Cancer-Nano-Interaction: Comparison
Please note this is a comparison between Version 1 by Muthukumaran Packirisamy and Version 2 by Conner Chen.

In the targeted therapy, nanoparticles (NPs) with specific properties, nanomedicine, are designed to specifically transport therapeutic agents to tumor sites and to release under controlled conditions. This strategy could potentially overcome the limitations of conventional methods and improve the cancer treatment outcomes by distinguishing malignant cells from non-malignant cells and selectively kill malignant cells. Bio-distribution, biocompatibility, biodegradability, and systemic clearance are the general challenges of using NPs in the targeted therapy. An effective NP-based drug delivery system should predict and control the fate of NPs in the biological environment. To develop and achieve a sound and efficient NPs-based system, we need to enhance our understanding of the nano-bio-interaction (NBI) happening between nanomaterials and a complex heterogeneous biological environment. At the cellular level, the NBI occurs at the interface of NPs surface and cell membrane. The interaction behavior of NPs is highly dependent on the physical and chemical properties of NPs.

  • nano-bio-interaction
  • nanoparticle
  • mechanobiological properties
  • cancer cells
  • cell mechanics
  • migratory index

1. Basic Components of Cells and Biomechanics

Cells are the basic functional unit of living organisms. Unlike plant cells (prokaryotic), animal cells (eukaryotic) do not have enclosing cell walls, and they are surrounded only by cell membranes [1][80] (Figure 1a). The cellular membrane is a thin (5–10 nm thickness) and permeable lipid bilayer that controls the flow and movements of ions and molecules between the interior of cells (cytosol) and the extracellular environment [2][76]. To retain the structural integrity of cells, a specialized cellular structure is required. This cellular structure, the cytoskeleton, determines the mechanobiological properties of cells (such as stiffness) and influences the shape, division, and functions of cells [3][4][81,82]. In addition to cytoskeletal proteins, other cellular components such as the membrane, the nucleus, and the cytoplasm could impact the mechanic of cells to some extent [5][21].
Figure 1. (a) Schematic showing typical eukaryote cells, (b) microtubules (they are in curved format), (c) actin filaments or long stress fibers (they are in linear format), and (d) intermediate filament (they are extending from the nucleus to the periphery of cells).
A schematic drawing of a eukaryotic cell is illustrated in Figure 1a. The nucleus, the mitochondria, endoplasmic reticulum, Golgi apparatus, and cytoskeleton are the main components of the internal part of a typical eukaryotic cell [6][83]. The nucleus of the cell is the largest organelle among subcellular components [7][84] and is located within the central region of cells and includes two regions: the internal region containing DNA and proteins and the outer boundary of the nucleus or karyotheca, which is a lipid bilayer similar to the membrane of cells. Regulating the gene expression is the main role of the nucleus, and to some extent, contributes to the cell mechanics [8][85]. The cytoplasm of eukaryotic cells includes all the material within the cell and outside the nucleus, such as proteins, protein complexes, and organelles [9][86]. Cells are dynamic living systems, and their mechanobiological properties allow them to sense microenvironmental changes and convert stimuli and changes into biological signals [5][10][21,87].
The cytoskeleton is made with a complex network of protein fibers and biopolymers embedded in the cytoplasm. In addition to maintaining the integrity of cells, the cytoskeleton provides pathways for molecular motor proteins to shuttle cargo between different regions of cells and generate and transmit cellular forces [3][81]. In response to the mechanical changes in their microenvironments, cells can either reinforce their cytoskeleton by polymerizing their structural proteins or fluidize their cytoskeleton to reduce their stiffness. Microtubule (MT), intermediate filament, and actin filament (F-actin) are three major fiber parts of the cytoskeleton [11][88] (see Figure 1b–d). MTs (diameter ≈25 nm), composed of two subunits (α and β tubulins), are stiff and hollow structures of the cytoskeleton, radiating outward from the central organelle. Intermediate filaments provide the strength, integrity, and organization of both the cell and nucleus. The intermediate filaments (diameter ≈10 nm) are composed of various proteins known as protofilaments (protein lamin, vimentin, keratin). These proteins are bundled around each other in a rope-like structure to form the final intermediate filaments. Intermediate filaments have a Young’s modulus between 1 and 5 GPa, and their length is between 1 and 3 μm. Intermediate filaments within the cytoplasm act as “stress absorbers” and organizes the position of organelles in cells [12][89].
Actin filaments are the main structural component of cytoskeleton, and with the help of non-muscle myosin II proteins, they provide the required forces for the movement and contraction of cells [13][14][90,91]. The actin filaments are composed of two different actin chains: F-actin and G-actin, which are twisted around each other. G-actin monomers are polymerized to form stiff F-actin with a modulus elasticity between 1 and 2 GPa [2][15][76,92]. The dimeter of F-actin varies from 5 to 9 nm and has a length in the order of ten micrometers. F-actin filaments are linked to each other during cell migration to form branches at a 70-degree angle from the original filament, enabling the cell membrane to protrude outward [2][76]. With the aid of non-muscle myosin II, two or more F-actin filaments are bundled in parallel to provide stress fibers. Myosin II is a molecular motor protein that makes F-actin filaments slide past each other to generate forces within cells [13][90]. Myosins directly impact cell mechanics, elasticity, cells adhesion, and mechanosensing [16][93]. The force generated by myosin is transmitted through focal adhesions, aggregates of cytoplasmic proteins at the inner surface of the membrane, to the interface of the integrin and extracellular matrix, and these forces are considered as traction forces to help cells move forward during cell migration [2][12][76,89].
Among these three different components of the cytoskeleton, actin filament plays the most important role in the structural integrity and deformability of the cell. Intermediate filaments are also able to tolerate some reasonable extent of deformations by engaging in shear stress. MTs play an important role in the cytoskeleton stability but contribute less to the mechanical integrity than the two other filaments [3][17][18][81,94,95].

