Journal of Physics D: Applied Physics ACCEPTED MANUSCRIPT • OPEN ACCESS Between life and death: strategies to reduce phototoxicity in super- resolution microscopy To cite this article before publication: Kalina L. Tosheva et al 2020 J. Phys. D: Appl. Phys. in press https://doi.org/10.1088/1361-6463/ab6b95 Manuscript version: Accepted Manuscript Accepted Manuscript is “the version of the article accepted for publication including all changes made as a result of the peer review process, and which may also include the addition to the article by IOP Publishing of a header, an article ID, a cover sheet and/or an ‘Accepted Manuscript’ watermark, but excluding any other editing, typesetting or other changes made by IOP Publishing and/or its licensors” This Accepted Manuscript is © 2020 IOP Publishing Ltd. As the Version of Record of this article is going to be / has been published on a gold open access basis under a CC BY 3.0 licence, this Accepted Manuscript is available for reuse under a CC BY 3.0 licence immediately. Everyone is permitted to use all or part of the original content in this article, provided that they adhere to all the terms of the licence https://creativecommons.org/licences/by/3.0 Although reasonable endeavours have been taken to obtain all necessary permissions from third parties to include their copyrighted content within this article, their full citation and copyright line may not be present in this Accepted Manuscript version. Before using any content from this article, please refer to the Version of Record on IOPscience once published for full citation and copyright details, as permissions may be required. All third party content is fully copyright protected and is not published on a gold open access basis under a CC BY licence, unless that is specifically stated in the figure caption in the Version of Record. View the article online for updates and enhancements. This content was downloaded from IP address 5.188.216.148 on 26/01/2020 at 06:47 ----!@#$NewPage!@#$---- Between life and death: strategies to 1 reduce phototoxicity in Super-Resolution 2 Microscopy 3 4 Kalina L. Tosheva1, Yue Yuan1, Pedro Matos Pereira2, Siân Culley1,3, and Ricardo Henriques1,3 5 1MRC Laboratory for Molecular Cell Biology, University College London, London, UK 6 2ITQB-NOVA, Oeiras, Portugal 7 3The Francis Crick Institute, London, UK 8 9 Super-Resolution Microscopy enables non-invasive, molecule-specific imaging of the internal structure and 10 dynamics of cells with sub-diffraction limit spatial resolution. One of its major limitations is the requirement for high- 11 intensity illumination, generating considerable cellular phototoxicity. This factor considerably limits the capacity for 12 live-cell observations, particularly for extended periods of time. Here, we give an overview of new developments in 13 hardware, software and probe chemistry aiming to reduce phototoxicity. Additionally, we discuss how the choice of 14 biological model and sample environment impacts the capacity for live-cell observations. 15 16 Phototoxicity | Photodamage | Super-Resolution Microscopy | Fluorescence 17 Correspondence: s.culley@ucl.ac.uk, r.henriques@ucl.ac.uk 18 Page 1 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- Introduction 19 The spatial resolution of an imaging system is 20 defined as the capacity to distinguish closely 21 separated features; in light microscopy, this is 22 limited by diffraction to ~ 200-300 nm 23 . Consequently, microscopy approaches developed 24 to achieve resolutions beyond this limit are termed 25 ‘Super-Resolution Microscopy’ (SRM) [1]. SRM 26 techniques that have recently gained popularity, 27 such as Photoactivated Localisation Microscopy 28 (PALM) [2], Stochastic Optical Reconstruction 29 Microscopy (STORM) [3], Structured Illumination 30 Microscopy (SIM) [4] and Stimulated Emission 31 Depletion (STED) Microscopy [5], have enabled 32 biological discoveries inaccessible to conventional 33 microscopy [6]–[9]. Alongside increased spatial 34 resolution, SRM retains many desirable features of 35 light microscopy techniques, including molecule- 36 specific labelling and the potential for live-cell 37 imaging, unavailable to other high-resolution 38 techniques, such as electron microscopy. However, 39 the live-cell imaging potential of SRM has remained 40 largely untapped as the requirements of most SRM 41 techniques pose various challenges for exploring 42 dynamic processes under physiological conditions. 43 In contrast, such limitations are absent when using 44 fixed specimens. 45 Resolution increase in SRM is generally achieved at 46 the cost of high-intensity illumination [10]. These 47 requirements result in photobleaching, defined as 48 irreversible loss of fluorescence during imaging. 49 However, of greater importance to live-cell imaging 50 is sample health. Thereby, the dependency of SRM 51 on illumination intensities orders of magnitude 52 higher than conventional microscopy (W/cm2 - 53 GW/cm2 compared to mW/cm2 - W/cm2) makes 54 phototoxicity the biggest concern when employing 55 these techniques [10], [11]. In the context of 56 microscopy, phototoxicity is a broad term 57 encompassing physical and chemical reactions 58 caused by the interaction between light and cellular 59 components, with detrimental effects on the latter 60 [12], [13]. Correct biological interpretations from live- 61 cell imaging can only be achieved if the observed 62 phenomena progress with minimal perturbation [14]. 63 A multitude of properties of the sample and the 64 imaging can influence phototoxicity and can thus be 65 optimised for improving SRM for live-cell imaging 66 (Fig. 1). 67 68 Fig. 1 Summary of the factors that can be optimised to reduce 69 phototoxicity in Super-Resolution Microscopy. 70 On a molecular level, the main causes of 71 phototoxicity are photochemical processes that 72 directly damage intracellular components or lead to 73 the production of toxic products within the cell or in 74 its direct environment [15], [16]. The detrimental 75 effects of ultraviolet (UV) light on cells is particularly 76 well characterised; illumination with UV light can 77 trigger the so-called 'UV-response' (Fig. 2aFig. 1 78 Summary of the factors that can be optimised to reduce 79 phototoxicity in Super-Resolution Microscopy.) [17], [18], 80 Page 2 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- DNA-strand breaks [19], [20], and thymidine 81 dimerisations [21] (Fig. 2bFig. 1 Summary of the factors 82 that can be optimised to reduce phototoxicity in Super- 83 Resolution Microscopy.), leading to mutations and 84 downstream apoptosis [22], [23]. Additionally, both 85 UV and visible wavelengths can excite other 86 endogenous photoactive molecules in the cell, such 87 as NAD(P)H [24], flavins [25], [26] and 88 porphyrins[27], [28]. Furthermore, in fluorescence 89 microscopy there are phototoxic effects associated 90 with the fluorescent molecules required for labelling 91 structures [15], [29]. Upon illumination, both 92 endogenous and exogenous photoactive molecules 93 can be excited to reactive states (most commonly 94 long-lived triplet states) capable of undergoing redox 95 reactions that lead to formation of reactive oxygen 96 species (ROS) (Fig. 2cFig. 1 Summary of the factors that 97 can be optimised to reduce phototoxicity in Super-Resolution 98 Microscopy.). ROS are considered the major 99 contributors to phototoxicity [12], [13]. Their 100 production can occur via direct reaction between the 101 excited molecule and environmental molecular 102 oxygen or via reactions with other neighbouring 103 molecules that generate free radicals [30]. ROS 104 have a broad range of negative effects ranging from 105 oxidising proteins, lipids, and DNA, as well as 106 systematic effects such as disrupting the redox 107 homeostasis, signalling pathways and cell cycle 108 [12], [31]. Notably, ROS production correlates with 109 illumination intensity and photobleaching [12], [15], 110 both of which are issues present in SRM. As a result, 111 there is considerable interest in developing SRM 112 technologies for improved sample health. Here, we 113 will outline the progress in hardware, software and 114 probe development as well as choices in biological 115 model and sample preparation that can help 116 improve live-cell SRM (Fig. 1) 117 Quantifying phototoxicity in microscopy 118 Measuring phototoxicity in microscopy is not a trivial 119 problem, as evidenced by the sparsity of the 120 121 Fig. 2 Interactions of light with cellular components leading to 122 phototoxicity. a UV light can trigger apoptosis by inducing Fas 123 receptor-mediated signalling pathways. b UV light can directly 124 damage DNA by causing strand breakage (top) or thymidine 125 dimerisation (bottom), causing mutations and inducing DNA 126 damage responses. c UV and visible wavelengths can excite 127 photoactive molecules leading to chemical generation of ROS, 128 which can then damage cellular components. 129 available literature [12], [13]. This is not entirely 130 surprising, as phototoxicity is mediated by many 131 factors (Fig. 1). These include illumination 132 wavelength, intensity and duration of illumination, 133 the illumination regime (e.g. LED illumination vs. 134 laser illumination, laser-scanning vs. light-sheet), 135 and the number of imaged 3D-planes [32]–[37]. 136 Additionally, illumination tolerance can vary 137 substantially between specimens (see Biological 138 Page 3 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- models and sample preparation section), and 139 experimental stress can influence a specimen’s 140 sensitivity to illumination [14]. For example, a 141 procedure as routine as transfection or the addition 142 of a drug has been shown to dramatically increase 143 cellular sensitivity to light [10], [38]. Therefore, steps 144 must be taken to reduce avoidable experimental 145 perturbations which can influence the well-being of 146 the sample in an illumination-independent manner, 147 e. g. suboptimal environmental conditions 148 (temperature, pH, etc.) [39] or complex sample 149 mounting. 150 How does one approach a problem as versatile as 151 measuring phototoxicity? An intuitive and common 152 way of assessing photodamage is by measuring 153 photobleaching [40]–[43]. However, phototoxicity 154 and photobleaching are two separate processes; 155 while toxic ROS are produced during 156 photobleaching, they can also be generated 157 independently of this process [15], [44]. Therefore, 158 phototoxicity can commence prior to a detectable 159 reduction in fluorescence, making photobleaching 160 an unreliable read-out for photodamage in the 161 context of live-cell imaging [12]. More importantly, 162 photobleaching rates give no information on the 163 health and viability of the specimen. Thus, a better 164 phototoxicity measure would have a read-out related 165 to the properties of the sample itself, rather than the 166 properties of the fluorescence [34]. 