[go: up one dir, main page]

Skip to main content
Elsevier Sponsored Documents logoLink to Elsevier Sponsored Documents
. 2018 May 29;23(9):2732–2743. doi: 10.1016/j.celrep.2018.04.094

Human Pluripotent Stem-Cell-Derived Cortical Neurons Integrate Functionally into the Lesioned Adult Murine Visual Cortex in an Area-Specific Way

Ira Espuny-Camacho 1,2,3,4,5, Kimmo A Michelsen 1, Daniele Linaro 1,2,3,6, Angéline Bilheu 1, Sandra Acosta-Verdugo 1,11, Adèle Herpoel 1, Michele Giugliano 6,7,8, Afsaneh Gaillard 9, Pierre Vanderhaeghen 1,2,3,10,12,
PMCID: PMC5990494  PMID: 29847802

Summary

The transplantation of pluripotent stem-cell-derived neurons constitutes a promising avenue for the treatment of several brain diseases. However, their potential for the repair of the cerebral cortex remains unclear, given its complexity and neuronal diversity. Here, we show that human visual cortical cells differentiated from embryonic stem cells can be transplanted and can integrate successfully into the lesioned mouse adult visual cortex. The transplanted human neurons expressed the appropriate repertoire of markers of six cortical layers, projected axons to specific visual cortical targets, and were synaptically active within the adult brain. Moreover, transplant maturation and integration were much less efficient following transplantation into the lesioned motor cortex, as previously observed for transplanted mouse cortical neurons. These data constitute an important milestone for the potential use of human PSC-derived cortical cells for the reassembly of cortical circuits and emphasize the importance of cortical areal identity for successful transplantation.

Keywords: brain repair, cerebral cortex, pluripotent stem cells, brain transplantation, neural wiring

Graphical Abstract

graphic file with name fx1.jpg

Highlights

  • Human PSC-derived cortical neurons efficiently integrate into the adult mouse brain

  • PSC-derived human neurons reestablish axonal pathways in the lesioned adult cortex

  • Restoration of cortical pathways requires a donor and recipient area-identity match


Espuny-Camacho et al. show that transplanted ESC-derived human cortical neurons integrate functionally into the lesioned adult mouse brain. Transplanted neurons display visual cortical identity and show specific restoration of damaged cortical pathways following transplantation into the visual but not the motor cortex, suggesting the importance of areal-identity match for successful cortical repair.

Introduction

Neurodegenerative diseases, brain injury, and stroke are major causes of neuronal cell loss in the adult brain. Together with in vivo reprogramming (Arlotta and Berninger, 2014, Gascón et al., 2017), transplantation of neural cells is a promising avenue for the replacement of lost neurons and damaged neural circuits (Barker et al., 2015, Gage and Temple, 2013, Goldman, 2016, Tabar and Studer, 2014). An ideal cell transplant approach should lead to the replacement of the lost neuronal subtypes and neural circuits in a comprehensive and specific way. Compared, for instance, with the replacement of substantia nigra neurons in Parkinson disease, this seems to be particularly challenging for the cerebral cortex, both conceptually and technically, given its unparalleled neuronal diversity, complex connectivity, and function. However, several independent studies have demonstrated the potential of transplanted mouse cortical cells, whether derived from mouse embryonic tissue or embryonic stem cells, for the replacement of lost neurons following a cortical lesion in the adult mouse (Falkner et al., 2016, Gaillard et al., 2007, Michelsen et al., 2015, Péron et al., 2017). Such transplanted cells display specific patterns of synaptic inputs, making them function in a highly similar way to endogenous neurons (Falkner et al., 2016). They also present surprisingly high levels of specificity in terms of cortex areal identity. For instance, replacement of lesioned motor cortex with embryonic motor cortex tissue (Gaillard et al., 2007) can lead to the selective re-establishment of motor axonal pathways, but the use of transplants derived from the visual cortex does not lead to any efficient repair. Similarly, the transplantation of mouse visual cortex-like cells derived from embryonic stem cells (ESCs) (Gaspard et al., 2008) can lead to the efficient replacement of lesioned axonal pathways of the visual cortex but not the motor cortex (Michelsen et al., 2015). Thus, successful transplantation in these cases was achieved only if there was a match between the areal identity (frontal versus occipital) of the lesioned and the transplanted cortical cells (Michelsen et al., 2015).