2. Techniques for Mechanobiological Characterizations

Various techniques can be implemented to measure the mechanobiological properties of single cells, such as viscosity and elasticity. The elastic modulus and viscosity modulus are typically used to express the mechanical properties of cells [19][11][25,88]. In the elastic modulus, the applied forces are related to cell deformation, while in the viscosity, time-dependent stress relaxation is measured in response to a step displacement [5][20][21,96]. Sufficient and controlled forces need to be applied to the cells to measure their mechanical properties. Based on the types of forces, different microrheological tools have been developed to measure mechanical properties. The most used methods for experimental measurements are shown in Figure 2. Classical methods such as Atomic Force Microscopy (AFM) [21][97], micropipette aspiration (MA) [22][98], optical tweezer (OP) [23][99], and magnetic twisting cytometry (MTC) [24][100] are preferred because of their high-resolution measurements. However, they are tedious, and the measurements take a long time. With MEMS (micro-electromechanical systems) [25][101] and microfluidic devices [26][102], mechanobiological properties can be measured at a higher speed, but their resolution is not as high as that of classical methods, and most of them are able only to measure deformability-related parameters, not elastic and viscosity modulus [27][103]. To enhance the accuracy of these methods, in parallel to experimental measurements, computational analyses need to be carried out; however, they may impose a level of complexity [2][76]Table 1 shows the limitation and advantages of different techniques. The technique can be chosen based on the type of cells and depending on the specific desired information.
Figure 2. Different tools for the mechanical characterizations of living cells, Atomic Force Microscopy (AFM), optical tweezer (OP), micropipette aspiration (MA), magnetic twisting cytometry (MTC), MEMS, and microfluidic techniques.
Table 1. Different techniques for mechanobiological measurements of cells.
Techniques Cell Type Mechanical Stimuli Important Parameters Advantages Limitations
Classical Techniques Atomic Force Microscopy (AFM) MCF7 [28][104];

Human bladder [20][96]
Cantilever micro indention Tip deflection, Young’s modulus High-resolution measurement; Provids both structural and mechanical information for local, whole, and interior measurements [29][21][23,97] Low throughput; Mechanical hitting of AFM tip may affect cell activities and position of probe; Requires a high-resolution microscope
Micropipette aspiration (MA) Human cartilage [22][98];

Colon cancer cells [30][105]
Negative force Young’s modulus Low-cost and well-established method Limited spatial resolution; Low throughput; For suspended cells only
Magnetic twisting cytometry (MTC) Melanoma [24][100];