167 There are several in vitro assays for post-imaging 168 assessment of the health and viability of a specimen 169 that can be used to indicate whether phototoxicity 170 occurred (Fig. 3a). These include detection of toxic 171 ROS, fragmentation and oxidation of DNA strands, 172 reduced metabolic activity, loss of membrane 173 integrity and the expression of stress- and 174 apoptosis-related proteins [45]–[50]. The 175 advantages are that these assays provide an 176 inexpensive and simple specimen viability 177 evaluation. Thus, different illumination conditions 178 can be tested and viability can be assessed each 179 time. However, for such assays the measurement is 180 limited to a single timepoint and imaging cannot be 181 recommenced after performing the assay. 182 183 Fig. 3 Methods for measuring phototoxicity. a `Destructive 184 read-outs' are techniques prohibiting further imaging of the 185 sample. These include blotting for phosphorylated forms of 186 proteins present in damage-activated pathways [51] and flow 187 cytometry for determining the population of cells expressing, 188 for example, apoptotic markers such as annexin V. b 189 `Fluorescent reporters' are additional indicators added to the 190 sample during imaging whose fluorescence signal changes in 191 response to e.g. intracellular Ca2+ concentration (top) or 192 mitochondrial membrane potential (bottom). `Label-free 193 methods' of quantifying phototoxicity involve: c short-term 194 observation of cell division and morphology and d proliferation 195 of cells in culture following imaging. 196 A more dynamic and practical approach entails 197 monitoring changes in relevant biological 198 parameters during imaging (Fig. 3b, c). Cellular 199 Page 4 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- processes which are particularly photosensitive (i.e. 200 rapidly perturbed by light) are excellent read-outs. 201 For example, a commonly employed method is 202 measuring changes in cytosolic calcium 203 concentration using calcium-sensitive fluorescent 204 probes [50], [52]–[54] (Fig. 3b, top). This strategy 205 was used to evaluate live-cell STED microscopy by 206 monitoring differences in intracellular calcium 207 concentration between control cells and STED- 208 imaged cells. This method showed that while there 209 is little difference between calcium concentration in 210 control and STED-imaged cells when using 211 excitation and STED-lasers with wavelengths >600 212 nm, responses indicative of cell damage were 213 observed with shorter illumination wavelengths and 214 when longer STED-laser dwell times were used [29]. 215 Other processes exist that make suitable read-outs 216 for phototoxicity, including changes in mitochondrial 217 membrane potential [41], [51] (Fig. 3b, bottom), 218 reduction of chromosome movement [55] and 219 slowing of microtubule growth [10]. It is worth 220 highlighting that, regardless of the process chosen, 221 care must be taken when employing fluorescent 222 probes for visualising these read-outs [46], [56]. 223 There are image-based phototoxicity 224 measurements that can be performed without 225 fluorescent labels. These often rely on identifying 226 changes in cell morphology indicative of entry into 227 apoptosis, such as blebbing or cell rounding [10], 228 [14], [51], [57], for example by using transmitted light 229 imaging (Fig. 3c). This approach was recently used 230 to train a deep convolutional neural network, 231 referred to as ‘DeadNet’, with the objective to 232 automate phototoxicity detection and quantification 233 from transmitted light images [58]. However, despite 234 widespread use, relying on morphology as a read- 235 out has two limitations: first, even experienced 236 researchers can struggle to identify subtle changes 237 in morphology, thus biasing the results (e.g. by 238 annotating ambiguous cases incorrectly [58]; 239 second, when changes become obvious, they 240 usually represent an extreme phenotype indicative 241 of irreversible damage. Thus, they cannot account 242 for early damage that may arise even as cells 243 display a healthy morphology [13], [39]. 244 In this context, a read-out that deserves special 245 mention is cell division (Fig. 3c, d): a well- 246 characterised biological process with easily 247 identifiable phases. It is highly regulated and 248 sensitive to various perturbations, including 249 illumination and changes in ROS concentrations 250 [15], [31]. This makes cell cycle an excellent read- 251 out for detection and quantification of phototoxicity 252 [39], with both continuous (Fig. 3c) and endpoint 253 (Fig. 3d) measurements possible. Delay in mitotic 254 progression has been used successfully to detect 255 perturbations in the health of both cultured cells and 256 developing embryos [32]–[35]. Additionally, 257 evaluating colony formation or number of cell 258 divisions after illumination (typically assessed after 259 a period of one or more cell cycles) can be indicative 260 of long-lasting damage [12], [29] (Fig. 3d). This 261 approach was used to perform extensive 262 characterisation of photodamage under illumination 263 conditions commonly used in single-molecule 264 localisation microscopy (SMLM) [10]. The viability of 265 several different cell lines was determined 20-24 h 266 post illumination, a strong correlation between 267 Page 5 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- shorter illumination wavelengths and increased cell 268 death was shown, particularly at high intensities. 269 However, results also suggested that long-term cell 270 viability is possible even with illumination 271 wavelengths as short as 405 nm, provided the 272 integrated light dose is small, preferably with 273 continuous rather than pulsed illumination. 274 Naturally, a limitation exists in employing these 275 methods to assess phototoxicity in post-mitotic 276 systems, e.g. primary neuron cultures. However, for 277 relevant models, choosing mitosis as a read-out has 278 the significant advantage of allowing phototoxicity 279 assessment based on label-free transmitted light 280 images [10], [29], [33], minimising the introducing 281 additional damage during evaluation. 282 From reports of phototoxicity in literature, several 283 conclusions can be drawn to guide live-cell friendly 284 SRM. Firstly, red-shifted wavelengths are preferable 285 to shorter wavelengths. In particular, UV 286 wavelengths should be avoided wherever possible 287 [10], [29], [33]. Furthermore, several studies 288 demonstrate that lower intensity illumination with 289 longer exposure is less damaging than short intense 290 bursts or pulses of illumination [10], [34], [40]. Most 291 importantly, a recurrent message throughout the 292 literature is that higher illumination intensities are 293 more damaging than corresponding imaging 294 conditions with lower illumination intensities. We 295 anticipate that real-time phototoxicity 296 measurements will become commonplace in both 297 diffraction-limited microscopy and SRM, and that 298 future SRM techniques will be accompanied by a 299 thorough description of how they impact living 300 Fig. 4 Low phototoxicity fluorescent probes and labelling for live-cell Super-Resolution Microscopy. Various recently-developed fluorescent protein- (a) and synthetic fluorophore- (b) based methods for labelling in live-cell super-resolution. All labels are shown attached to a microtubule as an example of an intracellular structure, with the exception of the Cer-HMSiR membrane dye in b. Page 6 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- samples. Concomitantly, for SRM users, awareness 301 of strategies for minimising phototoxicity is crucial. 302 Fluorescent probe development for live-cell 303 Super-Resolution Microscopy 304 SRM techniques have distinct requirements for 305 fluorescent probes. SIM quality relies on collecting 306 images of high Signal-to-Noise Ratio (SNR), 307 generally achieved by labelling with fluorophores of 308 high brightness and resistance to photobleaching. In 309 STED, fluorophores must not only be bright but also 310 possess a large Stokes-shift and stimulated 311 emission cross-section at the STED wavelength 312 [59]. SMLM techniques have the most demanding 313 labelling requirements - fluorophores must be 314 capable of cycling between ‘on’ and ‘off’ states with 315 appropriate kinetics, a high quantum yield in the on- 316 state, and a very low quantum yield in the off-state. 317 Several fluorophores and probes have been 318 developed specifically for SRM [60], [61]. However, 319 while many specialised fluorophores exist for fixed 320 specimens [62], there are far fewer options available 321 for live-cell imaging. An inappropriate choice of 322 fluorophore for live-cell SRM will not only lead to low 323 quality images downstream [63], but also inevitably 324 impact acquisition settings and hence phototoxicity 325 [10], [64]. 326 As for most fluorescence microscopy techniques, 327 the two classes of fluorophores used in SRM are 328 fluorescent proteins (FPs) (Fig. 4a) and synthetic 329 fluorophores (SFs) (Fig. 4b). FPs are the usual 330 choice for live-cell imaging as they can be fused to 331 a target of interest via genetic encoding, but at the 332 cost of reduced brightness compared to SFs. The 333 recent development of bright and photobleaching- 334 resistant FPs has expanded the options for SIM and 335 STED (Fig. 4a, left). Examples of these new FPs are 336 mNeonGreen (λex=506 nm) [65], mScarlet 337 (λex=569 nm) [66] and mGarnet (λex=598 nm) [67]. 338 SMLM techniques generally require 339 photoswitchable fluorophores (e.g. mEos3.2, 340 rsKame) [68], [69]. Despite the availability of several 341 photoswitchable FPs, their use in live-cell imaging 342 remains challenging [10], [64]. The chief reason is 343 that transitions between off- and on-states are 344 typically modulated by UV illumination. The 345 combination of this with high intensity excitation for 346 detection of molecular positions results in a short 347 window for live-cell SMLM studies. To reduce 348 phototoxicity in SMLM, FPs that do not require UV 349 pumping for photoswitching are being developed 350 (Fig. 4a, centre), with one such example being 351 SPOON [70]. Primed conversion is another 352 promising UV-independent approach to induce 353 photoswitching (Fig. 4a, right) [71]. Thereby a 354 combination of blue and near-infrared illumination 355 induces photoconversion in Dendra2 and the newly 356 developed primed-conversion protein pr-mEos2 357 [71], [72]. Recently, a general mechanism for primed 358 conversion was described, which is anticipated to 359 accelerate the development of more FPs that can be 360 photoconverted with this live-cell friendly approach 361 [73]. FPs for other specific SRM techniques have 362 also been developed (e.g. Skylan-NS for non-linear 363 SIM or GMars for REversible Saturable/switchable 364 OpticaL Fluorescence Transitions, RESOLFT) [74], 365 [75]. 