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 61 62

The American Journal of Pathology, Vol. -, No. -, - 2015

ajp.amjpathol.org

Regenerative Medicine Theme Issue

REVIEW miRNA Control of Tissue Repair and Regeneration Q13

Chandan K. Sen and Subhadip Ghatak From the Center for Regenerative Medicine and Cell-Based Therapies and the Department of Surgery, The Ohio State University Wexner Medical Center, Columbus, Ohio Accepted for publication April 2, 2015. Address correspondence to Chandan K. Sen, Ph.D., The Ohio State University Wexner Medical Center, 473 W. 12th Ave., Columbus, OH 43210. E-mail: chandan.sen@osumc. edu.

Q3

Tissue repair and regeneration rely on the function of miRNA, molecular silencers that enact posttranscriptional gene silencing of coding genes. Disruption of miRNA homeostasis is developmentally lethal, indicating that fetal tissue development is tightly controlled by miRNAs. Multiple critical facets of adult tissue repair are subject to control by miRNAs as well. Sources of cell pool for tissue repair and regeneration are diverse and provided by processes, including cellular dedifferentiation, transdifferentiation, and reprogramming. Each of these processes is regulated by miRNAs. Furthermore, induced pluripotency may be achieved by miRNA-based strategies independent of transcription factor manipulation. The observation that miRNA does not integrate into the genome makes miRNA-based therapeutic strategies translationally valuable. Tools to manipulate cellular and tissue miRNA levels include mimics and inhibitors that may be specifically targeted to cells of interest at the injury site. Here, we discuss the extraordinary importance of miRNAs in tissue repair and regeneration based on emergent reports and rapid advances in miRNA-based therapeutics. (Am J Pathol 2015, -: 1e13; http://dx.doi.org/10.1016/j.ajpath.2015.04.001)

Injury-responsive coding genes are recognized as a main driver of wound healing and tissue regeneration.1 After injury, tissue healing is initiated either by regeneration or repair or by a combination of both. Although robust tissue regeneration is observed in a number of lower vertebrates, including urodele amphibians and teleost fish, mammalian tissue regeneration is limited, particularly in adults.2 A cornerstone in the process of regeneration is the expression of injury-inducible coding genes at the site of tissue injury. However, the simultaneous expression of an array of injury responsive coding genes, after injury, complicates signaling networks. Post-transcriptional gene silencing (PTGS) may be viewed as a filter that is aimed at selectively advancing limited sets of injury-responsive coding genes toward protein expression to streamline the repair and regeneration process. Earlier, approximately 97% of human DNA was considered as junk because it did not encode for protein.3 However, the current literature recognizes a critical role of noncoding DNA in biology. What used to be known as junk DNA is now known to produce approximately 22nucleotide long evolutionarily conserved and functionally

Q2

critical single-stranded miRNA molecules that specifically silence mRNA function predominantly by targeting the 30 untranslated regions of mRNA.4 miRNAs play a main role in PTGS by sequence-specific inhibition of gene expression5 that may occur at different levels such as chromatin modification and silencing, translational repression, and mRNA degradation. Thus, miRNA biology determines the biological functionality of coding genes. According to the current edition of miRbase, a central online repository for miRNAs, there are 1881 hairpin precursors and 2588 mature miRNAs enlisted for humans. This list is rapidly expanding. Because the function of miRNA determines the functional fate of mRNA, understanding the importance of miRNA in the context of tissue repair and regeneration becomes critically important. Here, we discuss the developments recorded in the current literature that underscore the rapidly unfolding importance Supported by NIH/National Institute of General Medical Sciences grants GM069589, GM077185, GM108014, and NR013898 (C.K.S.). Q1 Disclosures: None declared. This article is part of a review series on regenerative medicine.

Copyright ª 2015 American Society for Investigative Pathology. Published by Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.ajpath.2015.04.001

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124

125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186

Sen and Ghatak of miRNA and related processes in tissue regeneration and repair. Q4

Q5

Development is the Key to Regeneration In mammals, the early in utero state has provided some of the most impressive evidence of tissue regeneration.6 Such remarkable regenerative potential is markedly blunted or lost in the postnatal period and more so during adulthood and old age. Comparison of global expression pattern of mature miRNA from two groups of mouse fetal skin has provided critical insight.7 The skin at embryonic day (E)16, when the healing is regenerative and scarless, was associated with global miRNA repression compared with the skin of E19 when the healing featured adult phenotype, including scar formation.7 Of note in this context is the observation that miRNA biogenesis is substantially blunted at E16, indicative of dampened PTGS in the fetal skin.7 Consistently, miRNA function is also globally suppressed in mouse oocytes and early embryo.8 The global suppression of miRNA during mouse oocyte-to-embryo transition is likely facilitated by the expression of highly conserved RNA binding proteins Lin28a and Lin28b that are abundant during embryogenesis. Lin28a and Lin28b repress miRNA biogenesis.9 Lin28, a putative pluripotent factor, downregulates let-7 miRNA by inducing the uridylation of let-7 precursor (preelet-7) by using a noncanonical poly (A) polymerase, TUTase4.10 These reports lead to the hypothesis that fetal tissue development is enabled by transiently silencing miRNA-dependent PTGS. Such silencing of the silencer unleashes the expression of a larger number of coding genes necessary for tissue development.11 After spinal cord injury in Wistar rats, miRNA expression is markedly subdued from day 3 after injury with gradual rebound of down-regulated miRNA at 7 days after injury.12 Suppression of miRNA expression after spinal cord injury was verified in an independent study in which downregulated miRNAs were reported up to 14 days after injury.13 The pattern of miRNA suppression after injury holds across organ systems. Partial hepatectomy is followed by down-regulation of 70% of the miRNA within 24 hours of injury. Of particular interest in this study was the observation that miRNA down-regulation was preceded by a transient upregulation of miRNA 3 hours after injury. This acute-phase response of selective miRNA up-regulation was interpreted as a response aimed at priming and commitment to liver regeneration. The subsequent genome-wide down-regulation of miRNAs was required for efficient recovery of liver cell mass to support tissue regeneration. In patients undergoing auxiliary liver transplantation, down-regulation of miRNA is associated with induced cell proliferation that supported tissue regeneration.14 Failed regeneration in a similar cohort was attributed to expression of specific miRNA which enforced cell cycle inhibition and DNA methylation.14 Taken together, these observations support the hypothesis that

2

injury-induced down-regulation of miRNAs may serve as an intrinsic physiologic strategy to transiently de-silence coding genes necessary for the tissue repair and regeneration processes. In adult tissue, resident stem cells (SCs) are usually postmitotic and quiescent and are substantially different from embryonic SCs (ESCs), which are actively proliferating and differentiating.15 However, external stimuli in the form of injury may suppress tissue-specific miRNAs, which in turn may lead to alterations in adult resident SCs such that they start to differentiate, akin to their embryonic counterpart, with the objective to enact tissue repair. This is one example of how regeneration is largely inspired by mechanistic underpinning of embryo development. Some of the widespread principles and processes that drive tissue regeneration include cellular differentiation, dedifferentiation, transdifferentiation, and reprogramming.16

Dedifferentiation, Transdifferentiation, and Cellular Reprogramming Unlike lower vertebrates, mammalian tissue regeneration or repair processes rely on pre-existing stem and other cells with plasticity to replace the injured tissue.17 The sources of such cell populations are limited. In response to a strong microenvironmental cue triggered by the injury, behavior of some of the cells in close proximity of the injury site changes to accommodate the emergence of new cell types. Plasticity of surviving cells at or close to the injury site provides additional pools of cell populations that may contribute to the repair or regeneration process. With the use of a fluorescent-labeled tracking method, it is observed that terminally differentiated cells may revert back to a less differentiated and more plastic state.18 The process wherein terminally differentiated cells regress from a specialized function to a simpler less differentiated state, a more SC-like state, is referred to as dedifferentiation.19 Mammalian dedifferentiation is evident in myoblasts, renal cells, oligodendrocyte precursor cells, and even in germ cells.19 The process of dedifferentiation commonly involves re-entry of regenerating cells into the cell cycle process.20 Dedifferentiation allows the cell to proliferate again before re-differentiating, permitting a variety of cell types to be replaced at the site of injury. The process of cell reversion to a more pluripotent state is demonstrated experimentally by transferring somatic cell nuclei to eggs and fusing somatic cells with pluripotent cells.21 Nuclei of fertilized egg or the pluripotent SCs (PSCs) contain some putative reprogramming factors that are capable of erasing the memory of the differentiated somatic cell. More recently we learned that a combination of only four transcription factors (ie, Oct3/4, Sox2, Klf4, and c-Myc) is sufficient to revert differentiated somatic cells into an embryonic fate somewhat akin to that of ESCs.22 The critical importance of transcription factor networks in determining stemness and cell fate was thus recognized. The emergent field of nuclear reprograming