MCF7 [31][106]
Force is applied by magnetic beads Stiffness and Young’s modulus Inducing little heat and photodamages compared to optical tweezer [32][10] Resolution limitation; Inducing non-uniform stress; Beads are localized randomly on cell; Attachment angle affects the displacement
Optical tweezers (OP) RBC [23][33][99,107] Laser-induced surface force Deformation index Without physical contact Only for suspended cells; Damaging consequence of optical heating on cells;

Limited magnitude of forces
Parallel plate Epithelial ovarian cancer [29][23];

MCF7 [31][34][106,108]
Shear stress Aspect ratio Homogeneity of the applied shear stress; Simplicity; Ability to study cell population Need bulky devices; Large amount of reagents; Difficult to visualize deformation
Microfluidic Techniques Fluid-induced deformation PBMCs [26][102] Fluid

shear stress
Deformation index, size High throughput; Simultaneously, other chemical assays can be done; The measurment can be done continuously; Contactless deformation; Applicable for both suspended and adhered cells Needing expensive high-speed camera for imaging
Constriction-induced deformation K562 [35][109];

MDA-MB-231 [36][110]
Mechanical squeezing Passage time, entry times, stiffness Wide-ranging applications in cell deformation; Applicable for different geometry structures;

Adjustable dimension for different cell types
Clogging and channel blockage; Possible effects of friction between cell and channel’s wall on measurements; Ignoring the effects of membrane rigidity and viscosity
Aspiration-induced deformation Neutrophils [37][24] Negative pressure Young’s modulus, cortical tension Straightforward method; Well-established mathematical model Leaking problem; Rectangle-like cross-section of microfluidic channels; Time-consuming process; Requiring high-vacuum pressure
Optical stretcher MCF7 [31][106];

MCF-7, MCF-10, MDA-MB-231 [38][111];

Red blood cells [23][99];

Melanoma cells [39][112]
Optically-induced surface forces Deformation index, cell elasticity No physical contact; Relatively high-throughput measurements Alignment problem;

Optical heating;

Thermal damage
Electrical-induced deformation MCF-10A, MCF-7 [40][113] Electroporation-induced swelling Deformation index, size of cells Fast heat dissipation; Better resolution;

Automation and parallelization of test with reduced amount of samples
High energy consumption and high voltage
MEMS Techniques Suspended microcantilever Circulating tumor cells [41][114]; Fibroblast [25][101] External actuator Frequency of cantilever, passage time, transit time All-inclusive systems; Parallel analysis; Better quality factor; Automation Fabrication is expensive; Non-transparent channels; High stiffness of silicon; calibration process
MEMS resonator MCF7 [42][115] External actuator Frequency of cantilever High throughput Expensive fabrication; Requiring external electrical system; Only for adherent cells