366 Page 7 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- The second alternative, SFs (Fig. 4b), are small 367 chemically synthesised probes. These have higher 368 quantum yields and are more robust against 369 photobleaching than FPs [76]–[79]. While there are 370 some cell-permeable SFs that can be used to label 371 specific proteins (e.g. fluorogens such as SiR- 372 tubulin and SiR-actin) (Fig. 4b, left) [80], [81] or cell 373 compartments directly (e.g. Membright, ER-Tracker 374 or MitoTracker) (Fig. 4b, centre) [77], [82]–[84], 375 additional `linker' molecules are normally required to 376 associate SFs with the structure of interest. These 377 linkers must bind the target structure with high 378 affinity and specificity (e.g. antibodies and 379 DNA/RNA scaffolds, usually using amine- or thiol- 380 reactive derivatives of the SF) [85]. However, many 381 of these high-affinity linkers and SFs are not cell- 382 permeable, which limits their use in live-cell SRM to 383 labelling of cell-surface molecules. If genetic 384 encoding is possible and preferable, cell-permeable 385 SFs can be combined with flexible self-labelling 386 systems, such as SNAP-tag, Halo-tag or FlAsH (Fig. 387 4b, right) [86]–[89]. An elegant example of such an 388 approach is the use of Cox8A-SNAP fusion labelled 389 with SNAP-Cell SiR for STED. This has enabled the 390 visualisation of the dynamics of mitochondrial 391 cristae with ~70 nm resolution [90]. 392 SFs have also been engineered for live-cell SRM. 393 Spontaneously blinking synthetic fluorophores (e.g. 394 HMSiR) have been recently developed (Fig. 4b, 395 center). They do not require UV irradiation or 396 cytotoxic additives (such as thiol) to induce 397 photoswitching [91], [92]. High photostability SFs 398 have also been developed, enabling live-cell STED 399 [79], [93]–[95]. 400 A final regime for live-cell SRM-compatible labelling 401 is based on site-specific conjugation of fluorophores 402 to a target of interest, through genetic code 403 modifications and click chemistry (Fig. 4b, right) 404 [96]–[98]. These approaches combine the benefits 405 of site-specific labelling (as is the case for FPs) with 406 no requirement for protein expression and bright 407 labels (as is the case for SFs). 408 409 Biological models and sample preparation 410 Care should be taken when selecting a biological 411 model for SRM. Cellular sensitivity to light exposure 412 can vary based on cell type and species [10], [14], 413 [45], and in the case of whole organisms, 414 developmental stage [13], [34]. Phototoxicity has 415 been documented for different cell types, ranging 416 from primary cells [13], [45] to various immortalised 417 cell lines [10], [26], [38], [99]. One such study 418 focuses on immortalised cell lines, where it shows 419 that COS-7 and U2OS cells exhibit similar 420 photosensitivity, whereas HeLa cells are 421 substantially more robust, potentially making the 422 latter a more suitable system for live-cell SRM 423 studies [10]. Another study illustrated the effect of 424 photodamage on primary cells from rat central 425 nervous system [45]. Here, illumination with blue 426 light could induce morphological changes, 427 differentiation or cell death depending on the cell 428 type. 429 When imaging whole organisms, earlier 430 developmental stages from the same species tend 431 to be more photosensitive than later [12]. 432 Page 8 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- Furthermore, different model organisms display 433 variable photosensitivity. For example, fruit fly 434 embryos and nematode worms have higher 435 illumination tolerances than zebrafish embryos, 436 corals or cultured cells [13], [14]. Even within the 437 same cell, different intracellular structures exhibit 438 different responses to illumination [29], [100]. 439 Photodamage can be mitigated through additional 440 sample preparation steps. Established strategies 441 centre on preventing photobleaching by modifying 442 the sample environment. As photobleaching can 443 contribute to phototoxicity via ROS production [44], 444 strategies to reduce photobleaching could also help 445 ameliorate phototoxicity [15], [29], [101]. One 446 strategy is to modify the environmental conditions 447 prior to or during imaging. A prime example is 448 removal of oxygen, the main effector of 449 photobleaching [102], from the culture medium. This 450 can be achieved by bubbling nitrogen through the 451 medium during imaging. This yields an increased 452 photostability [103], [104] and, since oxygen is 453 directly involved in the production of ROS, also 454 reduces light-dependent oxidative stress on the 455 sample. It has also been shown that growing cells in 456 a hypoxic environment (3% oxygen) yielded a 25% 457 increase in mitosis entry after blue light irradiation 458 [33]. Other approaches to reduce oxygen in the 459 medium involve the addition of commercially 460 available oxygen-scavengers such as the Oxyrase® 461 enzyme complex (developed by Oxyrase, Inc., 462 Mansfield, Ohio). In combination with suitable 463 substrates, such as D/L-lactate or D/L-succinate, 464 these enzymes catalytically reduce the 465 concentration of oxygen and free radicals present in 466 the medium, thus minimising photobleaching and 467 phototoxicity [105], [106]. While these approaches 468 could improve live-cell SRM, it should be noted that 469 they are only suitable for specimens which can 470 tolerate hypoxia or anoxia. Notably, some 471 fluorophores used in SRM require oxygen 472 scavenger systems to photoswitch, however, these 473 buffers typically use cytotoxic compounds such as 474 thiols, making them unsuitable for live-cell imaging. 475 A different strategy for reduction of ROS during 476 imaging involves supplementing the media with 477 antioxidants. Antioxidants are molecules that 478 prevent oxidation in a biological context [107]. 479 Among antioxidants, Trolox, the soluble form of 480 vitamin E, has been shown to have a protective 481 effect for a number of cell lines due to its ROS- 482 neutralising properties [108]. The presence of the 483 antioxidant in the sample medium has been shown 484 to increase the number of post-illumination mitotic 485 cells by up to 38% compared to cells illuminated 486 without Trolox [33]. However, this molecule is not 487 suitable for SMLM, as it has been shown to inhibit 488 fluorophore blinking [109]. Another antioxidant used 489 in microscopy is rutin, a plant flavonoid shown to 490 reduce EGFP reddening [110], [111], although no 491 direct reduction of phototoxicity was demonstrated. 492 A notable example of a medium additive for live-cell 493 imaging is the vitamin- and antioxidant-rich 494 'Supplements for Optogenetic Survival' (SOS). SOS 495 has been shown to increase viability and reduce 496 photodamage in several cell types of the rat central 497 nervous system [45]. 498 There are chemicals used in mounting media, such 499 as various antioxidants, triplet-state quenchers and 500 Page 9 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- radical scavengers, that can be used for 501 photobleaching reduction and ROS neutralisation. 502 These include ascorbic acid [112], n-propyl gallate 503 [112]–[114], p-phenylenediamine [114]–[116], 1,4- 504 diazobicyclo(2,2,2)-octane (DABCO) [114], [117], 505 mercaptoethylamine (MEA) and cyclooctatetraene 506 (COT) [112]. Their presence in mounting media for 507 reduction of photobleaching is well characterised 508 [112], [115], [118], however there is no 509 comprehensive study on the use of these chemicals 510 in live-cell imaging. As a result, there is little 511 information regarding biocompatible working 512 concentrations or biological side effects. Therefore, 513 while potentially useful, they require further 514 exploration prior to use in live-cell SRM. 515 Some substances commonly used as supplements 516 are known also to cause phototoxicity, such as 517 molecules with benzene rings which are intrinsically 518 fluorescent [111]. For example, common cell media 519 components, such as riboflavin and pyridoxal, can 520 enhance oxidative reddening of GFPs; this effect 521 accounts for a considerable part of GFP 522 photobleaching [119]. Depleting these substances 523 increases GFP photostability, indirectly reducing 524 photodamage [110]. Additionally, the combination of 525 riboflavin and tryptophan in media generates ROS 526 and induces cytotoxicity upon illumination, whereas 527 their removal alleviates this effect [120], [121]. 528 Finally, the study that established the SOS 529 supplement [45] used it in combination with the 530 photoinert media NEUMO and MEMO, which also 531 lack riboflavin. These media were specifically 532 developed to prevent phototoxicity of nervous 533 system cells. A confounding example is 4-(2- 534 hydroxyethyl)-1-piperazineethanesulfonic acid 535 (HEPES), commonly used as a replacement for 536 carbon dioxide buffering during imaging [39]. 537 However, early reports demonstrated that HEPES- 538 buffered media exposed to low-intensity white light 539 can generate toxic hydrogen peroxide with 540 detrimental effects on thymocyte or T-cell culture 541 [122], [123]. 542 There is still a lack of information on SRM sample 543 preparation reducing phototoxicity. Many principles 544 can be transferred from conventional fluorescence 545 imaging. These include assessing photosensitivity 546 of the biological model, environmental conditions, 547 and attention to media composition. 548 Hardware developments for improved live- 549 cell imaging 550 The microscope configuration has a substantial 551 impact on the amount of photodamage experienced 552 by a specimen. Fig. 5a shows the common 553 illumination regimes for conventional microscopy 554 and SRM (widefield for SIM and SMLM, confocal for 555 STED). Basic optimisations of the microscope body, 556 for example minimising photon loss in the detection 557 path by using high-quality filters and sensitive 558 detectors, will reduce the illumination burden to 559 achieve suitable SNR [101]. In SRM approaches, 560 microscopes are built with high-quality components, 561 often having bespoke solutions to maximize signal 562 detection [124], [125]. In addition, the ever-present 563 phototoxic high-intensity illumination requirements 564 of most SRM techniques can be further ameliorated 565 using dedicated hardware designs. Interestingly, a 566 recent study shown low-illumination live-cell SRM 567 Page 10 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- immediately followed by in situ fixation of the sample 568 and high-illumination SRM [126]. This approach 569 combines the collection of temporal information in 570 living-cells with a mild resolution increase, then 571 capture of higher resolution for a specific timepoint 572 upon fixation. 