From a translational viewpoint, the ability of human pluripotent stem cells (PSCs) to contribute to the repair of cortical lesions is of paramount importance, given the limited availability of fetal material. We and others have shown that human ESCs and induced PSCs (iPSC) can be differentiated into pyramidal glutamatergic cortical neurons from all cortical layers (van den Ameele et al., 2014, Eiraku et al., 2008, Espuny-Camacho et al., 2013, Shi et al., 2012). The default differentiation of human ESCs and iPSCs cultured in the absence of any morphogens but in the presence of Noggin for human ectoderm acquisition recapitulates several main hallmarks of in vivo corticogenesis, such as temporal patterning in vitro (Espuny-Camacho et al., 2013). Moreover, upon transplantation into newborn recipient mice, the cortical neurons send specific patterns of cortical axonal projections at far distances from the graft location and are integrated in mouse neuronal networks (Espuny-Camacho et al., 2013).

Human ESC-derived neurons were recently shown to establish functional synapses following transplantation into damaged cortical areas in the adult mouse (Tornero et al., 2013, Tornero et al., 2017), but the specificity of the cortical fate of the transplanted cells and of their axonal input/output remains to be explored. Here, we investigated whether and how human ESC-derived cortical neurons corresponding mostly to a visual-like identity (Espuny-Camacho et al., 2013) transplanted into the lesioned adult murine cortex could integrate into the lesioned area and participate in the reassembly of cortical circuits. We found that the human neurons transplanted into the lesioned cortex acquire the molecular and axonal projection characteristics of all six cortical layers, while displaying a high degree of visual areal specificity. They also display features of functional neurons in terms of synaptic connectivity. The success of transplantation is highly dependent on a match of (visual) areal identity between the lesioned and the transplanted neurons. These results imply that human ESC-derived cortical neurons also can efficiently differentiate and establish cortical-specific neural connections in the less permissive environment of the adult lesioned brain.

Results

Human PSC-Derived Cortical Neurons Integrate into the Adult Lesioned Murine Cortex following Transplantation

To determine whether transplanted human PSC-derived cortical neurons can integrate in the lesioned adult mouse cortex, we used focal cortical lesioning mediated by injection of neurotoxic ibotenic acid, as used for the transplantation of mouse ESC-derived cortical neurons (Michelsen et al., 2015) (Figure 1A). Three days later, we transplanted cortical-like cells differentiated from GFP-expressing human ESCs (Espuny-Camacho et al., 2013) into the same lesioned area. This in vitro corticogenesis model generates human neurons from all different cortical layers in a temporal manner. We showed that such human cortical cells transplanted into immunodeficient newborn mice generate neurons with a visual-like areal identity at early stages that efficiently integrate into mouse neural networks in vivo (Espuny-Camacho et al., 2013). We, therefore, first focused on transplantation into the lesioned visual cortex, as performed previously for mouse ESCs (Michelsen et al., 2015). Following 3 weeks of in vitro differentiation, GFP-expressing human ESC-derived cortical cells were dissociated and transplanted in the lesioned visual cortex of immunodeficient adult mice. The presence of the human transplants was assessed by GFP immunostainings at 1–2 and 6 months post-transplantation (MPT) and found in 11 of 21 (52%) animals that underwent transplantation. As expected, transplants were located in the cortex or in the underlying white matter (Figures 1B, 2A, S1A, S1B, S2E, and S3A). We then compared the transplants at 2 and 6 MPT for maturation and integration.

Figure 1.

Figure 1

Grafted Human ESC-Derived Neurons in the Adult Murine Cortex Express Telencephalic and Cortical Layer-Specific Markers

(A) Experimental scheme of the lesioning/transplantation experiments in the visual cortex.

