–1265.
102. Yeon, J.H.; Park, J.-K. Cytotoxicity test based on electrochemical impedance measurement of HepG2
cultured in microfabricated cell chip. Anal. Biochem. 2005, 341, 308–315
103. Zhu, X.; Feng, C.; Ye, Z.; Chen, Y.; Li, G. Fabrication of magneto-controlled moveable architecture to
develop reusable electrochemical biosensors. Sci. Rep. 2015, 4, 4169.
104. Damiati, S.; Peacock, M.; Leonhardt, S.; Damiati, L.; Baghdadi, M.A.; Becker, H.; Kodzius, R.; Schuster, B.
Embedded Disposable Functionalized Electrochemical Biosensor with a 3D-Printed Flow-Cell for
Detection of Hepatic Oval Cells. Genes 2018, 9, 89.
105. Sun, D.; Lu, J.; Chen, Z.; Yu, Y.; Mo, M. A repeatable assembling and disassembling electrochemical
aptamer cytosensor for ultrasensitive and highly selective detection of human liver cancer cells. Anal. Chim.
Acta 2015, 885, 166–173.
Sensors 2020, 20, 1721 21 of 22
106. Han, L.; Liu, P.; Petrenko, V.A.; Liu, A. A Label-Free Electrochemical Impedance Cytosensor Based on
Specific Peptide-Fused Phage Selected from Landscape Phage Library. Sci. Rep. 2016, 6, 22199.
107. Baniukevic, J.; Kirlyte, J.; Ramanavicius, A.; Ramanaviciene, A. Application of oriented and random
antibody immobilization methods in immunosensor design. Sens. Actuators B Chem. 2013, 189, 217–223.
108. Deng, X.; Chen, M.; Fu, Q.; Smeets, N.; Fu, X.; Zhang, Z.; Filipe, C.; Hoare, T. A Highly Sensitive
Immunosorbent Assay Based on Biotinylated Graphene Oxide and the Quartz Crystal Microbalance. ACS
Appl. Mater. Interfaces 2016, 8, 1893–1902.
109. Bae, Y.; Oh, M.; Lee, W.; Lee, W.; Choi, J. Study on orientation of immunoglobulin G on protein G layer.
Biosens. Bioelectron. 2005, 21, 103–110.
110. Kausaite-Minkstimiene, A.; Ramanaviciene, A.; Kirlyte, J.; Ramanavicius, A. Comparative Study of
Random and Oriented Antibody Immobilization Techniques on the Binding Capacity of Immunosensor.
Anal. Chem. 2010, 82, 6401–6408.
111. Xiao, Y.; Lin, M.; Jiang, X.; Ye, J.; Guo, T.; Shi, Y.; Bian, X. The Recent Advances on Liver Cancer Stem Cells:
Biomarkers, Separation, and Therapy. Anal. Cell Pathol. 2017, 2017, 5108653.
112. Eliasson, M.; Olsson, A.; Palmcrantz, E.; Wiberg, K.; Inganas, M.; Guss, B.; Lindberg, M.; Uhlen, M.
Chimeric IgG-binding receptors engineered from staphylococcal protein A and streptococcal protein. J.
Biol. Chem. 1988, 263, 4323–4327.
113. Völlenkle, C.; Weigert, S.; Ilk, N.; Egelseer, E.; Weber, V.; Loth, F.; Falkenhagen, D.; Sleytr, U.B.; Sara, M.
Construction of a Functional S-Layer Fusion Protein Comprising an Immunoglobulin G-Binding Domain
for Development of Specific Adsorbents for Extracorporeal Blood Purification. Appl. Environ. Microbiol.
2004, 70, 1514–1521.
114. Allard, W.J.; Matera, J.; Miller, M.C.; Repollet, M.; Connelly, M.C.; Rao, C.; Tibbe, A.G.; Uhr, J.W.;
Terstappen, L.W. Tumor cells circulate in the peripheral blood of all major carcinomas but not in healthy
subjects or patients with nonmalignant diseases. Clin. Cancer Res. 2004, 10, 6897–6904.