ajp.amjpathol.org

-

The American Journal of Pathology

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

187 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248

249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310

miRNA and Regenerative Medicine has provided critical insight into a number of fundamental cell biology processes, including transdifferentiation or direct reprogramming. Although the two terms are often used interchangeably, there are important distinctions between the two. Transdifferentiation is a spontaneous change of cell fate that typically happens in vivo under conditions of stress or pathologic insult. In contrast, direct reprogramming is induced by defined factors such as chemical compounds, transcription factors, or other artificial engineering methods.23 In other words, direct reprogramming could be defined as transdifferentiation that can only happen under stringent artificial control conditions.24 Transdifferentiation advances the notion of dedifferentiation to one step further whereby all of the cells regress to a point from where they may differentiate to multiple specialized cell types. Transdifferentiation may also bypass the pluripotent state. With the use of defined transcription factors such as Ascl1/Brn2/Myt1l, Gata4/Mef2c/Tbx5, and Hnf4a/ Foxa, fibroblasts may be reprogrammed to neurons, cardiomyocytes, and hepatocytes, respectively.25 Such inducible transdifferentiation process bypasses the pluripotent state and is therefore often referred to as direct reprogramming.26 Under normal physiologic conditions, the transdifferentiation process seems to be rare in mammals, although a closer look is warranted. However, during the early postnatal development smooth muscle of the mouse esophagus may give rise to skeletal muscle tissue.27 In adults, dedifferentiation and transdifferentiation may occur naturally during the early phase of tissue repair. These subtle changes in cellular plasticity are driven by an orchestrated regulation of gene expression networks to reconstruct tissues in a sophisticated manner that avoids undesirable consequences such as neoplasia. Although the reappearance of developmental phenotype is reported during tissue regeneration,28 the molecular underpinnings remain largely elusive. What fills part of this void is the emergent observations that span the diverse phylogenetic tree of the animal kingdom, demonstrating that miRNAs play a key role in tissue repair and regeneration.29 Given their capacity to determine the functional fate of coding genes and therefore a large network of downstream signaling pathways,30 miRNAs are rapidly emerging as potential candidates for the therapeutic modulation of tissue repair and regeneration. In this context, it is of extraordinary importance that the turnover of miRNA within the cell determines its functional fate.

miRNA Turnover Determines Cell Fate Tissue regeneration is markedly compromised with aging, and the underlying reasons remain largely elusive.31 Lin28a improve tissue repair by suppressing let-7 biogenesis and enhancing oxidative metabolism.31 Several mature miRNAs exhibit tissue-specific expression pattern, although their primary transcripts show no important changes in expression pattern.32 Although several aspects of global miRNA biogenesis pathways are well understood,5 little is known

The American Journal of Pathology

-

311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 miRNA Biogenesis 345 346 For a detailed review of miRNA biogenesis, please refer to 34,37 347 the recent literature. Broadly, miRNA biogenesis in348 volves two sequential processing events that trim the tran349 script to yield a mature functional miRNA. Apart from the 350 known canonical and noncanonical pathways, miRNA 351 biogenesis from the loop sequence, tRNA-derived RNA 352 fragments and the miRNA arm selection have now become 353 evident. Several kilobase long pri-miRNAs along with the 354 stem loop structure are cropped into a 60- to 100-nt hairpin355 structured precursor (the pre-miRNA) by RNase III Drosha 356 357 and the DGCR8 protein that constitutes the multiprotein 358 microprocessor complex along with several auxiliary co359 factors that coordinate activity and specificity of Drosha cleavage (Figure 1). Binding of DGCR8 to the pri-miRNA ½F1 360 361 facilitates precise positioning and orientation of Drosha’s 362 catalytic center approximately 11 nt upstream of the 363 stemeloop structure. In this way a double-stranded pre364 0 miRNA with a 2-nt 3 overhang is generated. The expres365 sion and activity of Drosha and DGCR8 are subject to 366 stringent regulation by multiple factors that in turn influence 367 intracellular turnover of miRNAs. For example, DGCR8 368 369 stabilizes Drosha, whereas Drosha controls DGCR8 levels 370 by cleaving hairpins present in the DGCR8 mRNA, thereby 371 inducing its degradation. The pre-miRNA is exported to the 372

about the specific mechanisms that regulate the expression of individual miRNAs that exhibit distinctive expression patterns in certain tissues.33 Interestingly, each step of the general miRNA biogenesis pathway seems to be subject to fine-tuning to allow for exquisite control of expression of specific individual miRNAs. Studies have shown that a battery of proteins take part in the regulation of miRNA biogenesis either by interacting with the Drosha-DiGeorge syndrome critical region gene (DGCR8) microprocessor complex, Dicer, or the miRNA precursors.34 At the microprocessor level, the p68 and p72 helicases regulate the biogenesis of approximately one-third murine pri-miRNA.35 Similarly, individual miRNA processing, such as priemiR21 or priemiR-7, is greatly influenced by other proteine protein interactions such as the SMADep68 complex or SMADeSMAD nuclear interacting protein 1 and splicing factor SF2/ASF.36 Similarly, heterogeneous nuclear ribonucleoprotein A1 facilitates Drosha-mediated processing by binding to the loop regions.36 The KH-type splicing regu- Q6 latory protein enhances Drosha and Dicer processing by binding to the GGG triplet motifs in their terminal loops of the pri- and pre-miRNA.36 Although the effect of Lin28 repressor is mostly restricted to let-7 family members, the nuclear factor NF90eNF45 heterodimer impairs DGCR8 binding by interacting with the stem of pri-miRNAs in a sequence-independent way.36 Finally, editing of primiRNAs or pre-miRNAs by adenosine deaminases 1 and 2 may affect mature miRNA accumulation and influence miRNA-target specificity.36

ajp.amjpathol.org

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

3

Figure 1 The miRNA biogenesis pathway. The pri-miRNA transcript is cleaved by the microprocessor complex Drosha-DGCR8 in the nucleus. The precursor hairpin pre-miRNA is exported from the nucleus by Exportin-5eRan-GTP. In the cytoplasm, the RNaseIII Dicer forms complex with the doublestranded RNA-binding protein TRBP and AGO-2 forms the RISC. The functional strand of the mature miRNA is loaded together with AGO-2 proteins into the RISC. The mature miRNA silences target mRNAs through mRNA cleavage, translational repression, or deadenylation, whereas the passenger strand (black) is degraded. AGO-2, argonaute-2; DGCR8, DiGeorge syndrome critical region gene 8; pre-miRNA, precursor miRNA; pri-miRNA, primary miRNA; RISC, RNA-induced silencing complex; TRBP, TAR RNA binding protein.

print & web 4C=FPO

373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434

Sen and Ghatak

cytoplasm by exportin-5 via a RAN-GTP dependent mechanism. The 2-nt 30 overhang plays a crucial role in recognizing the pre-miRNA by exportin-5, permitting the export of only correctly processed miRNA precursors into the cytoplasm. Disruption of enzymes such as Drosha or DGCR8, which are essential for the miRNA maturation, results in rapid proliferation of cancer cells.38 In the cytoplasm, pre-miRNAs are further cleaved near the terminal loop by another key RNase III enzyme called the Dicer. In mouse fetal skin, Dicer is substantially downregulated, indicating suppressed global miRNA biogenesis that is likely to prevent silencing of numerous coding genes aimed at tissue regeneration often evident at E16.7 Dicer acts as a molecular ruler that cleaves the pre-miRNA at a specific distance from the ends produced by Drosha cleavage and generating an approximate 22-nt double-stranded miRNA duplex (miR/miR*) with a 2-nt 30 overhang. Dicer, together with TAR RNA binding protein, protein kinase RNA activator, and argonaute (AGO) proteins, forms the RNAinduced silencing complex (RISC). It was initially thought that RISC incorporates one of the strands of the diced miRNA duplex with the lowest base-pairing stability at the 50 end. The other strand, originally termed miR*, is degraded on the basis of thermodynamic principles and energyindependent endonuclease-mediated selection with Ago2. Strands are present in the miR-RISC complex and are named as -5p or -3p, depending on their relative position in the hairpin sequence. Evidence shows that the target abundance stabilizes the less-dominant strand of miRNA, implying that the target transcript regulation may affect the selection of mature miRNAs. We now know that 3p-miRNA is successfully activated if the pre-miRNA is 50 -capped. In addition to arm selection, the loop region of the pre-miRNA hairpin may also be used for so-called loop-miRNA, suggesting that