2.1. Classical Methods

Classical methods provide a high-resolution measurement on the mechanobiology of single cells; however, they suffer from tedious, low-throughput, and long-processing measurements. Among various measuring techniques, AFM has been extensively used to study the mechanobiological properties of nano-treated single cells. This section will mostly focus on the AFM technique and briefly discuss other main classical methods.
Optical tweezer is one of the popular classic techniques for the manipulation and mechanical characterizations of suspended cells. In this technique, a focusing laser beam, introduced from a high numerical aperture objective, is utilized to trap single cells close to the beam focus. OP could apply time-varying stretching forces ranging from 0.1 to 100 pN onto the trapped cells to characterize mechanical properties. The mechanical properties of cells can be quantified by calibration techniques. Even though OP is an effective technique for mechanobiological measurements, it may induce unwanted detrimental effects to cells due to using a high-powered laser, altering the mechanics of cells [11][33][88,107].
Magnetic twisting cytometry is another well-established method for the mechanobiological characterization of living cells. In this technique, magnetic beads attached to the cells impose a quantified external force on the portion of cells under an external magnetic field. The magnetic-field-induced bead displacement is tracked to characterize the viscoelastic properties of cells. The applied stress can be controlled by translocating and regulating the external field. MTC offers various advantages over other methods. MTC generates both liner force and twisting torque, magnetic manipulation does not cause light-induced damage as in optical trapping, and MTC allows parallel simultaneous measurements. There are also some disadvantages associated with this method. It is not easy to control the region where beads are bound to cells, and more importantly, beads lose magnetization with time and need to be re-magnetized to maintain the torque applied [31][43][106,116].
Micropipette aspiration is a traditional method used to deform cells by imposing gentle suction to a micropipette. MP deforms individual cells in whole, and their deformations are measured to quantify the mechanobiological properties. This technique applies a small negative pressure into the glass micropipette with an inner diameter smaller than cells, causing them to deform and elongates a portion of them into the pipette. Several parameters such as the suction pressure, the diameter of pipette orifice, and the protrusion length of cells in the pipette are measured to derive the mechanobiological properties (stiffness) of aspirated cells. This technique has been used to measure the mechanical properties of numerous types of cells such as HeLa [44][117] and human leukocytes cells. Although MA offers a straightforward and well-established method for the mechanical characterization of individual cells, requiring special equipment and involving delicate procedures are the main challenges [45][46][47][75,118,119].
Over the past three decades, AFM has been used as a key tool for simultaneous morphological and mechanobiological characterizations of different living cells, such as human kidney cells [48][120], human bladder cancer [49][121], ovarian cancer cells [50][122], and breast cancer cells [51][123]. The AFM method was introduced in 1986 to imaging and manipulating matter at molecular and cellular scales [52][124]. AFM can be used in liquid environments, and it has a flexible cantilever (several micrometers) at the end to probe the sample topography and measure forces between the tip and sample with piconewton sensitivity. The AFM technique is not a high-throughput method, but it has a simple principle of operation, allowing users to adjust this technique to measure the desired mechanobiological property. However, it has several intricacies that make the acquisition of quantitative data complex. To apply deformations, the AFM tip is vertically indented into the cell until the pre-set loading force, and the applied force, which is proportional to the cantilever deflection, is recorded (Figure 3). The motion of the cantilever can be measured optically by a beam of laser or through sensing elements built into the cantilever itself. Then, the AFM tip is controlled to return to its original position. During the approach–retract process, the cantilever deflection versus the vertical displacement of the AFM probe is recorded. The approach curve along theoretical models can be used to extract the cellular Young’s modulus, while the retract curve can be used to quantify the adhesion force [21][53][54][55][97,125,126,127].
Figure 3. (a) AFM indentation and interpretation of the curve: (1) AFM above the cell surface, (2) AFM in contact with the cell surface, (3) motion of the AFM cantilever to contact the cell surface and indent into cell until the setpoint, (4) AFM tip detaching from the sample (AFM tip-cell adhesion), (5) returning to initial position. Reprinted with permission from [56][128]. Copyright Journal of Visualized Experiments 2013. (b) Cell elasticity measurement of human breast cancer cells with AFM and visualization of actin filaments in both cell lines. Reprinted with permission from [28][104]. Copyright Elsevier 2008.
There are different contact models to extract the mechanical properties from the AFM-obtained curve. The most commonly used models for estimating the cellular Young’s modulus include Hertz, Sneddon, Johnson–Kendall–Roberts (JKR), Derjaguin–Muller–Toporov (DMT), and Oliver–Pharr [29][23]. Each model can be used based on the different AFM tip geometries and sample properties. The Hertz is the most frequently used model to approximate the contact between the AFM tip and the sample. Three assumptions are considered for using the Hertz model: the AFM tip is a perfect sphere, linear strain–stress relationship (maximum 30% indentation of sample thickness), and the sample deformation is fully reversible. If these conditions are met, the Hertz model could extract mechanobiological properties by defining the contact point, which is difficult to determine, particularly for mammalian cells with complex surface morphologies [29][21][23,97]. With AFM, forces as small as 10-11 N can be measured.