115. De Rosa, M., Archaeal lipids: Structural features and supramolecular organization. Thin Solid Films 1996,
284–285, 13–17.
116. Hanford, M.J.; Peeples, T.L. Archaeal tetraether lipids: Unique structures and applications. Appl. Biochem.
Biotechnol. Enzym. Eng. Biotechnol. 2002, 97, 45–62.
117. Eme, L.; Spang, A.; Lombard, J.; Stairs, C.W.; Ettema, T.J.G. Archaea and the origin of eukaryotes. Nat. Rev.
Microbiol. 2017, 15, 711–723.
118. Engelhardt, H. Are S-layers exoskeletons? The basic function of protein surface layers revisited. J. Struct.
Biol. 2007, 160, 115–124.
119. Engelhardt, H. Mechanism of osmoprotection by archaeal S-layers: A theoretical study. J. Struct. Biol. 2007,
160, 190–199.
120. Schuster, B.; Sleytr, U.B. Composite S-layer lipid structures. J. Struct. Biol. 2009, 168, 207–216.
121. Schuster, B.; Pum, D.; Sleytr, U.B. Voltage clamp studies on S-layer-supported tetraether lipid membranes.
Biochim. Biophys. Acta 1998, 1369, 51–60.
122. Su, Z.; Abbasi, F.; Leitch, J.J.; Faragher, R.J.; Schwan, A.L.; Lipkowski, J. Mechanisms of alamethicin ion
channel inhibition by amiloride in zwitterionic tethered lipid bilayers. J. Electroanal. Chem. 2019, 848, 113281.
123. Schrems, A.; Larisch, V.D.; Stanetty, C.; Dutter, K.; Damiati, S.; Sleytr, U.B.; Schuster, B. Liposome fusion
on proteinaceous S-layer lattices triggered via β-diketone ligand-europium(iii) complex formation. Soft
Matter 2011, 7, 5514–5518.
124. Colombini, M. VDAC: The channel at the interface between mitochondria and the cytosol. Mol. Cell.
Biochem. 2004, 256–257, 107–115.
125. Colombini, M. VDAC structure, selectivity, and dynamics. Biochim. Biophys. Acta 2012, 1818, 1457–1465.
126. Colombini, M. The VDAC channel: Molecular basis for selectivity. Biochim. Biophys. Acta 2016, 1863, 2498–2502.
127. Colombini, M.; Blachly-Dyson, E.; Forte, M. VDAC, a channel in the outer mitochondrial membrane. Ion
Channels 1996, 4, 169–202.
128. Komarov, A.G.; Deng, D.; Craigen, W.J.; Colombini, M. New insights into the mechanism of permeation
through large channels. Biophys. J. 2005, 89, 3950–3959.
129. Lee, A.C.; Xu, X.; Colombini, M. The role of pyridine dinucleotides in regulating the permeability of the
mitochondrial outer membrane. J. Biol. Chem. 1996, 271, 26724–26731.
130. Rostovtseva, T.K.; Komarov, A.; Bezrukov, S.M.; Colombini, M. Dynamics of nucleotides in VDAC
channels: Structure-specific noise generation. Biophys. J. 2002, 82, 193–205.
Sensors 2020, 20, 1721 22 of 22
131. Zizi, M.; Forte, M.; Blachly-Dyson, E.; Colombini, M. NADH regulates the gating of VDAC, the
mitochondrial outer membrane channel. J. Biol. Chem.1994, 269, 1614-1616.
132. Knoll, W.; Naumann, R.; Friedrich, M.; Robertson, J.W.F.; Lösche, M.; Heinrich, F.; McGillivray, D.J.;
Schuster, B.; Gufler, P.C.; Pum, D.; et al Solid supported lipid membranes: New concepts for the biomimetic
functionalization of solid surfaces. Biointerphases 2008, 3, FA125–FA135.