4

double strands generated by typical miRNA processing are not absolutely needed for the functioning of certain miRNAs. The primary function of an miRNA is to target its cognate mRNA(s), a process governed by base pairing between the seed region of miRNA with the complementary sequences within the 30 -untranslated region of the targeted mRNA. Functional repression depends on the extent of complementarity. If the binding is perfect, it leads to mRNA cleavage and degradation. If the binding is imperfect, it may cause translational repression. In animal cells, miRNAs fine-tune protein translation both by direct and indirect influences. Direct effects are exerted at multiple phases of protein translation, including repression at the initiation, prevention of ribosome assembly for the initiation of translation, repression at postinitiation steps, and inhibition of elongation or termination of translation process. Indirect translational repression by miRNA may occur at the mRNA level and is caused by the destabilization and subsequent degradation and the compartmentalization and sequestration of the target mRNAs. In this mode of action, miRNAs destabilize their target mRNAs primarily through deadenylation. The compartmentalization and sequestration processes rely on cytoplasmic foci known as processing bodies where repressed mRNAs are sequestered with enriched translational repressors.

miRNA Degradation Although miRNAs are globally stable with half-lives that may extend up to days,39 it is now clear that the absolute levels of mature miRNAs are also controlled by cis- and trans-acting factors that directly regulate stability. Such specific modifications and exonucleases may profoundly influence the lifetime of any given miRNA. Stemming from these initial studies are three areas that warrant deeper

ajp.amjpathol.org

-

The American Journal of Pathology

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496

print & web 4C=FPO

497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558

miRNA and Regenerative Medicine

Figure 2

Enzymes that act on the pre-miRNA and mature miRNA to determine miRNA turnover. The pri-miRNA transcript is processed by Drosha to generate the pre-miRNA. The miRNA biogenesis depends on the balance of productive and abortive ribonucleases such as Dicer and MCPIP1, both of which act on the same pre-miRNA. Argonaute and GLD-2 act on mature miRNA for maturation and stabilization. Similarly PUP-2, XRN-2, and PNPaseold35 act on mature miRNA for degradation. MCPIP1, monocyte chemotactic protein-induced protein 1; PNPase, polynucleotide phosphorylase; pre-miRNA, precursor miRNA; pri-miRNA, primary miRNA.

investigation to further elucidate the causes and effects of altered miRNA homeostasis at the level of mature miRNA stability.40 These three areas are i) miRNA modifications, ii) relation between miRNA function and stability, and iii) targeted degradation of specific miRNA.40 On an as-needed basis, reversible muting of PTGS may ½F2 be enacted by several miRNases (Figure 2) or exoribonucleases that may degrade cellular miRNA, thereby regulating downstream functional impact. The question is how do miRNAs resist attack by ribonucleases? The silencing complex miRISC may shield its resident miRNA from cleavage by ribonucleases. In fact, the core components of silencing complexes, AGO proteins, act to enhance the abundance of miRNAs besides acting as a silencer.41 Structural biology data of small RNA and AGO suggest that AGO tightly binds to small RNA. Both 50 and 30 ends of small RNA are buried within the AGO protein.42 How the AGO protein unfolds to make the miRNA accessible for exoribonuclease degradation remains unknown.

miRNA Signature in SCs

Q7

Several studies have consistently documented that only a subset of miRNA is expressed in SCs. Deep sequencing studies have revealed that miRNA expression in naive ESCs exhibit a different signature compared with primed epiblast SCs with only one-third of the total miRNA being differentially expressed.43 Differential expression of several clusters, including miR-290/miR-295, miR-17/miR-92, miR-302/ miR-367, was reported along with a large repetitive cluster on chromosome 2.43 Although the term induced PSCs (iPSCs) is often referred to as nearly identical to their embryoderived counterparts, miRNA expression profiling between the human ESCs and human iPSCs demonstrates an important difference in a handful of miRNAs.44 Genome-wide mapping of miRNA expressed in ESCs and PSCs generated by different reprogramming strategies such as iPSCs or nuclear transfer ESCs, recognized 34 miRNAs to be differentially regulated in reprogrammed cells compared with ESCs, among which miR-24 and miR-370 were previously reported.45

The American Journal of Pathology

-

These reports demonstrate that SCs or pluripotent cells have their own distinct signature irrespective of their origin. This observation sheds light on the intriguing potential that miRNAs may function as key drivers in regeneration.46 Loss of Dicer is embryonically lethal because they lack almost all PSCs.47 In vitro, ESCs with Dicer deletion do not proliferate.48 DGCR8 ablation also blunts proliferation of ESCs which go on to accumulate in the G1 phase of the cell cycle.49 Current observations lead to the notion that an intact miRNA biogenesis pathway is required for tissue development and raises the possibility that early steps in embryonic development require the action of specific miRNAs to properly regulate SC function. Delivery of multiple members of the miR-290 family rescued proliferation in DGCR8-deficient ESCs, although they could not restore differentiation defects.50 These miRNAs are referred to as ESC-specific cell cycle-regulating miRNAs. Interestingly, proliferation and self-renewal may be blocked in ESCs by expression of let-7 miRNA, indicating that distinct subsets of miRNAs regulate SC maintenance, proliferation, and differentiation.51 Broadly, this is consistent with findings that overall miRNA complexity increases with development. More and more miRNAs are expressed as cells differentiate into tissues and organs.52 For let-7, this is additionally interesting because its expression is regulated by control of miRNA processing and maturation.53 Mature let-7 is readily detected in more differentiated cell types, whereas the precursor accumulates in ESCs and dedifferentiating tumor cells.51 Thus, the fact that miRNAs play important role in maintaining various cellular processes may be extrapolated to cellular reprogramming.

miRNAs in Cellular Reprogramming Takahashi and Yamanaka22 reported that differentiated cells may be reprogrammed to an embryonic-like state by ectopic expression of only four factors: Oct4, Sox2, Klf4, and c-Myc (OSKM, Yamanaka factors). Because the molecular mechanisms underlying such cellular reprogramming are on their way to be comprehensively understood, a growing