573 In the case of STED, the presence of a second high- 574 intensity laser beam (depletion laser) in addition to a 575 confocal excitation beam confers the high 576 phototoxicity of this method. However, the 577 properties of both beams can have a substantial 578 impact on sample photodamage. It has been shown 579 in confocal microscopy that nanosecond pulsed, 580 rather than continuous, excitation can reduce 581 photobleaching, and that averaging multiple fast 582 scans is less phototoxic than acquiring a single slow 583 scan (Fig. 5b, `Temporally adaptive illumination') 584 [41]. The properties of the excitation beam have also 585 been explored specifically in STED. For example, 586 reducing the pulsing rate of the excitation laser 587 allows time for long-lived triplet states to relax which 588 leads to decreased photobleaching [127]. Similarly 589 to confocal microscopy, scanning at a higher rate in 590 STED has been shown to reduce photobleaching 591 [42]; this is enabled by using fast resonant scanning 592 mirrors rather than slower galvanometer scanning 593 mirrors to scan the beam pair through the sample. 594 Another method described reducing phototoxicity in 595 STED is by using two-photon excitation (Fig. 5a 596 `Two-photon'). As two-photon excitation only excites 597 fluorophores within the focal volume of the beam 598 (rather than along the entire beam path, as is the 599 case in single-photon excitation), it is often 600 considered a more live-cell friendly imaging regime 601 [54], [128]. Indeed, live-cell STED has been 602 successfully demonstrated with two-photon 603 excitation [129], [130] although while the former 604 paper claims that there is no photodamage to the 605 sample, this is not quantified. It should be noted that 606 two-photon excitation does however increase local 607 heating, which can damage the sample in a non- 608 fluorophore mediated manner [131]. 609 Fig. 5 Hardware modalities for conventional and low-phototoxicity Super-Resolution Microscopy. a Microscopy illumination regimes for conventional fluorescence imaging. b Examples of regimes that reduce light dose to the sample by inhomogeneous illumination. c Examples of light-sheet microscopy geometries. Page 11 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- In STED microscopy with pulsed depletion lasers, 610 resolution scales non-linearly with beam intensity. 611 Thus, in order to obtain high resolution images, very 612 high (and phototoxic) depletion beam intensities are 613 required. A different approach to obtaining high 614 resolution STED images without this power 615 dependence is gSTED (gated-STED) [132]. gSTED 616 uses a continuous wave (CW) laser for the depletion 617 beam rather than a pulsed laser. When a CW 618 depletion beam is combined with a pulsed excitation 619 beam, spatial information about the underlying 620 fluorophore distribution becomes encoded in the 621 temporal information of emission on a nanosecond 622 timescale. By using time-gated detectors, photons 623 detected immediately after excitation can be 624 excluded from the final image, which improves 625 image resolution. By tuning the size of the time-gate, 626 gSTED can thus increase STED resolution 627 independent of increasing light dose to the sample 628 [133]. 629 SIM is generally considered the least phototoxic 630 SRM technique [134]. However, it still requires the 631 acquisition of several frames (often ≥ 9) at high SNR 632 in order to generate the final reconstructed image. 633 Several approaches have been developed to reduce 634 the number of frames required for a SIM 635 reconstruction, including pixel reassignment and 636 image scanning microscopy (ISM) methods. One 637 example is multifocal structured illumination 638 microscopy (MSIM, [135]), which combines 639 principles from SIM and confocal microscopy to 640 scan an array of spots across the sample for fast 641 live-cell imaging with resolution doubling (Fig. 5b, 642 `Multi-focal illumination'). Another method, rapid 643 non-linear ISM [136], combines ISM with two-photon 644 excitation and second-harmonic generation for low 645 phototoxicity imaging. A wide range of such SIM- 646 based techniques exist, and have been rigorously 647 compared elsewhere [134], [137]. It has been 648 demonstrated recently that using sub-millisecond 649 pulses as excitation in SIM (when combined with 650 novel analytics as described below) reduced 651 photobleaching and enables long-term live-cell 652 imaging [138]. 653 Techniques that restrict illumination to only the focal 654 plane of the sample are also preferable to those 655 which illuminate along the whole beam path. One 656 such example of this is TIRF (total internal reflection 657 fluorescence) microscopy, where only fluorophores 658 within a few hundred nanometers of the coverslip 659 are illuminated. While TIRF has been combined with 660 super-resolution modalities, such as SIM, and is 661 effective in reducing photodamage by axially 662 confining excitation [134], it is restrictive in that only 663 biological structures adjacent to the cell membrane 664 can be studied. 665 Light-sheet microscopy approaches similarly 666 confine illumination to a narrow band, but their 667 imaging geometries allow for investigation of 668 structures throughout the whole sample and not just 669 regions close to the coverslip. The majority of them 670 involve illuminating the sample with a thin sheet of 671 light and then detecting the fluorescence 672 perpendicular to the direction of sheet propagation 673 (Fig. 5c, `Gaussian light sheet') [139], [140]. This 674 confers low phototoxicity as only the part of the 675 sample being imaged is illuminated without the need 676 for non-linear optical processes (which is the case in 677 Page 12 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- two-photon microscopy). Indeed, light-sheet 678 microscopy was named the Nature Methods 679 technique of the year in 2014, in part due to its low 680 phototoxicity [141]. There are several ways in which 681 light-sheet microscopy schemes can yield super- 682 resolution with reduced phototoxicity. Super- 683 resolution in live samples has been demonstrated 684 using light-sheet microscopy by simply combining 685 this illumination geometry with SRM techniques 686 such as SMLM [142]–[144] and RESOLFT [145]. 687 However, the employed SRM methods still require 688 high-intensity illumination, and thus such composite 689 techniques do not exploit the inherent low 690 phototoxicity of light-sheet imaging. Therefore, a 691 more elegant approach involves illuminating the 692 sample with a light-sheet regime followed by the 693 application of SMLM analytics designed for ultra- 694 high-density datasets, which allows for reduction of 695 the illumination power ([146] and Analytics section, 696 see below). The more widely-explored method for 697 combining SRM and light-sheet microscopy has 698 been the use of novel methods for generating and 699 shaping the light-sheet. Bessel beams have been 700 used to generate thinner light-sheets [147], and 701 these beams have also been extended to 702 incorporate SIM [148]. The latter strategy has also 703 been demonstrated on a system with two 704 counterpropagating light-sheets formed using 705 standard Gaussian beams [149]. The most radical 706 and live-imaging-friendly light-sheet SRM technique 707 developed to date is lattice light-sheet microscopy 708 [150] (Fig. 5c, `Lattice light sheet'). This has 709 demonstrated 3D time-lapse super-resolution 710 imaging in both cultured cells and intact model 711 organisms with minimal phototoxicity. 712 An interesting approach to reducing the illumination 713 dose in SRM is using spatially varying illumination 714 depending on the structural content of the imaging 715 region (Fig. 5b, `Spatially adaptive illumination'). 716 This approach was originally demonstrated for 717 Fig. 6 Analytics to complement low-phototoxicity imaging regimes. a Top: typical SMLM images are successfully reconstructed from sparse blinking raw data acquired under high phototoxic illumination. Bottom: reducing phototoxic illumination leads to more emitting fluorophores per raw data frame. When reconstructed using conventional SMLM algorithms, these produce low-quality images containing artefacts. High density SMLM algorithms can produce better quality images from such datasets. b Top: typical SIM imaging involves acquiring 9-25 raw images (depending on the number of grating rotations and phases) at high SNR, which can be successfully reconstructed using conventional SIM algorithms. Bottom: decreasing the illumination intensity, and thus SNR of the raw images, leads to artefacts in images reconstructed using conventional methods. The Hessian SIM deconvolution algorithm can bypass this limitation [138]. c Deep neural networks can be trained to infer super-resolution information from e.g. low-resolution diffraction-limited or low-quality super-resolution images. In this example, a neural network can be trained on pairs of low resolution/super-resolution images of the trained structure (`Network training'). The trained network can then be applied to unseen low resolution images to infer the super-resolution equivalents (`Network inference'). Page 13 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- confocal imaging [48] and has since been extended 718 to SIM [151], RESOLFT [152] and indeed light-sheet 719 microscopy [141]. There is also a range of adaptive 720 illumination STED techniques that have been 721 developed [153]–[155], and while these 722 predominantly focus on reducing light dose in the 723 context of photobleaching, this will concomitantly 724 also impact the live-cell compatibility of these 725 techniques. 