133. Knoll, W.; Köper, I.; Naumann, R.; Sinner, E.K. Tethered bimolecular lipid membranes-A novel model
membrane platform. Electrochim. Acta 2008, 53, 6680–6689.
134. Sinner, E.K.; Knoll, W. Functional tethered membranes. Curr. Opin. Chem. Biol. 2001, 5, 705–711.
135. Sinner, E.K.; Ritz, S.; Naumann, R.; Schiller, S.; Knoll, W. Self-assembled tethered bimolecular lipid
membranes. Adv. Clin. Chem. 2009, 49,159–179.
136. Zaba, C.; Ritz, S.; Tan, C.-W.D.; Zayni, S.; Müller, M.; Reuning, U.; Sinner, E.K. Functional Cell Adhesion
Receptors (Integrins) in Polymeric Architectures. ChemBioChem 2015, 16, 1740–1743.
137. Zapf, T.; Tan, C.-W.D.; Reinelt, T.; Huber, C.; Shaohua, D.; Geifman-Shochat, S.; Paulsen, H.; Sinner, E.K.
Synthesis and Functional Reconstitution of Light-Harvesting Complex II into Polymeric Membrane
Architectures. Angew. Chem. Int. Edit. 2015, 54, 14664–14668.
138. Czernohlavek, C.; Schuster, B. Formation and characteristics of mixed lipid/polymer membranes on a
crystalline surface-layer protein lattice. Biointerphases 2020, 15, 011002.
139. Czernohlavek, C.; Schuster, B. Formation of planar hybrid lipid/polymer membranes anchored to an
S-layer protein lattice by vesicle binding and rupture. Soft Mater. 2020, doi: 10.1080/1539445X.2019.1708753.
140. Schuster, B.; Sleytr, U.B. Tailor-made crystalline structures of truncated S-layer proteins on
heteropolysaccharides. Soft Matter 2009, 5, 334–341.
141. Jackman, J.A.; Knoll, W.; Cho, N.-J. Biotechnology applications of tethered lipid bilayer membranes.
Materials 2012, 5, 2637.
142. Trojanowicz, M. Miniaturized biochemical sensing devices based on planar lipid membranes. Fresenius J.
Anal. Chem. 2001, 372, 246–260.
143. Howard, A.D.; McAllister, G.; Feighner, S.D.; Liu, Q.; Nargund, R.P.; Van der Ploeg, L.H.T.; Patchett, A.A.
Orphan G-protein-coupled receptors and natural ligand discovery. Trends Pharmacol. Sci. 2001, 22, 132–140.
144. Howard, A.D.; Wang, R.; Pong, S.S.; Mellin, T.N.; Strack, A.; Guan, X.M.; Zeng, Z.; Williams Jr, D.L.;
Feighner, S.D.; Nunes, C.N.; et al. Identification of receptors for neuromedin U and its role in feeding.
Nature 2000, 406, 70–74.
145. Walsh, K.B. Targeting cardiac potassium channels for state-of-the-art drug discovery. Expert Opin. Drug
Dis. 2015, 10, 157–169.
146. Zhao, Q.; Wang, Y.; Cui, H.; Du, X. Bio-inspired sensing and actuating materials. J. Mater. Chem. C 2019, 7,
6493–6511.
147. Liao, Z.; Wang, J.; Zhang, P.; Zhang, Y.; Miao, Y.; Gao, S.; Deng, Y.; Geng, L. Recent advances in microfluidic
chip integrated electronic biosensors for multiplexed detection. Biosens. Bioelectron. 2018, 121, 272–280.
148. Naranjo-Hernández, D.; Reina-Tosina, J.; Min, M. Fundamentals, recent advances, and future challenges in
bioimpedance devices for healthcare applications. J. Sensors 2019, 9210258. doi: 10.1155/2019/9210258
149. Primiceri, E.; Chiriacò, M.S.; Notarangelo, F.M.; Crocamo, A.; Ardissino, D.; Cereda, M.; Bramanti, A.P.;
Bianchessi, M.A.; Giannelli, G.; Maruccio, G. Key enabling technologies for point-of-care diagnostics.