(B) Immunofluorescence image showing the GFP+ transplant (green) and MAP2 (red) staining of the visual cortex and the transplant.

(C–J) Immunofluorescence images showing the GFP+ graft (green), beta 3 class III tubulin (blue), and the human markers HuNuclei and HuNCAM (red; C and D); the telencephalic marker FoxG1 (red; E); and the cortical layer markers Tbr1 (F), Foxp2 (G), CTIP2 (H), Brn2 (I), Satb2 (J) (red) at 6 MPT.

Scale bars: 500 μm (B), 50 μm (C–J).

Figure 2.

Figure 2

Human ESC-Derived Cortical Neurons Send Numerous Projections to Specific Cortical Targets following Transplantation into the Adult Murine Visual Cortex

(A) Camera lucida drawing representing the transplant located into the visual cortex and its projections at 6MPT.

(B–G) GFP+ axonal projections are detected by immunofluorescence in different regions of the mouse brain at 6MPT. Fibers are detected along the corpus callosum (CC) (B) and external capsule (EC) (C). Fibers are also detected in the striatum (CPu) (D), visual cortex (E), thalamus (LD) (F), and superior colliculus (SC) (G).

(F and G) Composite picture views made from the stacked confocal images showing a broad area from thalamus (F) and superior colliculus, midbrain (G). Ctx, cortex; CPu, caudate putamen; LV, lateral ventricle; Rt, reticulate thalamus; LD, laterodorsal thalamic nucleus; VB, ventrobasal complex; Po, posterior thalamic nuclear group.

Scale bars: 2 mm (A), 100 mm (B–G).

At 2 MPT, transplants were still rather immature, with the presence of neural progenitors, displayed in rosette-like structures and expressing markers of proliferation (Figures S1A, S1C, and S1E), while the expression of mature neuronal markers such as VGlut1 and NeuN remained low (Figures S1G and S1I). In contrast, at 6 MPT, the transplants, found in 58% of animals, presented a more mature pattern, with far fewer or no rosette-like structures and with few proliferative cells or progenitors (Figures S1B, S1D, and S1F). GFP+-transplanted cells were positive for the human specific markers HuNuclei and human neural cell adhesion molecule (NCAM), and broadly expressed the telencephalic marker FoxG1, the neuronal marker beta 3 class III tubulin (Figures 1C–1E), and mature neuronal markers such as VGlut1, NeuN, and MAP2 (Figures S1H, S1J, and 1B). Moreover, these neurons expressed a diverse repertoire of layer-specific cortical pyramidal neuronal markers, including deep layer VI; V markers Tbr1, Foxp2, and Ctip2; and upper layer IV-II markers Satb2 and Brn2 (Figures 1F–1J and S2I–S2L), in accordance with results previously obtained with human cortical cells transplanted in neonatal cortex (Espuny-Camacho et al., 2013).

We then examined qualitatively the patterns of axonal projections of the transplants. At 2 MPT, consistent with immaturity, the transplants displayed little neurite outgrowth within the host, with GFP+ axons that were still positive for immature markers such as doublecortin (DCX) (Figures S2A–S2D). Axonal processes were detected mostly close to the site of the graft in the ipsilateral cortex, with few GFP+ axons extending distantly along the corpus callosum, along the internal capsule, and in the striatum (Figures S2E–S2H). In contrast, at 6 MPT, in accordance with the expression of generic and specific markers of mature pyramidal neurons from all cortical layers, the transplants displayed robust axonal outgrowth, with a high number of GFP+ projections located at short- and long-range distances from the graft location, including ipsilateral and contralateral cortical projections (target of layer II-III cortical neurons), and subcortical projections to striatum (target of layer V cortical neurons), thalamus (target of layer VI cortical neurons), and midbrain/hindbrain (target of layer V cortical neurons) (Figure 2).