2.2. MEMS- and Microfluidic-Based Techniques

Although classical methods provide high-resolution measurements, single-cell analysis with classical methods is a very time-consuming process. MEMS-based approaches, including microfluidic techniques [57][129], could provide high-throughput alternatives that can clinically be used for the deformability characterization of individual cells [58][130]. Microfluidic-based systems could characterize the mechanobiological properties of thousands of cells in a short time. Their resolutions might not be competitive with classical tools, so they mostly focus on deformability-related parameters rather than elastic properties. In the following, a few prominent techniques are discussed.
Researchers at MIT University developed a suspended MEMS resonator [25][101] to characterize the mechanobiological properties of ≈105 single cells per h by integrating a constriction channel to the device at the apex of a micro-cantilever. By measuring the velocity and transit time of cells passing through the constriction channel, they evaluated the stiffness and friction of the cells. In another study, a MEMS resonator was proposed by Corbin et al. [42][115] to quantify the mechanobiological properties of human breast cancer cells (Figure 4a). They modeled the MEMS platform and the cells as a two-degrees-of-freedom system to estimate the mechanobiological properties of cells through the vibrational behavior of the microsystem. Then, they studied the shift resonant frequency of the system after and before chemically fixing the adherent cells to the resonating platform to predict their viscosity and elasticity. MEMS systems offer automated and rapid measurements; however, for mechanobiological measurement, they suffer from non-transparency and high stiffness compared to living cells [59][131].
Figure 4. Different microfluidic and MEMS techniques for the mechanobiological characterization of cells: (a) Modeling cells as a two-degrees-of-freedom system and measuring their viscoelastic properties using a MEMS resonator. Reprinted with permission from [42][115]. Copyright Royal Society of Chemistry 2015. (b) Constriction channels to induce mechanical deformation onto oocyte cells and measuring their deformations as they pass through the tight channel. Adapted with permission from [60][135]. Copyright Springer Nature 2015. (c) A micro-aspiration integrated into a constriction channel for quantifying the deformability properties of cells by measuring the threshold pressures. Reprinted with permission from [61][136]. Copyright Royal Society of Chemistry 2012. (d) Hydrodynamic stretching of cells and high-throughput assay to measure the index of cells and investigate the deformability of cells. Reprinted with permission from [26][102]. Copyright (2012) National Academy of Sciences.
In contrary to MEMS (normally made of silicon) systems, polymer-based microsystems offer more advantages. The mechanical properties of cells are closer to the mechanical properties of these polymers, so their in vivo microenvironment can be mimicked better. Due to the optical transparency of the polymer, the behavior of living cells and their deformations can be monitored with light microscopy at the same time [27][62][103,132]. With the aid of microfluidic devices, fast mechanobiological assays can be performed using reduced quantities of samples. Microfluidic techniques can be classified based on the mechanical stimuli used to deform the cells. Monitoring the cell movement as it passes through a constriction channel is one of the most straightforward techniques for studying the mechanobiological properties of living cells (Figure 4b). Under a hydraulic pressure difference, target cells are squeezed by the wall of the channel, which is marginally smaller than the diameter of the cell. With the aid of the constriction channel, various parameters such as entry time, passage time, elongation, and recovery time can be quantified. Clogging and channel blockage are the main limitations of these devices [63][64][133,134].
Deformation can be made with the aspiration technique in which the concept of conventional MA is mimicked to measure the mechanobiological properties of the cell (Figure 4c). A cell is partially aspirated into a microfluidic channel and deformed through a series of funnel-shaped constrictions. Meanwhile, the elongation of the cell is measured by a microscope and camera to infer the rheological properties of living cells [37][61][24,136]. Living cells can also be exposed to the hydrodynamic forces and deformation by designing microchannels in which various fluid stress stimuli are generated [34][108] (Figure 4d). In contrast to the mechanical confinement-induced deformation, cells can be deformed by shear stress within microchannels with a larger diameter than the cell’s diameter. The deformation index (DI) or stretch ratio is defined as the ratio of both axes of the cross-sectional area of the deformed cell and can be quantified by high-speed imaging. Using a high-speed imaging camera is one of the limitations of microfluidic-based fluid-induced deformation [25][41][101,114]. The optical stretcher is a popular method for the mechanobiological characterization of the suspended cells. This technique could be used to trap and stretch single cells based on the laser-induced momentum transfer. The stretching forces can be affected by the size, type of cells, refractive index, and laser power. Although optical stretching can measure the mechanobiological properties of cells, the imposing forces are not large enough to promote significant deformability to simulate in vivo conditions encountered by migrating cancer cells. Furthermore, the effects of the laser beam on the mechanobiological properties of cells are unknown and need further studies [23][58][99,130]. Electrical fields also can be implemented for the mechanobiological characterization of cells [65][40][22,113]. Whenever a single cell experiences an externally applied electrical field, it is swelled or expanded in size, which is a phenomenon known as electroporation. The electrical field increased the conductivity and permeability of the cell plasma membrane. The influx of small molecules through the open pores in the cell membrane causes the swelling and expansion of cells. Swelling ratios (before and after establishing voltage) of cells can be recorded to evaluate the deformability of cells.