726 Analytical approaches to live-cell Super- 727 Resolution Microscopy 728 Analytics can be used to extract super-resolution 729 information from images acquired at low 730 illumination, and thus low phototoxicity (Fig. 6). Such 731 techniques are generally based on SMLM principles 732 but improve its live-cell compatibility (Fig. 6a). In 733 SMLM, when high intensity illumination is used, 734 fluorophore blinking is sparse and thus the well- 735 separated single molecules are straightforward to 736 detect and localise with high accuracy and precision 737 [156], [157]. However, as intensity is decreased 738 towards a lower phototoxicity regime, blinking 739 becomes more dense and molecules become 740 increasingly overlapped. Such datasets require 741 specialised algorithms to extract molecule locations. 742 The first example of such an algorithm was Super- 743 Resolution Optical Fluctuation imaging (SOFI), 744 where the temporal statistics of fluorophore intensity 745 oscillations are used to generate images with sub- 746 diffraction resolution [158]. Indeed, SOFI has been 747 used to image live cells [159] although only for short 748 periods of time due to the requirement for UV 749 illumination to induce photoswitching. Another 750 algorithm developed for analysing datasets with 751 dense blinking is 3B [160], where super-resolution 752 images can be obtained from datasets imaged with 753 a xenon arc lamp rather than lasers. However, both 754 SOFI and 3B techniques still rely on 755 photoswitchable fluorophores, which have 756 drawbacks discussed above. The Super-Resolution 757 Radial Fluctuations (SRRF) algorithm allows for the 758 reconstruction of super-resolution images from 759 datasets containing non-photoswitchable 760 fluorophores such as GFP [161], [162]. SRRF has 761 been shown to work on datasets obtained with 762 confocal and LED-illuminated microscopes, with the 763 latter enabling continuous live-cell imaging for >30 764 minutes [163]. However, SRRF cannot retrieve 765 resolutions in these regimes as high as those 766 achievable with photoswitchable fluorophores. A 767 promising new development for analysing high- 768 density datasets is Haar wavelet kernel (HAWK) 769 [164]. HAWK is a pre-processing algorithm that 770 separates fluorophores in time; this creates an 771 artificial lower-density dataset, which can then be 772 analysed using any SMLM algorithm. 773 While most analytical developments for live-cell 774 SRM centre on SMLM-based techniques, there are 775 also analytics for enabling lower phototoxicity 776 imaging in SIM and STED. Hessian-SIM is a 777 deconvolution algorithm that can obtain high-quality 778 SIM images from raw data acquired at low signal-to- 779 noise ratio (Fig. 6b) [138]. This overcomes a 780 substantial barrier in SIM, in that conventional SIM 781 reconstruction algorithms perform poorly on low- 782 illumination datasets, leading to artefacts within the 783 resulting images. Approaches have also been 784 Page 14 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- proposed for low-power STED microscopy based on 785 reconstructing images with knowledge of 786 fluorescence lifetime changes induced by the STED 787 beam [75], [165]. 788 A rapidly evolving field in microscopy image analysis 789 is the use of machine learning (ML)-based 790 techniques [166], [167]. Such techniques are used 791 for diverse applications including object 792 segmentation, denoising, and structure prediction, 793 and these can also be extended to SRM (Fig. 6c). 794 One example is Content Aware Image Restoration 795 (CARE), where a neural network is trained on high 796 illumination intensity datasets (i.e. high 797 phototoxicity) and used to denoise corresponding 798 datasets acquired at much lower illumination 799 intensities [168]. CARE was demonstrated to 800 enhance resolution of GFP-tagged microtubules to 801 a similar extent to SRRF analysis of the same data, 802 but with higher quality and higher temporal 803 resolution. There are also specialised ML algorithms 804 for super-resolution applications. ANNA-PALM is a 805 method that, after training a neural network with 806 sparse SMLM data, can reconstruct super- 807 resolution images from dense data and a 808 correspondingly lower number of frames [169]. 809 While not demonstrated in live-cell data, this 810 technique could in theory alleviate phototoxicity with 811 minimal sacrifice to spatial resolution by imaging 812 photoswitchable FPs with lower illumination 813 intensity. Other ML-based techniques have also 814 allowed for prediction of enhanced resolution 815 images from low illumination diffraction-limited 816 images (Fig. 6c), for example: converting confocal 817 to Airyscan-type or STED-type images [75], [170]; or 818 widefield to SIM-type images [75]. 819 Discussion and outlook 820 High quality live-cell fluorescence microscopy 821 involves compromising between four key properties: 822 SNR, imaging speed, spatial resolution, and sample 823 health [12]. We present an overview of the 824 challenges faced on how to balance the latter two 825 properties in live-cell SRM, highlighting potential 826 strategies to maximise resolution while minimising 827 phototoxicity. 828 As commercial super-resolution systems become 829 commonplace in biological labs and open-source 830 microscope hardware becomes more widespread, 831 there is a growing desire to translate cell biology 832 experiments from conventional diffraction-limited 833 microscopes to higher resolution alternatives. 834 However, the cost of this increased resolution is 835 often the sample health. Users must be aware of 836 what phototoxicity is, how to detect it, and methods 837 that can be used to ameliorate it. Unfortunately, 838 there are very few dedicated studies discussing 839 phototoxicity specifically in SRM [10], [29]. 840 It is clear that there are several frontiers for 841 optimising SRM protocols for minimising 842 phototoxicity, and a much-needed development in 843 the field is a non-perturbing robust indicator of 844 sample health during imaging. Caution must be 845 taken when reporting and evaluating phototoxicity 846 as it would also require using uniform metrics for 847 data quality. There is already software available for 848 assessing the quality and resolution of SRM images 849 [171], [172] Comparative analytics for phototoxicity 850 Page 15 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- would thus provide a complete numerical framework 851 for experiment optimisation. 852 As super-resolution microscopes become 853 increasingly standard equipment in biological 854 research, users must be aware of their limitations in 855 live-cell imaging. Many of the suggestions offered in 856 this review for reducing phototoxicity remain under 857 active development, and it is imperative for users to 858 follow progress in hardware, analytics and 859 fluorophores to ensure that they are minimising 860 photodamage to samples. 861 Acknowledgements 862 This work was funded by grants from the UK 863 Biotechnology and Biological Sciences Research 864 Council (BB/R000697/1; BB/R021805/1; 865 BB/S507532/1) (R.H.), the UK Medical Research 866 Council (MR/K015826/1) (R.H.), the Wellcome Trust 867 (203276/Z/16/Z) (S.C., R.H). K.L.T. supported by a 868 4-year MRC Research Studentship and Y.Y. by a 4- 869 year BBSRC Research Studentship. We would like 870 to thank Dr. David Albrecht (Max Planck Institute for 871 the Science of Light, Erlangen, Germany), Dr. 872 Francesca Bottanelli (Freie Universität Berlin, Berlin, 873 Germnay), Dr. Agathe Chaigne (UCL, London, UK), 874 Dr. Gautam Dey (UCL, London, UK), Dr. Caron 875 Jacobs (University of Cape Town, Cape Town, 876 South Africa), Ms. Megan Jones (UCL, London, UK), 877 Dr. Christophe Leterrier (Aix Marseille Université, 878 Marseille, France), Dr. Apostolos Papandreou (UCL, 879 London, UK) and Dr. Uwe Schmidt (Max Planck 880 Institute of Molecular Cell Biology and Genetics, 881 Dresden, Germany) for valuable advice and 882 feedback. 883 884 ORCID IDs 885 Kalina L. Tosheva: 886 https://orcid.org/0000-0003-0999-5182 887 Yue Yuan: 888 https://orcid.org/0000-0002-6698-3009 889 Pedro Matos Pereira: 890 https://orcid.org/0000-0002-1426-9540 891 Siân Culley: 892 https://orcid.org/0000-0003-2112-0143 893 Ricardo Henriques: 894 https://orcid.org/0000-0002-2043-5234 895 Page 16 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- Bibliography 896 [1] L. Schermelleh et al., “Super-resolution microscopy demystified,” Nat. Cell Biol., vol. 21, no. 1, pp. 72–84, 897 2019. 898 [2] E. Betzig et al., “Imaging Intracellular Fluorescent Proteins at Nanometer Resolution,” Science (80-. )., 899 vol. 313, no. 5793, pp. 1642–1645, 2006. 900 [3] M. J. Rust, M. Bates, and X. Zhuang, “Sub-diffraction-limit imaging by stochastic optical reconstruction 901 microscopy (STORM),” Nat. Methods, vol. 3, no. 10, pp. 793–796, 2006. 902 [4] M. G. L. Gustafsson, “Surpassing the lateral resolution limit by a factor of two using structured 903 illumination microscopy,” J. Microsc., vol. 198, no. 2, pp. 82–87, May 2000. 904 [5] S. W. Hell and J. Wichmann, “Breaking the diffraction resolution limit by stimulated emission: stimulated- 905 emission-depletion fluorescence microscopy,” Opt. Lett., vol. 19, no. 11, pp. 780–782, Jun. 1994. 906 [6] K. Xu, G. Zhong, and X. Zhuang, “Actin, Spectrin, and Associated Proteins Form a Periodic Cytoskeletal 907 Structure in Axons,” Science (80-. )., vol. 339, no. 6118, pp. 452 LP – 456, Jan. 2013. 908 [7] J. Nixon-Abell et al., “Increased spatiotemporal resolution reveals highly dynamic dense tubular matrices 909 in the peripheral ER,” Science (80-. )., vol. 354, no. 6311, p. aaf3928, Oct. 2016. 910 [8] J. Chojnacki et al., “Envelope glycoprotein mobility on HIV-1 particles depends on the virus maturation 911 state,” Nat. Commun., vol. 8, no. 1, p. 545, 2017. 912 [9] T. Nerreter et al., “Super-resolution microscopy reveals ultra-low CD19 expression on myeloma cells that 913 triggers elimination by CD19 CAR-T,” Nat. Commun., vol. 10, no. 1, p. 3137, 2019. 914 [10] S. Wäldchen, J. Lehmann, T. Klein, S. van de Linde, and M. Sauer, “Light-induced cell damage in live- 915 cell super-resolution microscopy,” Sci. Rep., vol. 5, p. 15348, Oct. 2015. 916 [11] R. Henriques, C. Griffiths, E. Hesper Rego, and M. M. Mhlanga, “PALM and STORM: Unlocking live-cell 917 super-resolution,” Biopolymers, vol. 95, no. 5, pp. 322–331, May 2011. 918 [12] P. P. Laissue, R. A. Alghamdi, P. Tomancak, E. G. Reynaud, and H. Shroff, “Assessing phototoxicity in 919 live fluorescence imaging,” Nat. Methods, vol. 14, p. 657, Jun. 2017. 920 [13] J. Icha, M. Weber, J. C. Waters, and C. Norden, “Phototoxicity in live fluorescence microscopy, and how 921 to avoid it,” BioEssays, vol. 39, no. 8, p. 1700003, Aug. 2017. 922 [14] V. Magidson and A. Khodjakov, Circumventing photodamage in live-cell microscopy, 4th ed., vol. 114. 923 Elsevier Inc., 2013. 924 [15] R. Dixit and R. Cyr, “Cell damage and reactive oxygen species production induced by fluorescence 925 microscopy: effect on mitosis and guidelines for non-invasive fluorescence microscopy,” Plant J., vol. 36, 926 no. 2, pp. 280–290, 2003. 