Visual Cortical Transplants of Human PSC-Derived Cortical Neurons Display Visual-like Long-Range Axonal Projections in the Adult Brain

Next, we examined in more detail the pattern of projections from the transplants at 6 MPT, especially in terms of areal specificity. We found remarkable areal specificity of the axonal patterns, corresponding to visual and limbic cortex targets, within both cortical and subcortical structures (Figures 3A–3E). In the basal ganglia, transplant-derived projections preferentially targeted the dorsomedial striatum (DM) (a preferential target of visual cortex), rather than the dorsolateral striatum (DL) (preferential target of motor areas) (Figures 2A, 2D, 3A, 3B, and 3F). In the thalamus, transplant-derived axons were detected abundantly within visual/limbic thalamic nuclei (e.g., lateral geniculate nucleus [LG], lateral posterior thalamic nucleus [LP], laterodorsal thalamic nucleus [LD]), while only a minority of fibers were detected in other primary non-visual thalamic nuclei (e.g., ventrobasal complex [VB], posterior thalamic nucleus group [Po], ventrolateral thalamic nucleus [VL], medial geniculate nucleus [MG]) (Figures 2A, 2F, and 3C–3F). Within the midbrain/hindbrain, transplant-derived axonal fibers were detected mostly in the superior colliculus (SC) and periacqueductal gray matter (PAG), which receive mainly inputs from the visual cortex and the limbic cortex, while fewer or no projections were detected in non-visual targets such as the inferior colliculus (IC), pedunculopontine nucleus (PPTg), red nucleus (RedN), and pyramidal tract (Figures 2A, 2G, and 3F). In the contralateral cortex, we also observed projections at the level of the V1/V2 visual cortex and less so at the level of the motor cortex (Figure 2A). A comparison with data from the Allen Mouse Brain Connectivity Atlas following anterograde injection into the V1 visual cortex shows that projections of human transplanted neurons match the normal projections of the visual cortex (Oh et al., 2014).

Figure 3.

Figure 3

Human ESC-Derived Cortical Neurons Grafted onto the Visual Cortex Display Specific Visual Cortex-like Projections 6 MPT

(A and B) Immunofluorescence images showing numerous GFP+ axons in the visual cortical target DM striatum (A) and fewer in the DL striatum (B), which is the target of the motor cortex.

(C–E) GFP+ axons also are detected in the targets of the visual cortex LG (C), LP, and LD (D) (thalamus) and fewer in the somatosensory-motor cortical target VB (thalamus) (E). DM, dorsomedial striatum; DL dorsolateral striatum; LV, lateral ventricle; CC, corpus callosum; dLG, dorsolateral geniculate nucleus; vLG, ventrolateral geniculate nucleus; LD, laterodorsal thalamic nucleus; LP, lateral posterior thalamic nucleus; VB, ventrobasal complex; Rt, reticulate thalamus.

(F) Quantification of the proportion of GFP+ fibers detected in the striatum, thalamus, and midbrain/hindbrain of grafts located into the visual cortex at 6 MPT. Data are represented as mean percentages ± SEMs (n = 10 animals). Wilcoxon’s non-parametric test for paired data; ∗∗p < 0.01. LG, lateral geniculate nucleus; Po, posterior thalamic nuclear group; VL, ventrolateral thalamic nucleus; MG, medial geniculate nuclei of the thalamus; SC, superior colliculus; IC, inferior colliculus; PAG, periaqueductal grey matter; PPtg, pedunculopontine nucleus; RedN, red nucleus; Pyr, pyramidal tract.

Scale bars: 100 μm.

These data indicate that human PSC-derived cortical neurons transplanted into the lesioned adult visual cortex display a mostly visual-like pattern of axonal projections, which is in line with the patterns observed following neonatal xenotransplantation (Espuny-Camacho et al., 2013) and with what was observed with transplanted mouse ESC-derived visual-like cortical neurons (Michelsen et al., 2015, Gaspard et al., 2008).