3. Impacts of Nanoparticles on Structural Elements and Morphology of Cells

Any changes in the cytoskeletal structure of cells could lead to the alterations of the mechanobiological properties of cells. In order to understand the effects of NPs on the mechanobiological properties of cells, we need to study the physiochemical interactions between NPs and the three main filamentous proteins: intermediate filament, actin filament, and MT. In some studies, the disruption of subcellular structures has been reported due to the NPs uptake; however, the consequences of those changes to fundamental biological processes have been less investigated. The cytoskeleton is responsible for the basic functions of cells: (a) to preserve the morphology of cells, (b) to anchor organelles, (c) to physically connect cells to the microenvironment, (d) to produce internal forces for cells movement, (e) to help cells for division, and (f) endocytosis [3][81]. Therefore, any changes in the cytoskeleton organization could induce cellular dysfunction (see Table 2).
Most NPs are thought to penetrate cells through forming vesicles, and these membrane-bound vesicles transport NPs along MT to intracellular compartments. During this process, the NPs might have indirect interactions with cytoskeletal proteins and change their organizations. It is not clear how they interact with those proteins while they are encapsulated inside lysosomes and endosomes [66][17]. However, there are some evidence showing that NPs could directly interact with the cytoskeletal proteins. It has been found that carbon nanomaterials enter cells by adhesive interaction, enabling them to freely swim in the cytoplasm and directly interact with the subcellular structures of cells. For example, Lundqvist et al. [67][59] found the presence of MT in the protein corona formed around the SiO2 NPs, suggesting the direct NPs–proteins interactions. Direct or indirect interaction with NPs may negatively affect the biological functions [68][69][70][137,138,139]. Tian et al. [71][140] showed that single-wall carbon nanotubes could enter cells and alter cell morphology by disturbing the actin networks. They observed that these NPs cause an irregular actin network in comparison to untreated cells. Various NPs-related parameters such as the shape, size, surface chemistry, concentration, and incubation time are important in assessing the toxicity of nanomaterials in cytoskeleton. The shape of the NPs can induce different effects on the cytoskeletal structure of cells. It has been shown that unlike silica NPs with small aspect ratios, silica nanorods with large aspect ratios can largely change the organization of the actin filament, particularly in the vicinity of the cell membrane, resulting in serious damages to the cytoskeletal structures [72][73][74][141,142,143]. Ibrahim et al. [75][144] used different techniques such as SEM, TEM, and immunofluorescence analysis to study the cytoskeletal changes in osteoblast-like cells underexposure of titanium-based orthopedic and dental implants NPs (nano-Tio2). Smaller particles were found to be more disruptive to the actin and microtubule cytoskeletal network in comparison to larger particles. In another work, Holt et al. [76][145] used fluorescence lifetime microscopy to study the interactions of single-wall carbon nanotubes with HeLa cells. They showed that nanotubes preferentially interact with F-actin compared to G-actin and dramatically change their distribution. NPs even could disrupt the MT and actin network at non-toxic concentrations. Liu et al. [77][146] showed that bare gold NPs with the size of 20 nm alter the microfilament arrangement of endothelial cells more than NPs with the size of 5 nm. In this study, five types of gold NPs with different sizes and surface coatings were used to determine the viability and cytoskeletal change of endothelial cells. They found that gold NPs do not affect the viability of cells; however, the force balance between intracellular tension and paracellular forces is broken in 20 nm bare gold NPs-treated cells. In another study, the sub-lethal concentration of silver NPs was used to investigate cytoskeletal changes in neural cells [78][147]. They found that the percentage of AgNP-treated cells containing inclusions is doubled compared to control cells, indicating a significant disruption of actin filaments.