927 [16] K. Logg, K. Bodvard, A. Blomberg, and M. Käll, “Investigations on light-induced stress in fluorescence 928 microscopy using nuclear localization of the transcription factor Msn2p as a reporter,” FEMS Yeast Res., 929 Page 17 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- vol. 9, no. 6, pp. 875–884, 2009. 930 [17] Y. Devary, C. Rosette, J. A. DiDonato, and M. Karin, “NF-kappa B activation by ultraviolet light not 931 dependent on a nuclear signal,” Science (80-. )., vol. 261, no. 5127, pp. 1442–1445, 1993. 932 [18] A. Rehemtulla, C. A. Hamilton, A. M. Chinnaiyan, and V. M. Dixit, “Ultraviolet Radiation-induced 933 Apoptosis Is Mediated by Activation of CD-95 (Fas/APO-1),” J. Biol. Chem. , vol. 272, no. 41, pp. 25783– 934 25786, Oct. 1997. 935 [19] M. J. Peak, J. G. Peak, and B. A. Carnes, “Induction of Direct and Indirect Single‐Strand Breaks in 936 Human Cell Dna By Far‐ and Near‐Ultraviolet Radiations: Action Spectrum and Mechanisms,” 937 Photochem. Photobiol., vol. 45, no. 3, pp. 381–387, 1987. 938 [20] M. J. Peak and J. G. Peak, “SINGLE-STRAND BREAKS INDUCED IN BACILLUS SUBTILIS DNA BY 939 ULTRAVIOLET LIGHT: ACTION SPECTRUM and PROPERTIES,” Photochem. Photobiol., vol. 35, no. 5, 940 pp. 675–680, May 1982. 941 [21] B. Durbeej and L. A. Eriksson, “Reaction mechanism of thymine dimer formation in DNA induced by UV 942 light,” J. Photochem. Photobiol. A Chem., vol. 152, no. 1, pp. 95–101, 2002. 943 [22] D. E. Brash et al., “A role for sunlight in skin cancer: UV-induced p53 mutations in squamous cell 944 carcinoma.,” Proc. Natl. Acad. Sci., vol. 88, no. 22, pp. 10124–10128, 1991. 945 [23] H. Vrieling et al., “DNA strand specificity for UV-induced mutations in mammalian cells.,” Mol. Cell. Biol., 946 vol. 9, no. 3, pp. 1277–1283, 1989. 947 [24] M. L. Cunningham, J. S. Johnson, S. M. Giovanazzi, and M. J. Peak, “Photosensitized Production of 948 Superoxide Anion by Monochromatic (290–405 nm) Ultraviolet Irradiation of NADH and NADPH 949 Coenzymes,” Photochem. Photobiol., vol. 42, no. 2, pp. 125–128, 1985. 950 [25] M. Eichler, R. Lavi, A. Shainberg, and R. Lubart, “Flavins are source of visible-light-induced free radical 951 formation in cells,” Lasers Surg. Med., vol. 37, no. 4, pp. 314–319, Oct. 2005. 952 [26] P. E. Hockberger et al., “Activation of flavin-containing oxidases underlies light-induced production of 953 H2O2 in mammalian cells,” Proc. Natl. Acad. Sci., vol. 96, no. 11, pp. 6255–6260, 1999. 954 [27] G. Y. Fraikin, M. G. Strakhovskaya, and A. B. Rubin, “The role of membrane-bound porphyrin-type 955 compound as endogenous sensitizer in photodynamic damage to yeast plasma membranes,” J. 956 Photochem. Photobiol. B Biol., vol. 34, no. 2–3, pp. 129–135, 1996. 957 [28] F. Ricchelli, “Photophysical properties of porphyrins in biological membranes,” J. Photochem. Photobiol. 958 B Biol., vol. 29, no. 2, pp. 109–118, 1995. 959 [29] N. Kilian et al., “Assessing photodamage in live-cell STED microscopy,” Nat. Methods, vol. 15, no. 10, pp. 960 755–756, 2018. 961 [30] C. S. Foote, “DEFINITION OF TYPE I and TYPE II PHOTOSENSITIZED OXIDATION,” Photochem. 962 Photobiol., vol. 54, no. 5, p. 659, Nov. 1991. 963 Page 18 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- [31] W. C. Burhans and N. H. Heintz, “The cell cycle is a redox cycle: Linking phase-specific targets to cell 964 fate,” Free Radic. Biol. Med., vol. 47, no. 9, pp. 1282–1293, 2009. 965 [32] Z. Schilling et al., “Predictive-focus illumination for reducing photodamage in live-cell microscopy,” J. 966 Microsc., vol. 246, no. 2, pp. 160–167, May 2012. 967 [33] S. Douthwright and G. Sluder, “Live Cell Imaging: Assessing the Phototoxicity of 488 and 546 nm Light 968 and Methods to Alleviate it,” J. Cell. Physiol., vol. 232, no. 9, pp. 2461–2468, 2017. 969 [34] J.-Y. Tinevez et al., “Chapter fifteen - A Quantitative Method for Measuring Phototoxicity of a Live Cell 970 Imaging Microscope,” in Imaging and Spectroscopic Analysis of Living Cells, vol. 506, P. M. B. T.-M. in E. 971 Conn, Ed. Academic Press, 2012, pp. 291–309. 972 [35] P. M. Carlton et al., “Fast live simultaneous multiwavelength four-dimensional optical microscopy,” Proc. 973 Natl. Acad. Sci., vol. 107, no. 37, pp. 16016–16022, 2010. 974 [36] M. M. Knight, S. R. Roberts, D. A. Lee, and D. L. Bader, “Live cell imaging using confocal microscopy 975 induces intracellular calcium transients and cell death,” Am. J. Physiol. Physiol., vol. 284, no. 4, pp. 976 C1083–C1089, Apr. 2003. 977 [37] Y. Wu et al., “Inverted selective plane illumination microscopy (iSPIM) enables coupled cell identity 978 lineaging and neurodevelopmental imaging in Caenorhabditis elegans,” Proc. Natl. Acad. Sci., vol. 108, 979 no. 43, pp. 17708–17713, 2011. 980 [38] H. Schneckenburger et al., “Light exposure and cell viability in fluorescence microscopy,” J. Microsc., vol. 981 245, no. 3, pp. 311–318, 2012. 982 [39] R. Cole, “Live-cell imaging,” Cell Adh. Migr., vol. 8, no. 5, pp. 452–459, Sep. 2014. 983 [40] F. Mubaid and C. M. Brown, “Less is More: Longer Exposure Times with Low Light Intensity is Less 984 Photo-Toxic,” Micros. Today, vol. 25, no. 6, pp. 26–35, 2017. 985 [41] C. Boudreau, T. L. Wee, Y. R. Duh, M. P. Couto, K. H. Ardakani, and C. M. Brown, “Excitation light dose 986 engineering to reduce photo-bleaching and photo-toxicity,” Sci. Rep., vol. 6, pp. 1–12, 2016. 987 [42] Y. Wu, X. Wu, R. Lu, J. Zhang, L. Toro, and E. Stefani, “Resonant Scanning with Large Field of View 988 Reduces Photobleaching and Enhances Fluorescence Yield in STED Microscopy,” Sci. Rep., vol. 5, p. 989 14766, Oct. 2015. 990 [43] E. A. Halabi, D. Pinotsi, and P. Rivera-Fuentes, “Photoregulated fluxional fluorophores for live-cell super- 991 resolution microscopy with no apparent photobleaching,” Nat. Commun., vol. 10, no. 1, p. 1232, 2019. 992 [44] K. Jacobson, Z. Rajfur, E. Vitriol, and K. Hahn, “Chromophore-assisted laser inactivation in cell biology,” 993 Trends Cell Biol., vol. 18, no. 9, pp. 443–450, 2008. 994 [45] J. H. Stockley et al., “Surpassing light-induced cell damage in vitro with novel cell culture media,” Sci. 995 Rep., vol. 7, no. 1, p. 849, 2017. 996 [46] J. W. Dobrucki, D. Feret, and A. Noatynska, “Scattering of exciting light by live cells in fluorescence 997 Page 19 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- confocal imaging: Phototoxic effects and relevance for FRAP studies,” Biophys. J., vol. 93, no. 5, pp. 998 1778–1786, 2007. 999 [47] C. Roechlecke, A. Schaller, L. Knels, and R. H. W. Funk, “The Influence of Sublethal Blue Light Exposure 1000 on Human {{RPE}} Cells,” Mol. Vis., vol. 15, pp. 1929–1938, 2009. 1001 [48] R. A. Hoebe, C. H. Van Oven, T. W. J. Gadella, P. B. Dhonukshe, C. J. F. Van Noorden, and E. M. M. 1002 Manders, “Controlled light-exposure microscopy reduces photobleaching and phototoxicity in 1003 fluorescence live-cell imaging,” Nat. Biotechnol., vol. 25, no. 2, pp. 249–253, 2007. 1004 [49] J. Ge et al., “Standard fluorescent imaging of live cells is highly genotoxic,” Cytom. Part A, vol. 83A, no. 1005 6, pp. 552–560, Jun. 2013. 1006 [50] D. H. Hawkins and H. Abrahamse, “The role of laser fluence in cell viability, proliferation, and membrane 1007 integrity of wounded human skin fibroblasts following helium-neon laser irradiation,” Lasers Surg. Med., 1008 vol. 38, no. 1, pp. 74–83, Jan. 2006. 1009 [51] Y. Kuse, K. Ogawa, K. Tsuruma, M. Shimazawa, and H. Hara, “Damage of photoreceptor-derived cells in 1010 culture induced by light emitting diode-derived blue light,” Sci. Rep., vol. 4, pp. 1–12, 2014. 1011 [52] A. McDonald, J. Harris, D. MacMillan, J. Dempster, and G. McConnell, “Light-induced Ca^2+ transients 1012 observed in widefield epi-fluorescence microscopy of excitable cells,” Biomed. Opt. Express, vol. 3, no. 6, 1013 p. 1266, 2012. 1014 [53] H. J. Koester, D. Baur, R. Uhl, and S. W. Hell, “Ca2+ fluorescence imaging with pico- and femtosecond 1015 two-photon excitation: Signal and photodamage,” Biophys. J., vol. 77, no. 4, pp. 2226–2236, 1999. 1016 [54] A. Hopt and E. Neher, “Highly Nonlinear Photodamage in Two-Photon Fluorescence Microscopy,” 1017 Biophys. J., vol. 80, no. 4, pp. 2029–2036, 2001. 1018 [55] C.-H. Chuang, A. E. Carpenter, B. Fuchsova, T. Johnson, P. de Lanerolle, and A. S. Belmont, “Long- 1019 Range Directional Movement of an Interphase Chromosome Site,” Curr. Biol., vol. 16, no. 8, pp. 825– 1020 831, 2006. 1021 [56] T. Minamikawa, D. A. Williams, D. N. Bowser, and P. Nagley, “Mitochondrial Permeability Transition and 1022 Swelling Can Occur Reversibly without Inducing Cell Death in Intact Human Cells,” Exp. Cell Res., vol. 1023 246, no. 1, pp. 26–37, 1999. 1024 [57] R. A. Hoebe, H. T. M. van der Voort, J. Stap, C. J. F. van Noorden, and E. M. M. Mamders, “Quantitative 1025 determination of the reduction of phototoxicity and photobleaching by controlled light exposure 1026 microscopy,” J. Microsc., vol. 231, no. 1, pp. 9–20, Jul. 2008. 1027 [58] D. Richmond, A. P.-T. Jost, T. Lambert, J. Waters, and H. Elliott, “DeadNet: Identifying Phototoxicity from 1028 Label-free Microscopy Images of Cells using Deep ConvNets,” CoRR, vol. abs/1701.0, 2017. 1029 [59] G. Vicidomini, P. Bianchini, and A. Diaspro, “STED super-resolved microscopy,” Nat. Methods, vol. 15, 1030 no. 3, pp. 173–182, 2018. 1031 Page 20 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- [60] R. Henriques and M. M. Mhlanga, “PALM and STORM: What hides beyond the Rayleigh limit?,” 1032 Biotechnol. J., vol. 4, no. 6, pp. 846–857, Jun. 2009. 1033 [61] J. C. Chozinski, Tyler J., Gagnon, Lauren A., Vaughan, “Twinkle, twinkle little star: photoswitchable 1034 fluorophores for super-resolution imaging,” FEBS Lett., vol. 588, p. 3603, 2014. 1035 [62] G. T. Dempsey, J. C. Vaughan, K. H. Chen, M. Bates, and X. Zhuang, “Evaluation of fluorophores for 1036 optimal performance in localization-based super-resolution imaging,” Nat. Methods, vol. 8, no. 12, pp. 1037 1027–1040, 2011. 1038 [63] J. V. Thevathasan et al., “Nuclear pores as versatile reference standards for quantitative superresolution 1039 microscopy,” bioRxiv, 2019. 1040 [64] S. van de Linde, I. Krstić, T. Prisner, S. Doose, M. Heilemann, and M. Sauer, “Photoinduced formation of 1041 reversible dye radicals and their impact on super-resolution imaging.,” Photochem. Photobiol. Sci., vol. 1042 10, no. 4, pp. 499–506, 2011. 1043 [65] N. C. Shaner et al., “A bright monomeric green fluorescent protein derived from Branchiostoma 1044 lanceolatum,” Nat. Methods, vol. 10, no. 5, pp. 407–409, 2013. 1045 [66] D. S. Bindels et al., “mScarlet: a bright monomeric red fluorescent protein for cellular imaging,” Nat. 1046 Methods, vol. 14, no. 1, pp. 53–56, 2017. 