Sensors 2018, 18, 3607.
150. Jurado-Sánchez, B. Microscale and nanoscale biosensors. Biosensors 2018, 8, 66.
151. Kim, J.; Campbell, A.S.; de Ávila, B.E.-F.; Wang, J. Wearable biosensors for healthcare monitoring. Nat.
Biotechnol. 2019, 37, 389–406.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... A typical electrochemical biosensor features three key electrodes: (i) Working electrode (WE) that is the heart of the sensor, where the biorecognition reaction occurs and the biomolecule-analyte interaction is translated into an electrical signal; [43] (ii) Reference electrode (RE), which is immersed in the solution but away from the reaction site, providing a stable point of reference for potential measurements; and (iii) Auxiliary electrode (AE) that is responsible for supplying current to the working electrode to facilitate efficient signal generation. Both the reference and auxiliary electrodes require high conductivity and stability [128]. The remarkable appeal of electrochemical biosensors for point-of-care applications lies in their unique characteristics, which were highlighted in the previous paragraph. ...
... (a) Different types of electrochemical nanobiosensors and (b) the different signals obtained(Banerjee et al., 2021;Curulli, 2021;Damiati & Schuster, 2020;Dincer et al., 2019;Martín-Yerga, 2019;Shen et al., 2014). ...
Chapter
Full-text available
In the quest for sustainable solutions, a groundbreaking innovation emerges: Nanobiohybrids for Advanced Wastewater Treatment and Energy Recovery. This compelling title delves into the forefront of integrated environmental technology, exploring the synergistic potential of nanotechnology and biotechnology in revolutionizing the way we manage wastewater and harness energy. Inside these pages, you'll: Find insightful explorations of process fundamentals of nanobiohybrids for water purification, microbial synthesis of nanocatalysts and energy storage devices. Grasp the mechanics of nanoparticle biosynthesis, biomolecule integration, and the dynamic interplay between nanotechnology and biology.Embark on fundamental components investigating environmental remediation for both wastewater treatment and energy recovery. Focusing on both components of nanobiohybrids: on the one hand organic, inorganic and magnetic nanoparticles, and on the other hand electroactive bacteria.Understand the implications of nanobiohybrids for environmental remediation, water and wastewater treatment to conserve resources and protect the environment. Explore how these solutions contribute to pollution control, resource recovery, sensing and photoelectrocatalysis. This book Nanobiohybrids for Advanced Wastewater Treatment and Energy Recovery is an indispensable resource for researchers, students, policymakers, and anyone intrigued by the intersection of nanoscience, nanotechnology and sustainability. As we stand at the brink of a new era in environmental engineering, this book empowers you to be at the forefront of change. Join the movement towards cleaner waters, abundant energy and a brighter future with nanobiohybrids. ISBN 9781789063585 (paperback) ISBN 9781789063592 (eBook) ISBN 9781789063608 (ePub)
... Any current that flows into the skin ISF via the working electrode leaves the solution via the counter electrode, completing the circuit. The reference electrode then measures the electrical potential of the working electrode 94 . For this design to function reliably, the system needs to consider the differences in electrical resistance between the dermis and epidermis. ...
Article
Microneedles (MNs) are microscopic needles that are applied to the skin in a minimally invasive way to facilitate transdermal drug delivery and/or uptake of interstitial fluid from the skin, which contains a variety of metabolites that can serve as biomarkers. The collection of interstitial fluid can be followed by post-sampling analysis or in situ real-time biosensing for disease diagnosis and drug monitoring. The painless and easy administration of MNs is appealing to patients, especially for long-term monitoring. In this Review, we discuss the use of MNs for biosensing purposes. We highlight the different types of MNs and sensing technologies used to develop MN-based biosensors. In addition, we discuss the potential to integrate MNs with wearable devices for real- time monitoring to improve point-of-care testing. Finally, we review the translational hurdles to be considered in bringing this technology from benchtop to bedside.