In vitro alternations in MT and F-actin concentrations and cytoskeletal destabilization have also been observed in cells, particularly under high concentrations of NPs. For example, Ogneva et al. [79][148] showed reduced F-actin content in silicon-treated mesenchymal stem cells compared to control cells (Figure 5a). Pisanic et al. [80][149] studied the effects of NPs concentrations on neuron cells. They found that by increasing the concentration of metal oxide NPs, the density of actin filaments is reduced, preventing them from getting mature under the stimulation of nerve growth factors. Mironava et al. [81][37] revealed that the cellular uptake of gold NPs disrupts actin fibers of human dermal fibroblast cells, and in contrast to the extended actin in control cells, in treated cells, actin filaments are broken and appeared as dotes (Figure 5b). However, no significant changes were found in actin or beta-tubulin protein levels. Choudhury et al. [82][150] studied the binding of nanosphere gold NPs to MT in the cell-free systems as well as in human lung carcinoma cells (A549) using Raman measurement, Fourier transform infrared (FTIR), and other imaging techniques. Their findings showed that gold NPs depending on their size and concentration might inhibit the polarization of MT. They also observed that MT networks are damaged and shrunken upon interaction with gold NPs compared to control cells.
Figure 5. (a) Mesenchymal stem cells treated with silica (Si) and silica–boron (SiB): F-actin detected with red TRITC-phalloidin staining, and DNA stained with blue DAPI. Actins in control cells are packed longitudinally, while they are arranged transversally in treated cells [79][148]. Copyright 2014, Open Access Springer Journals. (b) Fluorescent imaging of human dermal fibroblasts stained for F-actin after three days exposure to gold NPs. F-actins appeared to be in dotted format compared to control cells. Reprinted with permission from [81][37]. Copyright Informa UK Ltd. 2010. (c) SEM images of MDA-MB-231 cells treated with fullerenol NPs compared to control cells. Treated cells show shorter protrusions in comparison to control cells, and the concentration of actin fibers has reduced after the uptake of NPs [83][151]. Copyright 2019, Open Access, Journal of Nanobiotechnology.
NPs might have different affinities to different subcellular structures depending on their physicochemical properties [84][85][152,153]. Wen et al. [86][154] found that silver NPs tend to bind actin rather than tubules under electrostatic interactions. They used imaging techniques to visualize the organization of actin and tubulin proteins after treating with silver NPs (size 30 nm). They observed that the secondary structures of actins and tubules are changed due to the interaction with NPs, and alpha-helices of both proteins are decreased while their beta-sheets are increased. NPs-induced cytoskeletal changes could also cause significant morphological changes [87][88][155,156]. Rasel et al. [89][157] observed morphological changes of osteoblast cells after treating with boron nitride NPs, while they do not have adverse effects on the viability and the metabolism of cells. Ali et al. [90][158] showed that gold nanorods could change the cytoskeletal structure of oral squamous cell carcinoma. They observed morphological changes in cytoskeleton protrusions (filopodia and lamellipodia) when incubating cells with integrin-targeted gold nanorods. Qin et al. [83][151] found that the NPs-treated breast cancer cells have reduced the number and length of filopodia compared to control cells, causing them to lose their adhesion to the extracellular matrix (Figure 5c). Patra et al. [91][159] observed that gold NPs damage the cytoskeletal structure and induce profound morphological changes in human carcinoma cells (A549). Subbiah et al. [92][160] studied the morphological changes of A549, NIH3T3, and HS-5, and they found that silver NPs may induce changes in the topography of cells lines and treated cells appeared more rounded than untreated cells. Morphological changes could be influenced by concentration or incubation time. Wu et al. [93][161] proved that the density of filamentous proteins is reduced by increasing the concentration and exposure time of gold NPs in human aortic endothelial cells, causing topographic changes in the cell surfaces. In another work, Pernodet et al. [94][162] found that citrate-gold NPs profoundly affect the cell morphology of human dermal fibroblasts when the concentrations and exposure time are increased. They observed that the density of actin filament decreases in the presence of NPs by extending the exposure time, showing that the actin fibers are depolarized due to the cellular uptake of NPs.