1047 [67] A. Hense, B. Prunsche, P. Gao, Y. Ishitsuka, K. Nienhaus, and G. U. Nienhaus, “Monomeric Garnet, a 1048 far-red fluorescent protein for live-cell STED imaging,” Sci. Rep., vol. 5, pp. 1–10, 2015. 1049 [68] M. Zhang et al., “Rational design of true monomeric and bright photoactivatable fluorescent proteins,” 1050 Nat. Methods, vol. 9, no. 7, pp. 727–729, 2012. 1051 [69] A. B. Rosenbloom, S.-H. S.-H. Lee, M. To, A. Lee, J. Y. Shin, and C. Bustamante, “Optimized two-color 1052 super resolution imaging of Drp1 during mitochondrial fission with a slow-switching Dronpa variant,” Proc. 1053 Natl. Acad. Sci., vol. 111, no. 36, pp. 13093–13098, 2014. 1054 [70] Y. Arai et al., “Spontaneously Blinking Fluorescent Protein for Simple Single Laser Super-Resolution Live 1055 Cell Imaging,” ACS Chem. Biol., p. acschembio.8b00200, 2018. 1056 [71] W. P. Dempsey et al., “In vivo single-cell labeling by confined primed conversion,” Nat. Methods, vol. 12, 1057 no. 7, pp. 645–648, 2015. 1058 [72] M. A. Mohr and P. Pantazis, “Primed Conversion: The New Kid on the Block for Photoconversion,” 1059 Chem. – A Eur. J., vol. 24, no. 33, pp. 8268–8274, Jun. 2018. 1060 [73] B. Turkowyd et al., “A General Mechanism of Photoconversion of Green-to-Red Fluorescent Proteins 1061 Based on Blue and Infrared Light Reduces Phototoxicity in Live-Cell Single-Molecule Imaging,” Angew. 1062 Chemie - Int. Ed., vol. 56, no. 38, pp. 11634–11639, 2017. 1063 [74] X. Zhang et al., “Highly photostable, reversibly photoswitchable fluorescent protein with high contrast 1064 ratio for live-cell superresolution microscopy,” Proc. Natl. Acad. Sci., vol. 113, no. 37, pp. 10364–10369, 1065 Page 21 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- 2016. 1066 [75] S. Wang et al., “GMars-T Enabling Multimodal Subdiffraction Structural and Functional Fluorescence 1067 Imaging in Live Cells,” Anal. Chem., vol. 90, pp. 6626–6634, 2018. 1068 [76] M. V Sednev, C. A. Wurm, V. N. Belov, and S. W. Hell, “Carborhodol: A New Hybrid Fluorophore 1069 Obtained by Combination of Fluorescein and Carbopyronine Dye Cores,” Bioconjug. Chem., vol. 24, no. 1070 4, pp. 690–700, Apr. 2013. 1071 [77] J. B. Grimm et al., “A general method to fine-tune fluorophores for live-cell and in vivo imaging,” Nat. 1072 Methods, vol. 14, p. 987, Sep. 2017. 1073 [78] P. Mateos-Gil, S. Letschert, S. Doose, and M. Sauer, “Super-Resolution Imaging of Plasma Membrane 1074 Proteins with Click Chemistry,” Front. Cell Dev. Biol., vol. 4, no. September, p. 98, 2016. 1075 [79] A. D. Thompson et al., “Long-Term Live-Cell STED Nanoscopy of Primary and Cultured Cells with the 1076 Plasma Membrane HIDE Probe DiI-SiR.,” Angew. Chem. Int. Ed. Engl., vol. 56, no. 35, pp. 10408– 1077 10412, Aug. 2017. 1078 [80] G. Lukinavičius et al., “Fluorogenic probes for live-cell imaging of the cytoskeleton,” Nat. Methods, vol. 1079 11, no. 7, pp. 731–733, 2014. 1080 [81] E. Kozma and P. Kele, “Fluorogenic probes for super-resolution microscopy,” Org. Biomol. Chem., vol. 1081 17, no. 2, pp. 215–233, 2019. 1082 [82] S.-H. Shim et al., “Super-resolution fluorescence imaging of organelles in live cells with photoswitchable 1083 membrane probes,” Proc. Natl. Acad. Sci., vol. 109, no. 35, pp. 13978–13983, 2012. 1084 [83] M. Collot et al., “MemBright: A Family of Fluorescent Membrane Probes for Advanced Cellular Imaging 1085 and Neuroscience,” Cell Chem. Biol., vol. 26, no. 4, pp. 600-614.e7, 2019. 1086 [84] L. Wang, M. S. Frei, A. Salim, and K. Johnsson, “Small-Molecule Fluorescent Probes for Live-Cell Super- 1087 Resolution Microscopy,” J. Am. Chem. Soc., vol. 141, no. 7, pp. 2770–2781, Feb. 2019. 1088 [85] L. D. Lavis, “Teaching Old Dyes New Tricks: Biological Probes Built from Fluoresceins and Rhodamines,” 1089 Annu. Rev. Biochem., vol. 86, no. 1, pp. 825–843, 2017. 1090 [86] A. Gautier et al., “An Engineered Protein Tag for Multiprotein Labeling in Living Cells,” Chem. Biol., vol. 1091 15, no. 2, pp. 128–136, Feb. 2008. 1092 [87] G. V Los et al., “HaloTag: a novel protein labeling technology for cell imaging and protein analysis.,” ACS 1093 Chem. Biol., vol. 3, no. 6, pp. 373–82, Jun. 2008. 1094 [88] B. A. Griffin, S. R. Adams, and R. Y. Tsien, “Specific covalent labeling of recombinant protein molecules 1095 inside live cells.,” Science (80-. )., vol. 281, no. 5374, pp. 269–72, Jul. 1998. 1096 [89] M. Lelek et al., “Superresolution imaging of HIV in infected cells with FlAsH-PALM.,” Proc. Natl. Acad. 1097 Sci. U. S. A., vol. 109, no. 22, pp. 8564–9, May 2012. 1098 [90] T. Stephan, A. Roesch, D. Riedel, and S. Jakobs, “Live-cell STED nanoscopy of mitochondrial cristae,” 1099 Page 22 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- Sci. Rep., vol. 9, no. 1, p. 12419, 2019. 1100 [91] S. N. Uno, D. K. Tiwari, M. Kamiya, Y. Arai, T. Nagai, and Y. Urano, “A guide to use photocontrollable 1101 fluorescent proteins and synthetic smart fluorophores for nanoscopy,” Microscopy, vol. 64, no. 4, pp. 1102 263–277, 2015. 1103 [92] H. Takakura et al., “Long time-lapse nanoscopy with spontaneously blinking membrane probes,” Nat. 1104 Biotechnol., vol. 35, no. 8, pp. 773–780, 2017. 1105 [93] C. Wang, M. Taki, Y. Sato, A. Fukazawa, T. Higashiyama, and S. Yamaguchi, “Super-Photostable 1106 Phosphole-Based Dye for Multiple-Acquisition Stimulated Emission Depletion Imaging,” J. Am. Chem. 1107 Soc., vol. 139, no. 30, pp. 10374–10381, 2017. 1108 [94] L. Chu et al., “Two-color nanoscopy of organelles for extended times with HIDE probes,” bioRxiv, 2019. 1109 [95] A. N. Butkevich et al., “Hydroxylated Fluorescent Dyes for Live-Cell Labeling: Synthesis, Spectra and 1110 Super-Resolution STED.,” Chemistry (Easton)., vol. 23, no. 50, pp. 12114–12119, Sep. 2017. 1111 [96] I. Nikić et al., “Debugging Eukaryotic Genetic Code Expansion for Site-Specific Click-PAINT Super- 1112 Resolution Microscopy.,” Angew. Chem. Int. Ed. Engl., vol. 55, no. 52, pp. 16172–16176, Dec. 2016. 1113 [97] A. O. Khan, V. A. Simms, J. A. Pike, S. G. Thomas, and N. V Morgan, “CRISPR-Cas9 Mediated Labelling 1114 Allows for Single Molecule Imaging and Resolution.,” Sci. Rep., vol. 7, no. 1, p. 8450, Aug. 2017. 1115 [98] T. Schvartz et al., “Direct fluorescent-dye labeling of α-tubulin in mammalian cells for live cell and 1116 superresolution imaging.,” Mol. Biol. Cell, vol. 28, no. 21, pp. 2747–2756, Oct. 2017. 1117 [99] P. . B. T. . S. W. S. L. . W. R. . S. H. Wagner M.; Weber, “Light Dose is a Limiting Factor to Maintain Cell 1118 Viability in Fluorescence Microscopy and Single Molecule Detection,” Int. J. Mol. Sci., vol. 11, no. 3, pp. 1119 956–966, 2010. 1120 [100] M. E. Bulina, K. A. Lukyanov, O. V Britanova, D. Onichtchouk, S. Lukyanov, and D. M. Chudakov, 1121 “Chromophore-assisted light inactivation (CALI) using the phototoxic fluorescent protein KillerRed,” Nat. 1122 Protoc., vol. 1, no. 2, pp. 947–953, 2006. 1123 [101] M. M. Frigault, J. Lacoste, J. L. Swift, and C. M. Brown, “Live-cell microscopy - tips and tools,” J. Cell 1124 Sci., vol. 122, no. 6, pp. 753–767, 2009. 1125 [102] L. Song, E. J. Hennink, I. T. Young, and H. J. Tanke, “Photobleaching kinetics of fluorescein in 1126 quantitative fluorescence microscopy,” Biophys. J., vol. 68, no. 6, pp. 2588–2600, 1995. 1127 [103] J. W. Dobrucki, “Interaction of oxygen-sensitive luminescent probes Ru(phen)32+ and Ru(bipy)32+ with 1128 animal and plant cells in vitro: Mechanism of phototoxicity and conditions for non-invasive oxygen 1129 measurements,” J. Photochem. Photobiol. B Biol., vol. 65, no. 2, pp. 136–144, 2001. 1130 [104] J. A. Bloom and W. W. Webb, “Photodamage to intact erythrocyte membranes at high laser intensities: 1131 methods of assay and suppression.,” J. Histochem. Cytochem., vol. 32, no. 6, pp. 608–616, Jun. 1984. 1132 [105] C. M. Waterman-Storer, J. W. Sanger, and J. M. Sanger, “Dynamics of organelles in the mitotic spindles 1133 Page 23 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- of living cells: Membrane and microtubule interactions,” Cell Motil., vol. 26, no. 1, pp. 19–39, Jan. 1993. 1134 [106] W. Xiang, M. J. Roberti, J.-K. Hériché, S. Huet, S. Alexander, and J. Ellenberg, “Correlative live and 1135 super-resolution imaging reveals the dynamic structure of replication domains,” J. Cell Biol., vol. 217, no. 1136 6, pp. 1973–1984, Jun. 2018. 1137 [107] R. Kohen and A. Nyska, “Invited Review: Oxidation of Biological Systems: Oxidative Stress Phenomena, 1138 Antioxidants, Redox Reactions, and Methods for Their Quantification,” Toxicol. Pathol., vol. 30, no. 6, pp. 1139 620–650, Oct. 2002. 1140 [108] T.-W. Wu et al., “The cytoprotective effect of Trolox demonstrated with three types of human cells,” 1141 Biochem. Cell Biol., vol. 68, no. 10, pp. 1189–1194, Oct. 1990. 1142 [109] I. Rasnik, S. A. McKinney, and T. Ha, “Nonblinking and long-lasting single-molecule fluorescence 1143 imaging,” Nat. Methods, vol. 3, no. 11, pp. 891–893, 2006. 1144 [110] A. M. Bogdanov, E. I. Kudryavtseva, and K. A. Lukyanov, “Anti-Fading Media for Live Cell GFP Imaging,” 1145 PLoS One, vol. 7, no. 12, p. e53004, Dec. 2012. 1146 [111] P. Atkins and J. de Paula, Physical Chemistry, Eigth edit. New York, US: W.H. Freeman and Company, 1147 2006. 1148 [112] J. Widengren, A. Chmyrov, C. Eggeling, P.-Å. Löfdahl, and C. A. M. Seidel, “Strategies to Improve 1149 Photostabilities in Ultrasensitive Fluorescence Spectroscopy,” J. Phys. Chem. A, vol. 111, no. 3, pp. 429– 1150 440, Jan. 2007. 1151 [113] H. Giloh and J. W. Sedat, “Fluorescence microscopy: reduced photobleaching of rhodamine and 1152 fluorescein protein conjugates by n-propyl gallate,” Science (80-. )., vol. 217, no. 4566, pp. 1252–1255, 1153 Sep. 1982. 1154 [114] K. Valnes and P. Brandtzaeg, “Retardation of immunofluorescence fading during microscopy.,” J. 1155 Histochem. Cytochem., vol. 33, no. 8, pp. 755–761, Aug. 1985. 1156 [115] G. D. Johnson, R. S. Davidson, K. C. McNamee, G. Russell, D. Goodwin, and E. J. Holborow, “Fading of 1157 immunofluorescence during microscopy: a study of the phenomenon and its remedy,” J. Immunol. 1158 Methods, vol. 55, no. 2, pp. 231–242, 1982. 1159 [116] A. Longin, C. Souchier, M. Ffrench, and P. A. Bryon, “Comparison of anti-fading agents used in 1160 fluorescence microscopy: image analysis and laser confocal microscopy study.,” J. Histochem. 1161 Cytochem., vol. 41, no. 12, pp. 1833–1840, Dec. 1993. 1162 [117] A. N. Fletcher and M. E. Pietrak, “Laser dye stability, part 10. Effects of DABCO on flashlamp pumping of 1163 coumarin dyes,” Appl. Phys. B, vol. 37, no. 3, pp. 151–157, 1985. 1164 [118] M. Ono, T. Murakami, A. Kudo, M. Isshiki, H. Sawada, and A. Segawa, “Quantitative Comparison of Anti- 1165 Fading Mounting Media for Confocal Laser Scanning Microscopy,” J. Histochem. Cytochem., vol. 49, no. 1166 3, pp. 305–311, Mar. 2001. 