... They are self-contained integrated devices that provide qualitative and quantitative data about analyte by using a biological recognition element that is coupled to a transduction element. The purpose of these analytical devices is to rapidly provide accurate and reliable information about the analyte of interest in real time [3][4][5][6]. ...
Article
The main factor determining the functioning of the track biosensor is the influence of the composition of an ionic liquid flowing through the track on the ion current density at a certain external voltage. When any contamination enters the ionic liquid flowing through the track, the current density changes. This change determines the sensitivity of the biosensor. The amplifying electronic circuit of the device makes it possible to vary the sensitivity parameter over a wide range. In this work, it was found that the sensitivity of the biosensor depends on the mode in which it is operated. This means that the mode of operation must be selected depending on what sensitivity is required in each particular case.
Article
Full-text available
Cytokines are signaling proteins present in human fluids. Their detection is crucial in understanding immune system activity and supporting the diagnosis and monitoring of diseases such as sepsis, diabetes, and cancers. Nevertheless, investigating emporal profiles of these crucial biomarkers still relies on blood sampling (potentially uncomfortable for patients and only providing a snapshot) and laboratory analysis. Wearable electrochemical sensors represent next-generation diagnostic and health monitoring methodologies, effectively investigating biomarkers in a continuous manner in alternative body fluids (e.g., sweat), thereby holding potential for early disease detection and enabling remote and enhanced personalized health monitoring. This work comprehensively reviews the significant potential of continuous electrochemical detection of cytokines in sweat. First, an overview of the role of cytokines from a medical perspective is given. Then, the concept of electrochemical sensors is introduced and the current state of the art of electrochemical sensing of cytokines in sweat toward wearable applications is critically assessed.
Article
Full-text available
Vitamins are crucial micronutrients found in limited quantities in food, living organisms, and soil. Since most vitamins are not produced within the human body, a lack of these essential nutrients can result in various physiological disorders. Analyzing vitamins typically involves costly, time-consuming methods, requiring skilled personnel, automated equipment, and dedicated laboratory setups. The pressing need is for the development of efficient, portable, and user-friendly detection techniques that are cost-effective, addressing the challenges associated with traditional analytical approaches. In recent years, electrochemical sensors and electrochemical microfluidic devices have garnered prominence owing to their remarkable sensitivity, quick analysis, cost-effectiveness, and facile fabrication procedures. Electrochemical sensing and microfluidics are two distinct fields that are often integrated to create powerful and versatile sensing devices. The connection between them leverages the advantages of both fields to create highly efficient, miniaturized, and portable analytical systems. This interdisciplinary approach has led to the development of innovative devices with broad applications in various scientific, medical, and environmental domains. This review begins by outlining the importance of vitamins in human nutrition and health and emphasizing the need for precise and reliable sensing techniques. Owing to the limited literature available on electrochemical detection of vitamin B complexes, this review offers an in-depth analysis of modern electrochemical sensing of water-soluble vitamins, focusing on B1, B2, B6, B9, and B12. The challenges faced by researchers are addressed, including selectivity, sensitivity, interference, matrix effects, and calibration, while also exploring promising prospects such as nanomaterial integration, miniaturization, microfluidics-based IoTs, and innovative sensor designs.