Long-Range Projections and Transplant Maturation Require a Visual Areal Identity Match

We found that the efficient integration of transplanted mouse ESC-derived cortical neurons or murine embryonic cortical tissue was achieved only if there was a precise match in the areal identity between lesioned and transplanted cells (Gaillard et al., 2007, Michelsen et al., 2015). Therefore, we compared our results obtained with transplants into the visual cortex with the outcome of the transplantations into a non-matching area, the motor cortex at 6 MPT (Figures S3B and S4A). We detected a transplant in 23 of 37 (62%) animals grafted in the motor cortex, including in 82% at 6 MPT, a rate that is slightly higher than that obtained with visual cortex transplantation, while graft volume was not significantly different between visual and motor locations (Figure 4A). However, examination of the pattern of axonal projections revealed that while a similar proportion of transplants projected fibers to the striatum (91% in visual transplants versus 71% in motor-located transplants), fewer motor-located transplants (<50%) sent long-range projections to the thalamic or midbrain/hindbrain levels compared to transplants in the visual cortex (91% in visual transplants versus 47% in motor-located transplants) (Figure 4B). Moreover, quantification of the total number of axonal fibers revealed that motor-located transplants sent far fewer axonal projections to specific subcortical targets, such as LG, LP, LD, SC, and PAG, with the exception of the striatum when compared to visual grafts (Figures 4D and S4B–S4E). It is interesting, however, that the distribution of fibers found in the striatum, thalamus, and midbrain/hindbrain from motor-located transplants also corresponded to a visual areal identity, albeit at a much lower efficiency (Figure 4E).

Figure 4.

Figure 4

Successful Graft Integration of hPSC-Derived Neurons into the Lesioned Adult Murine Cortex Requires an Areal Identity Match

(A) Quantification of the graft volume of human transplants located in the visual cortex or the motor cortex. Data are represented as mean percentages ± SEMs, compared to the value in visual grafts as 100 (n = 8 animals with visual cortex grafts; n = 8 animals with motor cortex grafts). Student’s t test; ns = non-significant.

(B) Quantification of the proportion of human transplants presenting GFP+ fibers in striatum (yellow bars) and thalamus and midbrain (green bars) from the total number of grafts located in the visual cortex or in the motor cortex. Data are represented as mean percentages ± SEPs (n = 11 animals with visual cortex grafts and n = 17 animals with motor cortex grafts). Z test; ns = non-significant; p < 0.05.

(C) Schematic illustration of the pattern of axonal projections obtained following grafting of human visual cells into the adult visual cortex (V1) (left) or the motor cortex (M1) (right).

(D) Quantification of the total number of fibers in the striatum, thalamus, and midbrain/hindbrain of grafts located in the visual cortex or the motor cortex 6 MPT. Data are represented as SEMs (n = 11 animals with visual cortex grafts; n = 17 animals with motor cortex grafts). Mann-Whitney non-parametric test for unpaired data; p < 0.05; ∗∗p < 0.01; ∗∗∗p < 0.001.

(E) Quantification of the proportion of GFP+ fibers in the striatum, thalamus, and midbrain/hindbrain of grafts located in the motor cortex at 6 MPT. Data are represented as mean percentages ± SEMs (n = 8 animals with motor cortex grafts). Wilcoxon’s non-parametric test for paired data; p < 0.05. DM, dorsomedial striatum; DL, dorsolateral striatum; LG lateral geniculate nucleus; LP, lateral posterior thalamic nucleus; LD, laterodorsal thalamic nucleus; VB, ventrobasal complex; Po, posterior thalamic nuclear group; VL, ventrolateral thalamic nucleus; MG, medial geniculate nuclei of the thalamus; SC, superior colliculus; IC, inferior colliculus; PAG, periaqueductal grey matter; PPTg, pedunculopontine nucleus; RedN, red nucleus; Pyr, pyramidal tract.