ajp.amjpathol.org

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

5

559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620

Sen and Ghatak

print & web 4C=FPO

621 Figure 3 Schematic illustration depicting basic 622 regulatory network of miRNA interactions aimed at pluripotency. Ectopic expression of miR-93, miR623 106a-b, and miR-302s induces pluripotency in 624 differentiated cells. Apart from direct expression of 625 core pluripotency factors such as Oct4, Sox2, KLF4, 626 and c-Myc, miR-300a-b-c-d, miR-200c, miR-369s are 627 capable of inducing pluripotency. Overexpression of 628 miR-290 may replace c-Myc during induction of 629 pluripotency. Expression of Let-7 miRNA enacts exit 630 of cells from the cell cycle by repressing LIN28. 631 Expression of miR-21 and miR-34 prevents conver632 sion of differentiated cells to their pluripotent state. 633 634 635 number of studies point toward the importance of miRNA in production of iPSCs, the statement should be used with 636 cell reprogramming. During the early phase of OSKMextreme caution. Harui et al59 have reported that in estab637 induced iPSC generation, miR-302s and miR-17 families lished cell lines adenoviral vectors do integrate into the 638 are the most abundantly expressed human ESC miRNAs.54 genome. Thus, it may not be justified to extrapolate these 639 frequencies to the in vivo situation.60 A detailed comparison Disruption of global miRNA biogenesis by ablation of 640 Drosha, Dicer, or Ago2 substantially reduces OSKMof nonintegrating nonviral methods for reprogramming 641 54 induced iPSC colonies in mouse embryonic fibroblasts. human somatic cells to iPSCs is provided in Table 1. ½T1 642 Compared with these methods, miRNA-based strategies to Introduction of several members of miR-290 cluster may 643 generate iPSCs without involving genomic integration is a enhance the efficiency of OSK-induced cell reprogramming, 644 much safer choice and therefore is of extraordinary theracomparable with that achieved by OSKM, demonstrating 645 646 ½F3 that miR-290s may substitute c-Myc (Figure 3). In this peutic value (Table 1).57,61e70 647 context, c-Myc binds to the promoter region of the miR648 290/miR-295 cluster.55 Similar studies with ectopic 649 Is miR-29 a Master Regulator of Tissue expression of miR-93, miR-106b, and miR-302 cluster were 650 Q8 successful in enhancing cell reprogramming. Regeneration? 651 In the absence of OSK reprogramming factors, delivery of 652 miR-302a, miR-302b, miR-302c, or miR-302d alone fails to In the mammalian system, the extracellular matrix (ECM) 653 56 induce the expression of epithelial markers. Two indeplays a critical role in tissue regeneration. The ECM de654 pendent groups have now demonstrated that lentiviral termines cell behavior by manipulating cell shape, prolif655 expression of miR-302/miR-367 cluster and a cocktail of eration, migration, differentiation, and death.71 Scarless 656 miRNAs (miR-200c, miR-302a-d, miR-369-3p, and miRregenerative healing is an intrinsic property of the fetal 657 658 369-5p) may reprogram human fibroblasts to iPSCs effiskin.72 Major differences in the ECM of the fetal and the 57 659 ciently. The cells that re-expressed human ESC factors that adult skin account for the sharp contrast in healing pheno660 feature global gene expression analogous to human ESCs type between the fetus and adult skin.73 Fetal skin ECM 661 were named miRNA-iPSCs. This constitutes the first proof of exhibits predominance of type III collagen, elevated hyal662 principle to demonstrate that iPSCs may be obtained with uronic acid, and high ratio of matrix metalloproteinase to 663 miRNAs without the need for genomic integration of foreign tissue-derived inhibitor that facilitates cell movements, 664 DNA. Genomic integration increases the chances of recipient turnover of ECM modulators, and tissue remodeling.72,74 665 chromosomal locus modification after integration, thereby One of the hallmarks of fetal skin development is the 666 raising the possibility of making the recipient genomic loci lower abundance of miRNAs caused by mechanisms that 667 unstable.58 Although adenovirus-derived vectors and Sendai are aimed at silencing PTGS during tissue development, 668 virus were proclaimed as safe nonintegrating methods for enabling expression of otherwise silenced coding genes that 669 670 671 Table 1 Comparison of Nonintegrating Nonviral Methods for the Human Somatic Cell Reprogramming to iPSCs 672 Methods Time, days Efficiency, % Success, % Pros Cons References 673 miRNA 20 0.002 60e70 Zero footprint Validated for only one cell type 57,61,62 674 mRNA 20e25 0.6e4.4 20e30 Zero footprint Cost, technically challenging, labor intensive 61e64 675 Protein 56 0.001 NA Zero footprint Technically challenging, reprograming time 61,64e66 676 Episomal 30 0.0006e0.02 80e90 Zero footprint Low efficiency 61,62,67,68 677 Minicircles 15 0.005 NA Zero footprint Validated for only one cell type 61,64,68,69 678 Piggybac 14e28 0.02 NA Zero footprint No data show excision of transposon from 61,70 679 iPSC, licensing/patent issue 680 681 iPSC, induced pluripotent stem cell; NA, Not available. 682

6

ajp.amjpathol.org

-

The American Journal of Pathology

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744

745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806

miRNA and Regenerative Medicine

Q9

Q10

drive skin development. Of such miRNAs that are silenced during fetal development, low abundance of miR-29 was reported in skin across several species.7,75,76 In humans, the miR-29 family consists of four highly conserved and closely related transcripts: hsa-miR-29a and hsa-miR-29b-1 from chromosome 7 and hsa-miR-29b-2 and hsa-miR-29c from chromosome 1. The members of miR-29 family share sequence homology at nucleotide positions 2 to 7, the seed region responsible for determination of the fate of the coding genes. PTGS of multiple ECM proteins by the miR29a-c family is one notable example of a single miRNA family that regulates a large set of functionally convergent genes such as different collagen isoforms, laminin 1, fibrilin 1, elastin, matrix metalloproteinase 2, hyaluronic acid, chondroitin sulfate, and integrin 1.77 In silico analyses have revealed that the 30 -untranslated region of the 20 collagen genes have predicted conserved binding sites for the miR29a-c family.78 This property is unique to the miR-29 family because the predicted binding sites of the collagen genes are not attributed to any sequence homology in their 30 -untranslated region.78 In adults, members of the miR-29 family cause suppression of ECM genes, resulting in tissue repair with a scar. Dysregulation of ECM caused by lower miR-29 was implicated in the development of fibrosis in several organs, including the heart,77 kidney,79 lung,80 and liver.81 Furthermore, miR-29 influences tissue differentiation and senescence.82 Elevated expression of miR-29 may attenuate the inhibitory action of transforming growth factor-b on myogenesis by targeting histone deacetylase 4, a key inhibitor of muscle differentiation.83 In adults, the elevated expression of miR-29 suppress the expression of insulin-like growth factor-1, p85a, and B-myb that lead to induction of senescence of muscle progenitor cells in vivo, thereby contributing to the development of muscle atrophy and sarcopenia.84 Suppression of miR-29 is a productive strategy to accelerate skeletal myogenesis by targeting its repressor YY1.42 In bones, miR-29 regulates osteoblast differentiation by targeting antiosteogenic factors and ECM proteins85 and Wnt signaling antagonists.86

influencing targets at several levels in a regulatory network to deliver robust biological output. To enable this, a thorough understanding of miRNAemRNA interactions is necessary. Developing hypotheses aimed at understanding the intricate regulatory networks with the help of in silico tools (eg, Targetscan, PicTar, mi-RANDA) and validating such hypotheses experimentally represent powerful approaches. High-throughput methods such as argonaute highthroughput sequencing of RNA isolated by cross-linking immunoprecipitation allow the study of actual miRNAe mRNA interactions in any specific biological context. The ability to properly evaluate the effect of an miRNA on a specific biological pathway will pave the way for the identification of appropriate candidates for therapeutic interventions. Because tissue regeneration is a multifaceted process, miRNA therapies must consider the differential requirements of the developmental program through which participating cells will temporally progress. Appropriate dosing and strategic targeting of specific cell populations are necessary to obtain optimal functional outcome. Because the incorporation of miRNA in the RISC complex is random in mammalian cells, any alteration in the level of a particular miRNA may also affect the activity of other miRNA by competing for incorporation into the miRISC. Cancer is associated with dysregulation of miRNA. Thus, any alteration in the miRNA pool, if not properly targeted to the tissue of interest, poses the risk of neoplastic outcome. Because there is clear overlap between key genes that control pluripotency and differentiation with those genes that are integral to cancer,90 it is necessary to meticulously design miRNA-based cell reprogramming strategies for the purposes of tissue repair and regeneration. Challenges related to intelligent target selection coupled with precision in dosing and spatiotemporal modulation of miRNA targets must be addressed. There are several methods to deliver miRNA to cells, including direct injection and viral and nonviral-based methods (Figure 4). ½F4

Cell Reprogramming Using miRNA

Direct injection is one of the simplest methods to deliver miRNA and antagomiRNAs. The effect is typically shortlived because the oligonucleotides are rapidly degraded by nucleases in serum and cleared by the kidneys for their inability to bind to plasma proteins. However, a single bolus intravenous injection of locked nucleic acideanti-miRNA may be stable and hence productive for several weeks.91

Approaches and Caution Synthetic locked nucleic acids conjugate oligonucleotides are useful to modify the cellular levels of any particular miRNA.87 Furthermore, overexpression at the level of primiRNA or pre-miRNA with some flanking sequence may be more biologically relevant than simply using the mature miRNA sequence because the transcript may interact with the endogenous cellular machinery, ensuring correct processing and final functionality.88 A key challenge in the modulation of miRNA signaling for tissue regeneration is the selection of the right therapeutic candidates. An appropriately chosen miRNA may modulate numerous functionally convergent coding genes.89 The miRNA is capable of

The American Journal of Pathology

-

Direct Injection

Viral Delivery Virus-based miRNA delivery to cells is one of the most widely used techniques in vitro.92 The most commonly used are adenoviral, lentiviral, and retroviral vectors. Although the viral vectors ensure long-term expression of the miRNA of interest (or control of expression with inducible promoters), this approach is not friendly for clinical translation because of

ajp.amjpathol.org

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

7

807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868

print & web 4C=FPO

869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930

Sen and Ghatak

Figure 4

A comparative analysis of current approaches. Although direct injection and delivery with the use of viral vectors is widely used in in vivo and in vitro research for deciphering molecular mechanisms and pathway validation, many disadvantages limit their use in translational research for therapeutic intervention. Lipid nanoparticles are comparatively safe, but they are rapidly eliminated by the reticuloendothelial system. Nanoelectroporation offers precise delivery, better transfection efficiency, and less cell mortality.