Article
Full-text available
Sweat is easily accessible from the human skin’s surface. It is secreted by the eccrine glands and contains a wealth of physiological information, including metabolites and electrolytes like glucose and Na ions. Sweat is a particularly useful biofluid because of its easy and non-invasive access, unlike other biofluids, like blood. On the other hand, nanomaterials have started to show promise operation as a competitive substitute for biosensors and molecular sensors throughout the last 10 years. Among the most synthetic nanomaterials that are studied, applied, and discussed, carbon nanomaterials are special. They are desirable candidates for sensor applications because of their many intrinsic electrical, magnetic, and optical characteristics; their chemical diversity and simplicity of manipulation; their biocompatibility; and their effectiveness as a chemically resistant platform. Carbon nanofibers (CNFs), carbon dots (CDs), carbon nanotubes (CNTs), and graphene have been intensively investigated as molecular sensors or as components that can be integrated into devices. In this review, we summarize recent advances in the use of carbon nanomaterials as sweat sensors and consider how they can be utilized to detect a diverse range of analytes in sweat, such as glucose, ions, lactate, cortisol, uric acid, and pH. Graphical Abstract
Article
Full-text available
The implementation of self-assembled biomolecules on solid materials, in particular, sensor and electrode surfaces, gains increasing importance for the design of stable functional platforms, bioinspired materials, and biosensors. The present study reports on the formation of a planar hybrid lipid/polymer membrane on a crystalline surface layer protein (SLP) lattice. The latter acts as a connecting layer linking the biomolecules to the inorganic base plate. In this approach, chemically bound lipids provided hydrophobic anchoring moieties for the hybrid lipid/polymer membrane on the recrystallized SLP lattice. The rapid solvent exchange technique was the method of choice to generate the planar hybrid lipid/polymer membrane on the SLP lattice. The formation process and completeness of the latter were investigated by quartz crystal microbalance with dissipation monitoring and by an enzymatic assay using the protease subtilisin A, respectively. The present data provide evidence for the formation of a hybrid lipid/polymer membrane on an S-layer lattice with a diblock copolymer content of 30%. The hybrid lipid/polymer showed a higher stiffness compared to the pure lipid bilayer. Most interestingly, both the pure and hybrid membrane prevented the proteolytic degradation of the underlying S-layer protein by the action of subtilisin A. Hence, these results provide evidence for the formation of defect-free membranes anchored to the S-layer lattice.
Article
Full-text available
Exploitation of biomolecular and biomimetic components on solid surfaces gain increasing importance for the design of stable functional platforms. The present study performed by quartz crystal microbalance with dissipation monitoring (QCM-D) reports on the formation of planar hybrid lipid/polymer membranes anchored to a crystalline surface (S-) layer protein lattice. In this approach, hybrid lipid/polymer vesicles were chemically bound to the S-layer protein lattice. Subsequently, to form a hybrid planar layer rupture and fusion was triggered either by (1) β–diketone – europium ion complex formation or (2) successive application of calcium ions, lowering the pH from 9 to 4, and the detergent CHAPS. As determined by QCM-D, method 1 resulted for a polymer content of 5% in a planar membrane with some imbedded intact vesicles, whereas method 2 succeeded in planar hybrid membranes with a polymer content of even up to 70%. These results provide evidence for the effective formation of planar lipid/polymer membranes varying in their composition on an S-layer protein lattice.
Article
Full-text available
The S-layer is a proteinaceous surface lattice found in the cell envelope of bacteria and archaea. In most archaea, a glycosylated S-layer constitutes the sole cell wall and there is evidence that it contributes to cell shape maintenance and stress resilience. Here we use a gene-knockdown technology based on an endogenous CRISPR type III complex to gradually silence slaB, which encodes the S-layer membrane anchor in the hyperthermophilic archaeon Sulfolobus solfataricus. Silenced cells exhibit a reduced or peeled-off S-layer lattice, cell shape alterations and decreased surface glycosylation. These cells barely propagate but increase in diameter and DNA content, indicating impaired cell division; their phenotypes can be rescued through genetic complementation. Furthermore, S-layer depleted cells are less susceptible to infection with the virus SSV1. Our study highlights the usefulness of the CRISPR type III system for gene silencing in archaea, and supports that an intact S-layer is important for cell division and virus susceptibility. The S-layer is a proteinaceous envelope often found in bacterial and archaeal cells. Here, the authors use CRISPR-based technology to silence slaB, encoding the S-layer membrane anchor, to show that an intact S-layer is important for cell division and virus susceptibility in the archaeon Sulfolobus solfataricus.