1167 Page 24 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- [119] A. M. Bogdanov, E. A. Bogdanova, D. M. Chudakov, T. V Gorodnicheva, S. Lukyanov, and K. A. 1168 Lukyanov, “Cell culture medium affects GFP photostability: a solution,” Nat. Methods, vol. 6, no. 12, pp. 1169 859–860, 2009. 1170 [120] A. M. Edwards, E. Silva, B. Jofré, M. I. Becker, and A. E. De Ioannes, “Visible light effects on tumoral 1171 cells in a culture medium enriched with tryptophan and riboflavin,” J. Photochem. Photobiol. B Biol., vol. 1172 24, no. 3, pp. 179–186, 1994. 1173 [121] A. Grzelak, B. Rychlik, and G. Bartosz, “Light-dependent generation of reactive oxygen species in cell 1174 culture media,” Free Radic. Biol. Med., vol. 30, no. 12, pp. 1418–1425, 2001. 1175 [122] G. T. Spierenburg, F. T. J. J. Oerlemans, J. P. R. M. Van Laarhoven, and C. H. M. M. De Bruyn, 1176 “Phototoxicity of N-hydroxyethylpiperazine-N’-2-ethanesulfonic acid-buffered culture media for human 1177 leukemic cell lines,” Cancer Res., vol. 44, pp. 2253–2254, 1984. 1178 [123] J. S. Zigler, J. L. Lepe-Zuniga, B. Vistica, and I. Gery, “Analysis of the cytotoxic effects of light-exposed 1179 hepes-containing culture medium,” Vitr. Cell. Dev. Biol., vol. 21, no. 5, pp. 282–287, 1985. 1180 [124] P. Almada, S. Culley, and R. Henriques, “PALM and STORM: Into large fields and high-throughput 1181 microscopy with sCMOS detectors,” Methods, vol. 88, pp. 109–121, 2015. 1182 [125] S. Herbert, H. Soares, C. Zimmer, and R. Henriques, “Single-Molecule Localization Super-Resolution 1183 Microscopy: Deeper and Faster,” Microsc. Microanal., vol. 18, no. 6, pp. 1419–1429, 2012. 1184 [126] P. Almada et al., “Automating multimodal microscopy with NanoJ-Fluidics,” Nat. Commun., vol. 10, no. 1, 1185 p. 1223, 2019. 1186 [127] G. Donnert, C. Eggeling, and S. W. Hell, “Major signal increase in fluorescence microscopy through dark- 1187 state relaxation,” Nat. Methods, vol. 4, no. 1, pp. 81–86, 2007. 1188 [128] K. König, T. W. Becker, P. Fischer, I. Riemann, and K.-J. Halbhuber, “Pulse-length dependence of 1189 cellular response to intense near-infrared laser pulses in multiphoton microscopes,” Opt. Lett., vol. 24, 1190 no. 2, pp. 113–115, 1999. 1191 [129] K. T. Takasaki, J. B. Ding, and B. L. Sabatini, “Live-Cell Superresolution Imaging by Pulsed STED Two- 1192 Photon Excitation Microscopy,” Biophys. J., vol. 104, no. 4, pp. 770–777, 2013. 1193 [130] P. Bethge, R. Chéreau, E. Avignone, G. Marsicano, and U. V. Nägerl, “Two-Photon Excitation STED 1194 Microscopy in Two Colors in Acute Brain Slices,” Biophys. J., vol. 104, no. 4, pp. 778–785, 2013. 1195 [131] E. Schmidt and M. Oheim, “Two-photon imaging induces brain heating and calcium microdomain 1196 hyperactivity in cortical astrocytes,” bioRxiv, 2018. 1197 [132] B. Harke, J. Keller, C. K. Ullal, V. Westphal, A. Schönle, and S. W. Hell, “Resolution scaling in STED 1198 microscopy,” Opt. Express, vol. 16, no. 6, pp. 4154–4162, 2008. 1199 [133] G. Vicidomini et al., “Sharper low-power STED nanoscopy by time gating,” Nat. Methods, vol. 8, no. 7, 1200 pp. 571–573, 2011. 1201 Page 25 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- [134] Y. Wu and H. Shroff, “Faster, sharper, and deeper: structured illumination microscopy for biological 1202 imaging,” Nat. Methods, vol. 15, no. 12, pp. 1011–1019, 2018. 1203 [135] A. G. York et al., “Resolution doubling in live, multicellular organisms via multifocal structured illumination 1204 microscopy,” Nat. Methods, vol. 9, no. 7, pp. 749–754, 2012. 1205 [136] I. Gregor, M. Spiecker, R. Petrovsky, J. Großhans, R. Ros, and J. Enderlein, “Rapid nonlinear image 1206 scanning microscopy,” Nat. Methods, vol. 14, no. 11, pp. 1087–1089, 2017. 1207 [137] E. N. Ward and R. Pal, “Image scanning microscopy: an overview,” J. Microsc., vol. 266, no. 2, pp. 221– 1208 228, 2017. 1209 [138] X. Huang et al., “Fast, long-term, super-resolution imaging with Hessian structured illumination 1210 microscopy,” Nat. Biotechnol., vol. 36, p. 451, Apr. 2018. 1211 [139] E. G. Reynaud, U. Kržič, K. Greger, and E. H. K. Stelzer, “Light sheet‐based fluorescence microscopy: 1212 More dimensions, more photons, and less photodamage,” HFSP J., vol. 2, no. 5, pp. 266–275, Oct. 2008. 1213 [140] E. G. Reynaud, J. Peychl, J. Huisken, and P. Tomancak, “Guide to light-sheet microscopy for 1214 adventurous biologists,” Nat. Methods, vol. 12, no. 1, pp. 30–34, 2014. 1215 [141] R. M. Power and J. Huisken, “A guide to light-sheet fluorescence microscopy for multiscale imaging,” 1216 Nat. Methods, vol. 14, no. 4, pp. 360–373, 2017. 1217 [142] F. Cella Zanacchi et al., “Live-cell 3D super-resolution imaging in thick biological samples,” Nat. Methods, 1218 vol. 8, no. 12, pp. 1047–1049, 2011. 1219 [143] F. Cella Zanacchi, Z. Lavagnino, M. Faretta, L. Furia, and A. Diaspro, “Light-Sheet Confined Super- 1220 Resolution Using Two-Photon Photoactivation,” PLoS One, vol. 8, no. 7, p. e67667, Jul. 2013. 1221 [144] J. Kim et al., “Oblique-plane single-molecule localization microscopy for tissues and small intact animals,” 1222 Nat. Methods, vol. 16, no. 9, pp. 853–857, 2019. 1223 [145] P. Hoyer et al., “Breaking the diffraction limit of light-sheet fluorescence microscopy by RESOLFT,” Proc. 1224 Natl. Acad. Sci., vol. 113, no. 13, pp. 3442 LP – 3446, Mar. 2016. 1225 [146] Y. S. Hu et al., “Light-sheet Bayesian microscopy enables deep-cell super-resolution imaging of 1226 heterochromatin in live human embryonic stem cells,” Opt. Nanoscopy, vol. 2, no. 1, p. 7, 2013. 1227 [147] T. A. Planchon et al., “Rapid three-dimensional isotropic imaging of living cells using Bessel beam plane 1228 illumination,” Nat. Methods, vol. 8, no. 5, pp. 417–423, 2011. 1229 [148] L. Gao et al., “Noninvasive Imaging beyond the Diffraction Limit of 3D Dynamics in Thickly Fluorescent 1230 Specimens,” Cell, vol. 151, no. 6, pp. 1370–1385, 2012. 1231 [149] B.-J. Chang, V. D. Perez Meza, and E. H. K. Stelzer, “csiLSFM combines light-sheet fluorescence 1232 microscopy and coherent structured illumination for a lateral resolution below 100 nm,” Proc. Natl. Acad. 1233 Sci., vol. 114, no. 19, pp. 4869 LP – 4874, May 2017. 1234 [150] B.-C. Chen et al., “Lattice light-sheet microscopy: Imaging molecules to embryos at high spatiotemporal 1235 Page 26 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- resolution,” Science (80-. )., vol. 346, no. 6208, p. 1257998, Oct. 2014. 1236 [151] N. Chakrova, A. S. Canton, C. Danelon, S. Stallinga, and B. Rieger, “Adaptive illumination reduces 1237 photobleaching in structured illumination microscopy,” Biomed. Opt. Express, vol. 7, no. 10, pp. 4263– 1238 4274, 2016. 1239 [152] J. Dreier et al., “Smart scanning for low-illumination and fast RESOLFT nanoscopy in vivo,” Nat. 1240 Commun., vol. 10, no. 1, p. 556, 2019. 1241 [153] T. Staudt, A. Engler, E. Rittweger, B. Harke, J. Engelhardt, and S. W. Hell, “Far-field optical nanoscopy 1242 with reduced number of state transition cycles,” Opt. Express, vol. 19, no. 6, pp. 5644–5657, 2011. 1243 [154] F. Göttfert et al., “Strong signal increase in STED fluorescence microscopy by imaging regions of 1244 subdiffraction extent,” Proc. Natl. Acad. Sci., vol. 114, no. 9, pp. 2125 LP – 2130, Feb. 2017. 1245 [155] J. Heine, M. Reuss, B. Harke, E. D’Este, S. J. Sahl, and S. W. Hell, “Adaptive-illumination STED 1246 nanoscopy,” Proc. Natl. Acad. Sci., vol. 114, no. 37, pp. 9797–9802, 2017. 1247 [156] D. Sage et al., “Super-resolution fight club: assessment of 2D and 3D single-molecule localization 1248 microscopy software,” Nat. Methods, vol. 16, no. 5, pp. 387–395, 2019. 1249 [157] A. Jimenez, K. Friedl, and C. Leterrier, “About samples, giving examples: Optimized Single Molecule 1250 Localization Microscopy,” Methods, 2019. 1251 [158] T. Dertinger, R. Colyer, G. Iyer, S. Weiss, and J. Enderlein, “Fast, background-free, 3D super-resolution 1252 optical fluctuation imaging (SOFI),” Proc. Natl. Acad. Sci., vol. 106, no. 52, pp. 22287 LP – 22292, Dec. 1253 2009. 1254 [159] S. Geissbuehler et al., “Live-cell multiplane three-dimensional super-resolution optical fluctuation 1255 imaging,” Nat. Commun., vol. 5, pp. 1–7, 2014. 1256 [160] S. Cox et al., “Bayesian localization microscopy reveals nanoscale podosome dynamics,” Nat. Methods, 1257 vol. 9, no. 2, pp. 195–200, 2012. 1258 [161] N. Gustafsson, S. Culley, G. Ashdown, D. M. Owen, P. M. Pereira, and R. Henriques, “Fast live-cell 1259 conventional fluorophore nanoscopy with ImageJ through super-resolution radial fluctuations,” Nat. 1260 Commun., vol. 7, no. 1, p. 12471, 2016. 1261 [162] R. F. Laine et al., “NanoJ: a high-performance open-source super-resolution microscopy toolbox,” J. 1262 Phys. D. Appl. Phys., vol. 52, no. 16, p. 163001, Feb. 2019. 1263 [163] S. Culley, K. L. Tosheva, P. Matos Pereira, and R. Henriques, “SRRF: Universal live-cell super-resolution 1264 microscopy,” Int. J. Biochem. Cell Biol., vol. 101, pp. 74–79, 2018. 1265 [164] R. J. Marsh et al., “Artifact-free high-density localization microscopy analysis,” Nat. Methods, vol. 15, no. 1266 9, pp. 689–692, 2018. 1267 [165] R. J. Marsh, S. Culley, and A. J. Bain, “Low power super resolution fluorescence microscopy by lifetime 1268 modification and image reconstruction,” Opt. Express, vol. 22, no. 10, pp. 12327–12338, 2014. 1269 Page 27 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript ----!@#$NewPage!@#$---- [166] L. von Chamier, R. F. Laine, and R. Henriques, “Artificial intelligence for microscopy: what you should 1270 know,” Biochem. Soc. Trans., vol. 47, no. 4, pp. 1029–1040, Jul. 2019. 1271 [167] C. Belthangady and L. A. Royer, “Applications, promises, and pitfalls of deep learning for fluorescence 1272 image reconstruction,” Nat. Methods, 2019. 1273 [168] M. Weigert et al., “Content-aware image restoration: pushing the limits of fluorescence microscopy,” Nat. 1274 Methods, vol. 15, no. 12, pp. 1090–1097, 2018. 1275 [169] W. Ouyang, A. Aristov, M. Lelek, X. Hao, and C. Zimmer, “Deep learning massively accelerates super- 1276 resolution localization microscopy,” Nat. Biotechnol., vol. 36, p. 460, Apr. 2018. 1277 [170] L. Fang et al., “Deep Learning-Based Point-Scanning Super-Resolution Imaging,” bioRxiv, 2019. 1278 [171] S. Culley et al., “Quantitative mapping and minimization of super-resolution optical imaging artifacts,” Nat. 1279 Methods, vol. 15, p. 263, Feb. 2018. 1280 [172] G. Ball, J. Demmerle, R. Kaufmann, I. Davis, I. M. Dobbie, and L. Schermelleh, “SIMcheck: a Toolbox for 1281 Successful Super-resolution Structured Illumination Microscopy,” Sci. Rep., vol. 5, p. 15915, Nov. 2015. 1282 1283 Page 28 of 28 AUTHOR SUBMITTED MANUSCRIPT - JPhysD-122201.R1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Accepted Manuscript