8

ajp.amjpathol.org

-

The American Journal of Pathology

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:47 pm  EO: AJP14_0816

931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992

993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054

miRNA and Regenerative Medicine Table 2

Key Control Modules in Cell Engineering Toolbox

Events

Rationality

Cell proliferation Cell death Cell migration

To ensure their survival on implantation To ensure their survival on implantation To redirect cellular movement toward specific signals and sites in the body where the cells should execute their action To reprogram cellecell, small moleculecell, and biologicecell communication For on-demand production and secretion of small molecules and biologics by engineered cells, extending beyond those molecules that a cell naturally makes For quantitative control of therapeutic cellular responses

Cell communication Cell secretions Cell response

their inherent toxicity and immunogenicity concerns.93 Adeno-associated viral vectors are promising for in vivo delivery of miRNAs because of their observed lower levels of toxicity and seemingly limited off-target effects.

Lipid Nanoparticles In the biology of tissue repair and regeneration, controlled release of bioactive reagents such as growth/differentiation factors or interfering RNA is critical to directing cell fate. These bioactive reagents may be loaded in lipid nanoparticles and supported by scaffolds for sustained local delivery.94 Lipid nanoparticles, being a nonviral vector, provide more flexibility in formulation design for improved uptake by the cells. In addition, by incorporating targeting ligands, it is possible to specify the route of vector administration and enhance the delivery to specific tissues or cells. Despite having several advantages, one of the main limitations of this approach is the absence of a three-dimensional mechanical support component frequently required to promote tissue repair and regeneration.94

Nano-electroporation Nanochannel electroporation may be used as a potential method to deliver precise amounts of miRNA into living cells. These devices are made of microchannels connected by nanochannels. With the use of an optical tweezer the cell to be transfected is positioned in one microchannel, and the miRNA of interest may be placed in the second microchannel.95 An intense electric field is produced by delivering a voltage pulse between the microchannels over a small area on the cell membrane, allowing a precise amount of oligonucleotide to be electrophoretically driven through the nanochannel, the cell membrane, and into the cell cytoplasm without affecting cell viability.95 Dose control is achieved by adjusting the duration and number of pulses. Although there is no reported in vivo evidence yet, such a nanochannel electroporation device has the potential to support tissue repair and regeneration in vivo. These highly sophisticated miRNA delivery methods cannot only deliver miRNA to cells of interest but can also be used for preparing cells with a desired miRNA payload for efficient cell based therapies.

The American Journal of Pathology

-

miRNA in Cell-Based Therapies Live cells may be used to package and deliver miRNAs of interest to tissue repair and regeneration. In this approach, paracrine effects of such implanted cells may guide tissue repair and regeneration. The lifetime of such cells can be controlled by natural and engineered circuits such as introduction of death signaling pathways96 or multiple auxotrophies (metabolic dependencies). The key control modules in the cell engineering toolbox is provided in Table 2. In cellbased therapies, the notion of paracrine effects that drive repair and regeneration is well established. For example, the beneficial activities of mesenchymal SCs are mediated by the secretion of pro-regenerative and anti-inflammatory growth factors in cardiac repair97 or graft-versus-host disease.98 miRNAs may be shuttled between cells packaged in small membrane-bound extracellular vesicles (EVs) (eg, exosomes, shedding vesicles, apoptotic bodies) that can then amend gene expression of their target cells.99 Thus, as a natural extension we hypothesize that cells could be used to deliver miRNAs to promote tissue regeneration. Alternatively, these EVs could be isolated and used to deliver a proregenerative miRNA payload. There is some interest in the use of EVs for miRNA delivery. Overexpression of miR146a in human embryonic kidney 293T cells elevated the miRNA level both in the transfected cells and EVs released from the cells.100 This approach of using cells as delivery vehicles is related to previous work that used genetically engineered MSCs to deliver proteins and to modulate in vivo tissue response.101 An advantage, particularly when using MSCs for this strategy, is that they naturally home to sites of disease or injury; therefore, they may provide a natural targeting mechanism to deliver miRNAs to the site of interest.102 MSCs transfected with mimics of miR-124 and miR-145 have successfully delivered these miRNAs to glioma cells. MSCs loaded with miRNAs of interest are able to change gene expression and cellular behavior in vivo.103 The importance of EVs in explaining such observations remain to be understood.

Tissue Engineering The ease by which the miRNAs may be overexpressed or repressed with sharp molecular tools renders them highly

ajp.amjpathol.org

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:48 pm  EO: AJP14_0816

9

1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 ½T21078 1079 1080 Q11 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116

1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178

Sen and Ghatak suitable for tissue reconstruction from SCs. In some instances miRNA have overcome the need for growth factors. Currently, miRNAs are widely used in the tissue engineering of cartilage, bone, and skeletal muscle. Repairing the injured cartilage is challenging, but it holds great promise for the treatment of osteoarthritis and spinal disk herniation. Although repairing major cartilage injury by using biomaterial implant and MSCs has proven to be highly successful, the phenotype of SC-derived chondrocytes is highly unstable compared with their endogenous counterpart. Under in vivo conditions, the artificial implant undergoes terminal differentiation, followed by ossification.104 Addition of miR-133 prevents such unwanted ossification.105 Furthermore, addition of miR-675 and miR-221 promotes expression of collagen II and proliferation of mature chondrocytes. In bone engineering, seeding MSCs with miR-148b mimics and miR-489 inhibitor to the osteoinductive scaffold results in increased matrix deposition and calcification.106 Application of miR-1 and miR-206 improve the differentiation potential of human satellite cells that play an important role in the tissue engineering of the skeletal muscle.107 Biomaterial scaffolds play a key role in scar formation associated with wound healing. These ECM-derived scaffolds, in addition to serving as physical support to promote tissue organization, resist aggressive wound contraction108 and scar tissue formation.109 These scaffolds also serve as vehicles to deliver bioactive substances such as cytokines,110 growth factors,109 living cells,111 and vectors for gene therapy. Attenuated levels of collagen type I and collagen type III mRNA expression were observed when primary fibroblasts were cultured with scaffolds doped with miR-29b.112 Application of this scaffold to full-thickness wounds in vivo minimized wound contraction, improved collagen type III/I ratio, and led to a dose-dependent (0.5 versus 5 mg) increased ratio of metalloproteinase-8 to tissue inhibitor of metalloproteinase-1.112 Thus, a potential of combining exogenous miRNAs with collagen scaffolds to improve extracellular matrix remodeling after injury is evident.

Conclusion With cues from tissue development process, it is now evident that miRNAs play a critical role in tissue repair and regeneration. Such influence is executed either via endogenous repair mechanisms or by directing the activity of implanted cells at the injury site. Changes in cellular phenotype as evident during dedifferentiation, transdifferentiation, or cellular reprogramming are directly linked to miRNA function within the cell. miRNA-based approaches have opened novel avenues to produce substantial amounts of iPSC clones without the need for genomic integration. Albeit in its infancy, several lines of critical evidence have shown the promise of miRNA-based tissue reparative and regenerative therapies, including tissue engineering applications. Compared with contemporary methods of ectopic gene expression currently used in regenerative

10

medicine, strategies to transiently manipulate cellular miRNA by using targeted mimics and inhibitors hold substantial promise. However, off-target effects of therapeutic miRNA remain a relevant concern. Parallel advances in novel delivery technologies such as lipid nanoparticles and nanochannel electroporation are beginning to address some of the key stability and toxicity issues, thus making miRNA-based therapies more translationally relevant. Currently, there are 152 open clinical studies on miRNAs worldwide (Clinical Trails.gov, https://clinicaltrials.gov/ct2/results?termZmiRNAþandþ Tissueþrepair, last accessed March 30, 2015). Of these, six trials are on miRNA and tissue repair, of which three are based in the United States. The race to perfect miRNA-based therapies is one of high stakes and has the clear potential to transform reparative and regenerative medicine.