Article
Full-text available
Carbon nanotubes (CNTs) have already been considered for medical applications due to their small diameter and ability to penetrate cells and tissues. However, since CNTs are chemically inert and non-dispersible in water, they have to be chemically functionalized or coated with biomolecules to carry payloads or interact with the environment. Proteins, although often only randomly bound to the CNT surface, are preferred because they provide a better biocompatibility and present functional groups for binding additional molecules. A new approach to functionalize CNTs with a closed and precisely ordered protein layer is offered by bacterial surface layer (S-layer) proteins, which have already attracted much attention in the functionalization of surfaces. We could demonstrate that bacterial S-layer proteins (SbpA of Lysinibacillus sphaericus CCM 2177 and the recombinant fusion protein rSbpA31-1068GG comprising the S-layer protein and two copies of the IgG binding region of Protein G) can be used to disperse and functionalize oxidized multi walled CNTs. Following a simple protocol, a complete surface coverage with a long-range crystalline S-layer lattice can be obtained. When rSbpA31-1068GG was used for coating, the introduced functionality could be confirmed by binding gold labeled antibodies via the IgG binding domain of the fusion protein. Since a great variety of functional S-layer fusion proteins has already been described, our new technology has the potential for a broad spectrum of functionalized CNTs.
Article
Full-text available
This work develops a thorough review of bioimpedance systems for healthcare applications. The basis and fundamentals of bioimpedance measurements are described covering issues ranging from the hardware diagrams to the configurations and designs of the electrodes and from the mathematical models that describe the frequency behavior of the bioimpedance to the sources of noise and artifacts. Bioimpedance applications such as body composition assessment, impedance cardiography (ICG), transthoracic impedance pneumography, electrical impedance tomography (EIT), and skin conductance are described and analyzed. A breakdown of recent advances and future challenges of bioimpedance is also performed, addressing topics such as transducers for biosensors and Lab-on-Chip technology, measurements in implantable systems, characterization of new parameters and substances, and novel bioimpedance applications.
Article
Full-text available
S-layers commonly cover archaeal cell envelopes and are composed of proteins that self-assemble into a paracrystalline surface structure. Despite their detection in almost all archaea, there are few reports investigating the structural properties of these proteins, with no reports exploring this topic for halophilic S-layers. The objective of the present study was to investigate the secondary and tertiary organization of the Haloferax volcanii S-layer protein. Such investigations were performed using circular dichroism, fluorescence spectroscopy, dynamic light scattering and transmission electron microscopy. The protein secondary structure is centered on β-sheets and is affected by environmental pH, with higher disorder in more alkaline conditions. The pH can also affect the protein’s tertiary structure, with higher tryptophan side-chain exposure to the medium under the same conditions. The concentrations of Na, Mg and Ca ions in the environment also affect the protein structures, with small changes in α-helix and β-sheet content, as well as changes in tryptophan side chain exposure. These changes in turn influence the protein’s functional properties, with cell envelope preparations revealing striking differences when in different salt conditions. Thermal denaturation assays revealed that the protein is stable. It has been reported that the S-layer protein N-glycosylation process is affected by external factors and the present study indicates for the first time changes in the protein structure.
Chapter
Nanosciences are distinguished by the cross-fertilization of biology, chemistry, material sciences, and solid-state physics and, hence, open up a great variety of new opportunities for innovation. The technological utilization of self-assembly systems, wherein molecules spontaneously associate under equilibrium conditions into reproducible supramolecular structures, is one key challenge in nanosciences for life and non-life science applications. The attractiveness of such processes is due to their ability to build uniform, ultra-small functional units with predictable properties down to the nanometer scale. Moreover, newly developed techniques and methods open up the possibility to exploit these structures at meso- and macroscopic scale. An immense significance at innovative approaches for the self-assembly of supramolecular structures and devices with dimensions of a few to tens of nanometers constitutes the utilization of crystalline bacterial cell surface proteins. The latter have proven to be particularly suited as building blocks in a molecular construction kit comprising of all major classes of biological molecules. The controlled immobilization of biomolecules in an ordered fashion on solid substrates and their directed confinement in definite areas of nanometer dimensions are key requirements for many applications including the development of bioanalytical sensors, biochips, molecular electronics, biocompatible surfaces, and signal processing between functional membranes, cells, and integrated circuits.