In summary, in order to study the toxicity of nanomaterial in cells, the interaction of nanomaterial with subcellular structures, particularly cytoskeleton, needs to be taken into account. Nanomaterials, even under low concentration due to direct and indirect interactions with filamentous networks of cells, could change the main cellular structure and lead to mechanobiological changes in cells.
Table 2. Cytoskeletal changes due to the NPs–protein interactions.
Author Cell Type NPs Type Methods Cytoskeleton Changes
Pernodet et al., 2007 [94][162] CF-31 (human dermal fibroblast) Gold NPs (13 nm) TEM, Confocal Imaging, Migration Assay Modification in actin networks; NPs impaired motility and adhesion
Pi et al., 2013 [95][163] MCF-7 (breast cancer) Selenium NPs AFM, Confocal Microscopy The organization of F-actin is changed, and they are aggregated; Actin concentration is reduced
Choudhury 2013 [82][150] A549 (lung cancer) Citrate-capped Gold NPs (20–60 nm) Raman, FTIR, TEM, Darkfield Microscopy, UV-Visible Spectroscopy Inhibiting the polarization of MT; MT structures are damaged, affecting the dynamic equilibrium
Qin et al., 2018 [83][151] MDA-MB-231 (breast cancer) Fullerenol NPs SEM, Fluorescence Imaging, AFM, Scratch Assay The concentration of actin is reduced, the migration speed is reduced, disturbing actin assembly
Hot et al., 2012 [76][145] HeLa (cervical cancer) Single-wall carbon nanotube (1 ± 0.3 nm) Fluorescence Imaging Microscopy NPs cause cells to have shorter F-actin; Traction force is reduced; NPs do not affect G-actin and myosin II
Huang et al., 2010 [72][141] A375 (melanoma) Silica NPs (MSNs) TEM, Confocal Microscopy, Western Blot The actin structure is disorganized and disrupted with NPs; Cell migration is reduced
Patra et al., 2007 [91][159] A549 (lung cancer) Gold NPs Confocal Microscopy The morphology is changed; Treated cells are rounded compared to non-treated
Pisanic et al., 2007 [80][149] PC12M (brain) Fe2O3 NPs TEM, Western Blot, Fluorescent Microscopy Reduction in the formation of actin microfilaments; They are less organized; NPs diminish the ability for differentiation
Wu et al., 2012 [93][161] HAEC (aortic endothelial cells) Diesel exhaust particles (DEPs) AFM, Fluorescent Imaging Cells became degraded; Cellular cytoskeletal structures were impaired
Wen et al., 2013 [86][154] Acting and tubulin proteins (cell-free system) Silver NPs TEM, Hyperspectral Imaging, Inducing changes in the secondary structures; Silver NPs tend to bind actin vs. tubulin
Cooper et al., 2015 [78][147] B35 (neuroblastoma) Silver NPs Immunocytochemistry NPs induce F-actin inclusion, disrupting the actin function
Rasel et al., 2015 [89][157] Osteoblast cells Boron nitride NPs AFM, TEM, X-Ray They do not affect the morphology of cells
Liu et al., 2017 [77][146] HUVEC (Endothelial cells) Gold NPs-coated with PEG (20 nm) Fluorescent Microscopy, Traction Force Microscopy NPs re-arranged actin filaments; Inhibition of Rock activity reduced the polymerization of actin; Reducing the focal adhesion
Vieira et al., 2017 [96][164] CCD1072Sk (Normal cells-skin) Gold NPs and silver NPs Immunofluorescence Imaging, Cytofluorometry NPs impair the F-actin;Cytoskeletal reorganization; Cells lose the cell polarization (without losing their viability)
Ali et al., 2017 [90][158] HSC-3 (tongue cancer) Gold nanorods coated with PEG and REG Western Blot, DIC Microscopy, Scratch Assay The cytoskeletal proteins are rearranged; Cytoskeletal protrusions (filopodia and lamellipoda) are reduced
Beaudet et al., 2017 [97][48] HeLa (cervical cancer) AuNPs, Swarna Bhasma Fluorescent Imaging Larger particles disrupted the microtubules networks
Ibrahim et al., 2018 [75][144] SaOS-2 (bone cancer) TiO2 spherical NPs Hyperspectral Imaging, Fluorescent Imaging, Western Blot The actin and microtubule cytoskeletal networks are disorganized
Kralovec et al., 2020 [74][143] A549 (lung cancer) Fe3O4@SiO2 Fluorescent Imaging, Western Blot Severe disruption of the actin filament and microtubules
Kota et al., 2021 [98][165] VSMCs (vascular smooth muscle cells) ZIF-8 NPs AFM, Fluorescent Imaging, Polymerization Assay Morphological changes and cytoskeletal disorganization were observed; NPs caused changes in actin filaments at basal and apical surfaces.
Video Production Service