References 1. Roy S, Khanna S, Rink C, Biswas S, Sen CK: Characterization of the acute temporal changes in excisional murine cutaneous wound inflammation by screening of the wound-edge transcriptome. Physiol Genomics 2008, 34:162e184 2. Thatcher EJ, Patton JG: Small RNAs have a big impact on regeneration. RNA Biol 2010, 7:333e338 3. Nowak R: Mining treasures from ‘junk DNA’. Science 1994, 263: 608e610 4. Brodersen P, Voinnet O: Revisiting the principles of microRNA target recognition and mode of action. Nat Rev Mol Cell Biol 2009, 10:141e148 5. Ha M, Kim VN: Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol 2014, 15:509e524 6. Schwartzfarb E, Kirsner RS: Understanding scarring: scarless fetal wound healing as a model (comment). J Invest Dermatol 2012, 132: 260 7. Cheng J, Yu H, Deng S, Shen G: MicroRNA profiling in mid- and late-gestational fetal skin: implication for scarless wound healing. Tohoku J Exp Med 2010, 221:203e209 8. Suh N, Baehner L, Moltzahn F, Melton C, Shenoy A, Chen J, Blelloch R: MicroRNA function is globally suppressed in mouse oocytes and early embryos. Curr Biol 2010, 20:271e277 9. Flemr M, Moravec M, Libova V, Sedlacek R, Svoboda P: Lin28a is dormant, functional, and dispensable during mouse oocyte-to-embryo transition. Biol Reprod 2014, 90:131 10. Heo I, Joo C, Kim Y-K, Ha M, Yoon MJ, Cho J, Yeom KH, Han J, Kim VN: TUT4 in concert with Lin28 suppresses MicroRNA biogenesis through pre-microRNA uridylation. Cell 2009, 138: 696e708 11. Wang J, Conboy I: Embryonic vs. adult myogenesis: challenging the ‘regeneration recapitulates development’ paradigm. J Mol Cell Biol 2010, 2:1e4 12. Yunta M, Nieto-Diaz M, Esteban FJ, Caballero-Lopez M, NavarroRuiz R, Reigada D, Pita-Thomas DW, del Aguila A, MunozGaldeano T, Maza RM: MicroRNA dysregulation in the spinal cord following traumatic injury. PLoS One 2012, 7:e34534 13. Strickland ER, Woller SA, Garraway SM, Hook MA, Grau JW, Miranda RC: Regulatory effects of intermittent noxious stimulation on spinal cord injury-sensitive microRNAs and their presumptive targets following spinal cord contusion. Front Neural Circuits 2014, 8:117 14. Salehi S, Brereton HC, Arno MJ, Darling D, Quaglia A, O’Grady J, Heaton N, Aluvihare VR: Human liver regeneration is characterized by the coordinated expression of distinct microRNA governing cell cycle fate. Am J Transplant 2013, 13:1282e1295

ajp.amjpathol.org

-

The American Journal of Pathology

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:48 pm  EO: AJP14_0816

1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240

1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302

miRNA and Regenerative Medicine

Q12

15. Cheung TH, Rando TA: Molecular regulation of stem cell quiescence. Nat Rev Mol Cell Biol 2013, 14:329e340 16. Jopling C, Boue S, Izpisua Belmonte JC: Dedifferentiation, transdifferentiation and reprogramming: three routes to regeneration. Nat Rev Mol Cell Biol 2011, 12:79e89 17. Odelberg SJ: Cellular plasticity in vertebrate regeneration. Anat Rec B New Anat 2005, 287:25e35 18. Echeverri K, Clarke JD, Tanaka EM: In vivo imaging indicates muscle fiber dedifferentiation is a major contributor to the regenerating tail blastema. Dev Biol 2001, 236:151e164 19. Cai S, Fu X, Sheng Z: Dedifferentiation: a new approach in stem cell research. BioScience 2007, 57:655e662 20. Eguizabal C, Montserrat N, Veiga A, Izpisua Belmonte JC: Dedifferentiation, transdifferentiation, and reprogramming: future directions in regenerative medicine. Semin Reprod Med 2013, 31: 82e94 21. Cowan CA, Atienza J, Melton DA, Eggan K: Nuclear reprogramming of somatic cells after fusion with human embryonic stem cells. Science 2005, 309:1369e1373 22. Takahashi K, Yamanaka S: Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126:663e676 23. Graf T: Historical origins of transdifferentiation and reprogramming. Cell Stem Cell 2011, 9:504e516 24. Takahashi K: Cellular reprogrammingelowering gravity on Waddington’s epigenetic landscape. J Cell Sci 2012, 125:2553e2560 25. Ieda M: Direct reprogramming into desired cell types by defined factors. Keio J Med 2013, 62:74e82 26. Kelaini S, Cochrane A, Margariti A: Direct reprogramming of adult cells: avoiding the pluripotent state. Stem Cells Cloning 2014, 7:19e29 27. Patapoutian A, Wold BJ, Wagner RA: Evidence for developmentally programmed transdifferentiation in mouse esophageal muscle. Science 1995, 270:1818e1821 28. Poss KD: Advances in understanding tissue regenerative capacity and mechanisms in animals. Nat Rev Genet 2010, 11:710e722 29. Frith JE, Porrello ER, Cooper-White JJ: Concise review: new frontiers in microRNA-based tissue regeneration. Stem Cells Transl Med 2014, 3:969e976 30. Gusev Y: Computational methods for analysis of cellular functions and pathways collectively targeted by differentially expressed microRNA. Methods 2008, 44:61e72 31. Shyh-Chang N, Zhu H, Yvanka de Soysa T, Shinoda G, Seligson MT, Tsanov KM, Nguyen L, Asara JM, Cantley LC, Daley GQ: Lin28 enhances tissue repair by reprogramming cellular metabolism. Cell 2013, 155:778e792 32. Choudhury NR, de Lima Alves F, de Andres-Aguayo L, Graf T, Caceres JF, Rappsilber J, Michlewski G: Tissue-specific control of brain-enriched miR-7 biogenesis. Genes Dev 2013, 27:24e38 33. Davis BN, Hata A: Regulation of MicroRNA Biogenesis: a miRiad of mechanisms. Cell Commun Signal 2009, 7:18 34. Winter J, Jung S, Keller S, Gregory RI, Diederichs S: Many roads to maturity: microRNA biogenesis pathways and their regulation. Nat Cell Biol 2009, 11:228e234 35. O’Donnell KA, Wentzel EA, Zeller KI, Dang CV, Mendell JT: c-Myc-regulated microRNAs modulate E2F1 expression. Nature 2005, 435:839e843 36. Krol J, Loedige I, Filipowicz W: The widespread regulation of microRNA biogenesis, function and decay. Nat Rev Genet 2010, 11: 597e610 37. Ghatak S, Sen CK: MicroRNA biogenesis in regenerative medicine. MicroRNA in regenerative medicine. Edited by Sen CK. Oxford: Academic Press, 2015. pp. 3e46 38. Ventura A, Jacks T: MicroRNAs and cancer: short RNAs go a long way. Cell 2009, 136:586e591 39. van Rooij E, Sutherland LB, Qi X, Richardson JA, Hill J, Olson EN: Control of stress-dependent cardiac growth and gene expression by a microRNA. Science 2007, 316:575e579

The American Journal of Pathology

-

40. Kai ZS, Pasquinelli AE: MicroRNA assassins: factors that regulate the disappearance of miRNAs. Nat Struct Mol Biol 2010, 17:5e10 41. Diederichs S, Haber DA: Dual role for argonautes in microRNA processing and posttranscriptional regulation of microRNA expression. Cell 2007, 131:1097e1108 42. Wang H, Garzon R, Sun H, Ladner KJ, Singh R, Dahlman J, Cheng A, Hall BM, Qualman SJ, Chandler DS, Croce CM, Guttridge DC: NF-kappaB-YY1-miR-29 regulatory circuitry in skeletal myogenesis and rhabdomyosarcoma. Cancer Cell 2008, 14:369e381 43. Jouneau A, Ciaudo C, Sismeiro O, Brochard V, Jouneau L, Vandormael-Pournin S, Coppee JY, Zhou Q, Heard E, Antoniewski C, Cohen-Tannoudji M: Naive and primed murine pluripotent stem cells have distinct miRNA expression profiles. RNA 2012, 18:253e264 44. Wilson KD, Venkatasubrahmanyam S, Jia F, Sun N, Butte AJ, Wu JC: MicroRNA profiling of human-induced pluripotent stem cells. Stem Cells Dev 2009, 18:749e758 45. Zhao B, Yang D, Jiang J, Li J, Fan C, Huang M, Fan Y, Jin Y, Jin Y: Genome-wide mapping of miRNAs expressed in embryonic stem cells and pluripotent stem cells generated by different reprogramming strategies. BMC Genomics 2014, 15:488 46. Hatfield SD, Shcherbata HR, Fischer KA, Nakahara K, Carthew RW, Ruohola-Baker H: Stem cell division is regulated by the microRNA pathway. Nature 2005, 435:974e978 47. Bernstein E, Kim SY, Carmell MA, Murchison EP, Alcorn H, Li MZ, Mills AA, Elledge SJ, Anderson KV, Hannon GJ: Dicer is essential for mouse development. Nat Genet 2003, 35:215e217 48. Murchison EP, Partridge JF, Tam OH, Cheloufi S, Hannon GJ: Characterization of Dicer-deficient murine embryonic stem cells. Proc Natl Acad Sci U S A 2005, 102:12135e12140 49. Wang Y, Medvid R, Melton C, Jaenisch R, Blelloch R: DGCR8 is essential for microRNA biogenesis and silencing of embryonic stem cell self-renewal. Nat Genet 2007, 39:380e385 50. Wang Y, Sheng G, Juranek S, Tuschl T, Patel DJ: Structure of the guide-strand-containing argonaute silencing complex. Nature 2008, 456:209e213 51. Melton C, Judson RL, Blelloch R: Opposing microRNA families regulate self-renewal in mouse embryonic stem cells. Nature 2010, 463:621e626 52. Wienholds E, Plasterk RH: MicroRNA function in animal development. FEBS Lett 2005, 579:5911e5922 53. Lehrbach NJ, Armisen J, Lightfoot HL, Murfitt KJ, Bugaut A, Balasubramanian S, Miska EA: LIN-28 and the poly(U) polymerase PUP-2 regulate let-7 microRNA processing in Caenorhabditis elegans. Nat Struct Mol Biol 2009, 16:1016e1020 54. Li Z, Yang CS, Nakashima K, Rana TM: Small RNA-mediated regulation of iPS cell generation. EMBO J 2011, 30:823e834 55. Judson RL, Babiarz JE, Venere M, Blelloch R: Embryonic stem cellspecific microRNAs promote induced pluripotency. Nat Biotechnol 2009, 27:459e461 56. Subramanyam D, Lamouille S, Judson RL, Liu JY, Bucay N, Derynck R, Blelloch R: Multiple targets of miR-302 and miR-372 promote reprogramming of human fibroblasts to induced pluripotent stem cells. Nat Biotechnol 2011, 29:443e448 57. Miyoshi N, Ishii H, Nagano H, Haraguchi N, Dewi DL, Kano Y, Nishikawa S, Tanemura M, Mimori K, Tanaka F, Saito T, Nishimura J, Takemasa I, Mizushima T, Ikeda M, Yamamoto H, Sekimoto M, Doki Y, Mori M: Reprogramming of mouse and human cells to pluripotency using mature microRNAs. Cell Stem Cell 2011, 8:633e638 58. Wurtele H, Little KC, Chartrand P: Illegitimate DNA integration in mammalian cells. Gene Ther 2003, 10:1791e1799 59. Harui A, Suzuki S, Kochanek S, Mitani K: Frequency and stability of chromosomal integration of adenovirus vectors. J Virol 1999, 73: 6141e6146 60. Schagen FH, Rademaker HJ, Fallaux FJ, Hoeben RC: Insertion vectors for gene therapy. Gene Ther 2000, 7:271e272