Article
A novel electrochemical platform composed of water soluble polyethylene imine-reduced graphene oxide (PEI-rGO) composites and cholesterol oxidase (ChOx) based on layer-by-layer (LBL) technique has been triumphantly developed. The PEI-rGO composites could effectively improve the electron transfer ability between the assembled enzyme and the surface of electrode. Furthermore, LBL technique could immobilize more enzyme under mild conditions, which was benefit to keep the bioactivity of redox enzyme. TEM, UV–vis and FT-IR showed that the PEI-rGO composites behaved as individual nanosheet structure and were successfully prepared. The self-assembly process of (ChOx/PEI-rGO)n film was monitored by UV–vis spectroscopy, demonstrating that the ChOx had been immobilized orderly without denaturation. Owing to the good conductivity of the PEI-rGO composites, the electron transfer ability of the immobilized ChOx was greatly improved on glassy carbon electrode (GCE). The PEI/(ChOx/PEI-rGO)3/GCE showed excellent electrochemical performances toward the detection of cholesterol. Under optimized conditions, the fabricated electrode exhibited a wide linear range in the cholesterol concentration from 0.10 × 10⁻⁶ to 9.331 × 10⁻³ mol L⁻¹ with a detection limit of 0.021 μmol L⁻¹ estimated at a signal-to-noise ratio of 3. The apparent Michaelis-Menten constant (Km) was 0.0431 mmol L⁻¹. This result indicated that a good affinity between the immobilized ChOx and the cholesterol. Meanwhile, the biosensor presented good stability, reproducibility and anti-interference ability with good recoveries for real sample detection.
Article
The effect of amiloride, an acid-sensing ion channel (ASIC) blocker, on the formation of alamethicin ion channels in tethered zwitterionic phospholipid bilayers was investigated using electrochemical impedance spectroscopy (EIS) and polarization modulation infrared reflection absorption spectroscopy (PM-IRRAS). EIS measurements show a decrease in conductivity indicating that ion transported across the alamethicin-rich phospholipid membrane is inhibited by the presence of amiloride. The PM-IRRAS spectra indicate that amiloride has no effect on the secondary structure of the alamethicin peptide. The IR data suggests that amiloride does not block ion translocation through the alamethicin ion pore but prevents the insertion of alamethicin peptides into the biomembrane suppressing the formation of ion channels. The mechanism responsible for amiloride inhibition is explained in terms of the electrostatic interactions between the molecule and membrane surface. Amiloride is positively charged at physiological pH levels invoking a repulsive force on the positive pole of the alamethicin molecular dipole. These repulsive interactions prevent the insertion of the alamethicin peptides into the hydrophobic core of the lipid bilayer. The present paper offers a molecular view of the interactions that occur between amiloride, an ion channel inhibitor, with the lipids and ion channel forming peptides in model tBLMs.
Article
Biological systems contain various amazing examples that can display adaptive and active behaviors in response to external stimuli. The changes of skin color and/or shapes according to environmental changes are common examples of the adaptive and active behaviors of creatures for better survival. To mimic these outstanding sensing and actuating behaviors of creatures, a new generation of bio-inspired sensing and actuating materials, have been recently attracting extensive attentions. Based on stimuli-responsive materials, these bio-inspired sensing and actuating materials can also exhibit the capabilities of changing their colors and/or shapes upon specific external stimuli, which hold great promise in a variety of fields such as sensors, actuators, flexible electronics, soft robotics and various biomedical applications. This review summarizes the-state-of-the-art progress of bio-inspired sensing and actuating materials, with focuses on the design principles and the technologies for forming these materials, as well as the introduction of their potential for versatile applications. Finally, an outlook is presented for discussing the challenges and future direction for the bio-inspired sensing and actuating materials.