We then examined in more detail the cellular composition of the transplants in visual versus motor transplants. As expected, both visual and motor-located grafts showed broad expression of the pan-neuronal marker beta 3 class III tubulin (Figures 5A–5F). Notably, the percentage of Tbr1+ or Foxp2+ deep layer cortical neurons was not significantly different between visual versus motor-located grafts (Figure 5G). Alternatively, the expression of mature neuronal and mature glutamatergic markers NeuN and VGlut1 displayed lower levels of expression in motor transplants (Figures 5A–5D). Similarly, Couptf1, an early marker of the occipital cortex (Armentano et al., 2007) that is widely expressed in ESC-derived cortical cells in vitro and following their transplantation in the lesioned visual cortex (Espuny-Camacho et al., 2013, Gaspard et al., 2008), also was expressed at lower levels in the motor-located transplants (Figures 5E, 5F, and 5H). Collectively these data indicate that heterotypic transplants in the motor cortex, despite the fact that they contain similar numbers of cortical neurons, fail to re-establish long-range axonal projections and to mature properly at the molecular level.

Figure 5.

Figure 5

Visual Cortex Transplants of Human ESC-Derived Neurons Show Mature Neuronal, Glutamatergic, and Visual Cortical Markers at 6 MPT

(A–F) Immunofluorescence images showing the GFP+ graft (green), beta 3 class III tubulin (blue), and the mature neuronal marker NeuN (red; A and B); the glutamatergic marker VGlut1 (red; C and D); and the visual cortical marker Couptf1 (red; E and F) following transplantation into the motor cortex (A, C, and E) or the visual cortex (B, D, and F) at 6 MPT.

(G) Quantification of the percentage of Tbr1+ and Foxp2+ cells among total GFP+ cells found in grafts located in the motor or the visual cortex. Data are represented as mean percentages ± SEMs (n = 4 animals with visual cortex graft; n = 4 animals with motor cortex graft). Student’s t test; ns = non-significant.

(H) Quantification of the percentage of Couptf1+ cells among total GFP+ cells found in grafts located in the motor cortex or the visual cortex. Data are represented as means ± SEMs (n = 3 animals with visual cortex graft; n = 3 animals with motor cortex graft). Student’s t test; p < 0.05.

Scale bars: 50 μm.

Human Transplants in the Visual Cortex Are Synaptically Integrated into the Adult Mouse Brain

To determine whether the transplanted neurons actually integrate at the synapse level, we performed specific marker analyses and patch-clamp recordings. Marker analyses revealed robust expressions of maturation markers for dendrites, axons, and synapses (Figures 6A–6D).

Figure 6.

Figure 6

Functional Maturation of Transplanted Human ESC-Derived Cortical Neurons

(A–D) Immunofluorescence images showing the GFP+ graft (green) and the dendritic marker MAP2 (red; A); the axonal marker Tau (red; B); the presynaptic marker synaptophysin Syp (red; C), and the postsynaptic marker Homer1 (red; D). Scale bars: 50 μm (A and B) and 10 μm (C and D).

(E) Sample membrane voltage traces of a cell evoked by the intracellular injection of hyperpolarizing and depolarizing current steps (top) and of a 10-s-long ramp current (bottom).

(F and G) Single-cell passive (F) and active (G) membrane properties. Open circles correspond to individual cells, with the bar representing the population mean. Both passive and active properties are in the ranges that are observed typically in cortical pyramidal neurons.

(H) Membrane voltage traces recorded intracellularly from three different cells (blue, red, and black) in current-clamp mode—without injection of any offset current—display EPSPs (n = number of cells for which the specific properties were determined.

(I) Spontaneous EPSPs properties, extracted by fitting each spontaneous event with a double-exponential function. Open circles are averages of all of the events recorded in individual cells (n = 3 cells with significant spontaneous synaptic activity), with the bar representing the population mean.

(J) Sample membrane voltage responses of a neuron (same cell as in H, black traces) evoked by a train of 10 extracellular electrical stimuli repeated at 20 Hz, with a last “recovery” pulse at 500 ms delivered after the last pulse in the train. All of the traces represent the average responses for 20–30 repetitions. The top trace shows the evoked response during perfusion with standard aCSF. The application of selective competitive antagonists of α-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate (AMPA) and N-methyl-d-aspartate (NMDA) receptors (CNQX and AP-V, respectively) blocks the glutamatergic responses, significantly suppressing the evoked PSPs (central trace). The residual response was GABAA receptor mediated, as demonstrated by its abolishment following application of the GABAA receptor antagonist gabazine (bottom trace).