ajp.amjpathol.org

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:48 pm  EO: AJP14_0816

11

1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364

1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426

Sen and Ghatak 61. Rao MS, Malik N: Assessing iPSC reprogramming methods for their suitability in translational medicine. J Cell Biochem 2012, 113: 3061e3068 62. Schlaeger TM, Daheron L, Brickler TR, Entwisle S, Chan K, Cianci A, DeVine A, Ettenger A, Fitzgerald K, Godfrey M, Gupta D, McPherson J, Malwadkar P, Gupta M, Bell B, Doi A, Jung N, Li X, Lynes MS, Brookes E, Cherry AB, Demirbas D, Tsankov AM, Zon LI, Rubin LL, Feinberg AP, Meissner A, Cowan CA, Daley GQ: A comparison of non-integrating reprogramming methods. Nat Biotechnol 2015, 33:58e63 63. Warren L, Ni Y, Wang J, Guo X: Feeder-free derivation of human induced pluripotent stem cells with messenger RNA. Sci Rep 2012, 2: 657 64. Li J, Song W, Pan G, Zhou J: Advances in understanding the cell types and approaches used for generating induced pluripotent stem cells. J Hematol Oncol 2014, 7:50 65. Stadtfeld M, Hochedlinger K: Induced pluripotency: history, mechanisms, and applications. Genes Dev 2010, 24:2239e2263 66. Zhou H, Wu S, Joo JY, Zhu S, Han DW, Lin T, Trauger S, Bien G, Yao S, Zhu Y, Siuzdak G, Scholer HR, Duan L, Ding S: Generation of induced pluripotent stem cells using recombinant proteins. Cell Stem Cell 2009, 4:381e384 67. Yu J, Hu K, Smuga-Otto K, Tian S, Stewart R, Slukvin II, Thomson JA: Human induced pluripotent stem cells free of vector and transgene sequences. Science 2009, 324:797e801 68. O’Doherty R, Greiser U, Wang W: Nonviral methods for inducing pluripotency to cells. Biomed Res Int 2013, 2013:705902 69. Jia F, Wilson KD, Sun N, Gupta DM, Huang M, Li Z, Panetta NJ, Chen ZY, Robbins RC, Kay MA, Longaker MT, Wu JC: A nonviral minicircle vector for deriving human iPS cells. Nat Methods 2010, 7: 197e199 70. Woltjen K, Michael IP, Mohseni P, Desai R, Mileikovsky M, Hamalainen R, Cowling R, Wang W, Liu P, Gertsenstein M, Kaji K, Sung HK, Nagy A: piggyBac transposition reprograms fibroblasts to induced pluripotent stem cells. Nature 2009, 458:766e770 71. Daley WP, Peters SB, Larsen M: Extracellular matrix dynamics in development and regenerative medicine. J Cell Sci 2008, 121:255e264 72. Larson BJ, Longaker MT, Lorenz HP: Scarless fetal wound healing: a basic science review. Plast Reconstr Surg 2010, 126:1172e1180 73. Tang QM, Chen JL, Shen WL, Yin Z, Liu HH, Fang Z, Heng BC, Ouyang HW, Chen X: Fetal and adult fibroblasts display intrinsic differences in tendon tissue engineering and regeneration. Sci Rep 2014, 4:5515 74. Longaker MT, Whitby DJ, Ferguson MW, Lorenz HP, Harrison MR, Adzick NS: Adult skin wounds in the fetal environment heal with scar formation. Ann Surg 1994, 219:65e72 75. Huang TH, Zhu MJ, Li XY, Zhao SH: Discovery of porcine microRNAs and profiling from skeletal muscle tissues during development. PLoS One 2008, 3:e3225 76. Wessels JM, Edwards AK, Khalaj K, Kridli RT, Bidarimath M, Tayade C: The microRNAome of pregnancy: deciphering miRNA networks at the maternal-fetal interface. PLoS One 2013, 8:e72264 77. van Rooij E, Sutherland LB, Thatcher JE, DiMaio JM, Naseem RH, Marshall WS, Hill JA, Olson EN: Dysregulation of microRNAs after myocardial infarction reveals a role of miR-29 in cardiac fibrosis. Proc Natl Acad Sci U S A 2008, 105:13027e13032 78. Liu Y, Taylor NE, Lu L, Usa K, Cowley AW Jr, Ferreri NR, Yeo NC, Liang M: Renal medullary microRNAs in Dahl salt-sensitive rats: miR-29b regulates several collagens and related genes. Hypertension 2010, 55:974e982 79. Qin W, Chung AC, Huang XR, Meng XM, Hui DS, Yu CM, Sung JJ, Lan HY: TGF-beta/Smad3 signaling promotes renal fibrosis by inhibiting miR-29. J Am Soc Nephrol 2011, 22:1462e1474 80. Pandit KV, Milosevic J, Kaminski N: MicroRNAs in idiopathic pulmonary fibrosis. Transl Res 2011, 157:191e199 81. Roderburg C, Urban GW, Bettermann K, Vucur M, Zimmermann H, Schmidt S, Janssen J, Koppe C, Knolle P, Castoldi M, Tacke F,

12

82. 83.

84.

85.

86.

87.

88.

89.

90.

91.

92. 93. 94.

95.

96.

97.

98.

99.

Trautwein C, Luedde T: Micro-RNA profiling reveals a role for miR29 in human and murine liver fibrosis. Hepatology 2011, 53:209e218 Schraml E, Grillari J: From cellular senescence to age-associated diseases: the miRNA connection. Longev Healthspan 2012, 1:10 Winbanks CE, Wang B, Beyer C, Koh P, White L, Kantharidis P, Gregorevic P: TGF-beta regulates miR-206 and miR-29 to control myogenic differentiation through regulation of HDAC4. J Biol Chem 2011, 286:13805e13814 Hu Z, Klein JD, Mitch WE, Zhang L, Martinez I, Wang XH: MicroRNA-29 induces cellular senescence in aging muscle through multiple signaling pathways. Aging (Albany NY) 2014, 6:160e175 Li Z, Hassan MQ, Jafferji M, Aqeilan RI, Garzon R, Croce CM, van Wijnen AJ, Stein JL, Stein GS, Lian JB: Biological functions of miR29b contribute to positive regulation of osteoblast differentiation. J Biol Chem 2009, 284:15676e15684 Kapinas K, Kessler C, Ricks T, Gronowicz G, Delany AM: miR-29 modulates Wnt signaling in human osteoblasts through a positive feedback loop. J Biol Chem 2010, 285:25221e25231 Ameres SL, Horwich MD, Hung JH, Xu J, Ghildiyal M, Weng Z, Zamore PD: Target RNA-directed trimming and tailing of small silencing RNAs. Science 2010, 328:1534e1539 Park JE, Heo I, Tian Y, Simanshu DK, Chang H, Jee D, Patel DJ, Kim VN: Dicer recognizes the 5’ end of RNA for efficient and accurate processing. Nature 2011, 475:201e205 Lim LP, Lau NC, Garrett-Engele P, Grimson A, Schelter JM, Castle J, Bartel DP, Linsley PS, Johnson JM: Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 2005, 433:769e773 Muller FJ, Laurent LC, Kostka D, Ulitsky I, Williams R, Lu C, Park IH, Rao MS, Shamir R, Schwartz PH, Schmidt NO, Loring JF: Regulatory networks define phenotypic classes of human stem cell lines. Nature 2008, 455:401e405 Hullinger TG, Montgomery RL, Seto AG, Dickinson BA, Semus HM, Lynch JM, Dalby CM, Robinson K, Stack C, Latimer PA, Hare JM, Olson EN, van Rooij E: Inhibition of miR-15 protects against cardiac ischemic injury. Circ Res 2012, 110:71e81 van Rooij E, Purcell AL, Levin AA: Developing microRNA therapeutics. Circ Res 2012, 110:496e507 Liu YP, Berkhout B: miRNA cassettes in viral vectors: problems and solutions. Biochim Biophys Acta 2011, 1809:732e745 Santo VE, Gomes ME, Mano JF, Reis RL: From nano- to macroscale: nanotechnology approaches for spatially controlled delivery of bioactive factors for bone and cartilage engineering. Nanomedicine (Lond) 2012, 7:1045e1066 Boukany PE, Morss A, Liao WC, Henslee B, Jung H, Zhang X, Yu B, Wang X, Wu Y, Li L, Gao K, Hu X, Zhao X, Hemminger O, Lu W, Lafyatis GP, Lee LJ: Nanochannel electroporation delivers precise amounts of biomolecules into living cells. Nat Nanotechnol 2011, 6:747e754 Di Stasi A, Tey SK, Dotti G, Fujita Y, Kennedy-Nasser A, Martinez C, Straathof K, Liu E, Durett AG, Grilley B, Liu H, Cruz CR, Savoldo B, Gee AP, Schindler J, Krance RA, Heslop HE, Spencer DM, Rooney CM, Brenner MK: Inducible apoptosis as a safety switch for adoptive cell therapy. N Engl J Med 2011, 365:1673e1683 Leskela HV, Satta J, Oiva J, Eriksen H, Juha R, Korkiamaki P, Ivaska KK, Soini Y, Lehenkari P: Calcification and cellularity in human aortic heart valve tissue determine the differentiation of bonemarrow-derived cells. J Mol Cell Cardiol 2006, 41:642e649 Hoogduijn MJ, Popp F, Verbeek R, Masoodi M, Nicolaou A, Baan C, Dahlke MH: The immunomodulatory properties of mesenchymal stem cells and their use for immunotherapy. Int Immunopharmacol 2010, 10:1496e1500 Zhang Y, Liu D, Chen X, Li J, Li L, Bian Z, Sun F, Lu J, Yin Y, Cai X, Sun Q, Wang K, Ba Y, Wang Q, Wang D, Yang J, Liu P, Xu T, Yan Q, Zhang J, Zen K, Zhang CY: Secreted monocytic miR150 enhances targeted endothelial cell migration. Mol Cell 2010, 39: 133e144

ajp.amjpathol.org

-

The American Journal of Pathology

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:48 pm  EO: AJP14_0816

1427 1428 1429 1430 1431 1432 1433 1434 1435 1436 1437 1438 1439 1440 1441 1442 1443 1444 1445 1446 1447 1448 1449 1450 1451 1452 1453 1454 1455 1456 1457 1458 1459 1460 1461 1462 1463 1464 1465 1466 1467 1468 1469 1470 1471 1472 1473 1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 1484 1485 1486 1487 1488

1489 1490 1491 1492 1493 1494 1495 1496 1497 1498 1499 1500 1501 1502 1503 1504 1505 1506 1507 1508 1509 1510 1511 1512 1513 1514 1515

miRNA and Regenerative Medicine 100. Guduric-Fuchs J, O’Connor A, Camp B, O’Neill CL, Medina RJ, Simpson DA: Selective extracellular vesicle-mediated export of an overlapping set of microRNAs from multiple cell types. BMC Genomics 2012, 13:357 101. Haider H, Jiang S, Idris NM, Ashraf M: IGF-1-overexpressing mesenchymal stem cells accelerate bone marrow stem cell mobilization via paracrine activation of SDF-1alpha/CXCR4 signaling to promote myocardial repair. Circ Res 2008, 103:1300e1308 102. Chapel A, Bertho JM, Bensidhoum M, Fouillard L, Young RG, Frick J, Demarquay C, Cuvelier F, Mathieu E, Trompier F, Dudoignon N, Germain C, Mazurier C, Aigueperse J, Borneman J, Gorin NC, Gourmelon P, Thierry D: Mesenchymal stem cells home to injured tissues when co-infused with hematopoietic cells to treat a radiation-induced multi-organ failure syndrome. J Gene Med 2003, 5: 1028e1038 103. Lee HK, Finniss S, Cazacu S, Bucris E, Ziv-Av A, Xiang C, Bobbitt K, Rempel SA, Hasselbach L, Mikkelsen T, Slavin S, Brodie C: Mesenchymal stem cells deliver synthetic microRNA mimics to glioma cells and glioma stem cells and inhibit their cell migration and self-renewal. Oncotarget 2013, 4:346e361 104. Hollander AP, Dickinson SC, Kafienah W: Stem cells and cartilage development: complexities of a simple tissue. Stem Cells 2010, 28: 1992e1996 105. Zhang Y, Xie RL, Croce CM, Stein JL, Lian JB, van Wijnen AJ, Stein GS: A program of microRNAs controls osteogenic lineage progression by targeting transcription factor Runx2. Proc Natl Acad Sci U S A 2011, 108:9863e9868

The American Journal of Pathology

-

106. Mariner PD, Johannesen E, Anseth KS: Manipulation of miRNA activity accelerates osteogenic differentiation of hMSCs in engineered 3D scaffolds. J Tissue Eng Regen Med 2012, 6:314e324 107. Koning M, Werker PM, van der Schaft DW, Bank RA, Harmsen MC: MicroRNA-1 and microRNA-206 improve differentiation potential of human satellite cells: a novel approach for tissue engineering of skeletal muscle. Tissue Eng Part A 2012, 18:889e898 108. Yannas IV, Burke JF, Orgill DP, Skrabut EM: Wound tissue can utilize a polymeric template to synthesize a functional extension of skin. Science 1982, 215:174e176 109. Formiga FR, Pelacho B, Garbayo E, Abizanda G, Gavira JJ, SimonYarza T, Mazo M, Tamayo E, Jauquicoa C, Ortiz-de-Solorzano C, Prosper F, Blanco-Prieto MJ: Sustained release of VEGF through PLGA microparticles improves vasculogenesis and tissue remodeling in an acute myocardial ischemia-reperfusion model. J Control Release 2010, 147:30e37 110. Ghahary A, Tredget EE, Shen Q, Kilani RT, Scott PG, Takeuchi M: Liposome associated interferon-alpha-2b functions as an antifibrogenic factor in dermal wounds in the guinea pig. Mol Cell Biochem 2000, 208:129e137 111. Hoban DB, Newland B, Moloney TC, Howard L, Pandit A, Dowd E: The reduction in immunogenicity of neurotrophin overexpressing stem cells after intra-striatal transplantation by encapsulation in an in situ gelling collagen hydrogel. Biomaterials 2013, 34:9420e9429 112. Monaghan M, Browne S, Schenke-Layland K, Pandit A: A collagenbased scaffold delivering exogenous microrna-29B to modulate extracellular matrix remodeling. Mol Ther 2014, 22:786e796

ajp.amjpathol.org

REV 5.2.0 DTD  AJPA2054_proof  5 June 2015  7:48 pm  EO: AJP14_0816

13

1516 1517 1518 1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 1535 1536 1537 1538 1539 1540 1541 1542

miRNA control of tissue repair and regeneration.

Tissue repair and regeneration rely on the function of miRNA, molecular silencers that enact post-transcriptional gene silencing of coding genes. Disr...
2MB Sizes 3 Downloads 11 Views