To test functional maturation and synaptic integration, we then performed patch-clamp recordings of transplanted neurons from brain tissue slices prepared from late-stage transplanted animals (8–9 MPT) into the visual cortex. This revealed that the majority of the cells recorded (8 of 13) presented a regular accommodating firing pattern that is characteristic of pyramidal neurons (Figures 6E–6G, S5A, and S5C). Stuttering firing profile, characteristic of some types of cortical interneurons, also was detected in a minority of the cells (3 of 13) (Figure S5B), reflecting that a minority of putative GABAergic cells also was present in the transplants, as shown by Espuny-Camacho et al. (2013. To test whether the grafted human neurons received functional synaptic connections, we measured an electrophysiological correlate of incoming synaptic connectivity. We observed that the human neurons displayed at rest excitatory postsynaptic potentials (EPSPs), present in 3 of 13 cells (Figures 6H and 6I), confirming that the human grafted cells received functional synapses.

Finally, we tested whether the human cells displayed signatures of functional incoming synaptic activity from mouse neuronal networks. We measured their evoked response to a train of extracellular electrical stimuli applied onto mouse host cortical cells located at a distance from the graft. The human neurons (three cells of three tested) showed robust evoked synaptic responses that could be blocked by specific glutamatergic (cyanquixaline [CNQX] and 2-amino-5-phosphonopentanoic acid [AP-V]) and GABAergic (gabazine) receptor antagonists (Figure 6J).

These data indicate that human ESC-derived cortical neurons transplanted into the adult mouse cortex present a mature electrical neuronal profile and are effectively integrated into the glutamatergic and GABAergic adult mouse host circuitry at 8 MPT.

Discussion

Cell replacement has been proposed and tested as a promising therapeutic avenue for diseases characterized by specific neuronal loss, in particular, Parkinson disease (Aboody et al., 2011, Barker et al., 2015, Tabar and Studer, 2014), but whether this approach could be applied to the loss of a wide diversity of cells in complex structures such as the cerebral cortex is much less clear to date (Goldman, 2016). Here, we differentiated human visual-like cortical projection pyramidal neurons and transplanted them into the lesioned adult murine cortex to test whether these cells could contribute to the reassembly of damaged cortical areas. We found that the human cortical neurons integrated stably into the host-damaged cortex, expressed molecular markers specific for different cortical layer neurons, displayed robust and specific long-range subcortical axonal projections corresponding to the damaged area, and established functional synapses with the host brain.

Efficient brain repair should not only involve axonal outgrowth and the generation of synapses but also, most important, should result in appropriate specificity in connectivity that would resemble closely the damaged circuits that need to be reestablished. It is important to note in this context that we report that the pattern of axonal projections from human neurons transplanted into the lesioned visual cortex corresponds to a visual cortical identity, thus potentially contributing to the reassembly of damaged visual pathways. These findings are in accordance with what we previously reported with mouse ESC-derived cortical cells transplanted into the lesioned adult visual cortex (Michelsen et al., 2015) and with transplants prepared from mouse embryonic cortex (Gaillard et al., 2007).

These data suggest that the human cortical progenitors and neurons, similar to those in the mouse, acquired a visual cortical identity in vitro that was maintained following adult transplantation, thereby allowing area-specific re-establishment of visual axonal pathways.

Moreover, we observed that successful transplantation was obtained only following transplantation in the visual cortex, suggesting that areal identity match may be an important factor for successful cortical transplantation, as previously observed with transplanted mouse cortical cells (Gaillard et al., 2007, Michelsen et al., 2015). Human ESC-derived neurons were previously transplanted in damaged rodent cortex and displayed evidence of synaptic input from the host brain and improved functional recovery (Tornero et al., 2013, Tornero et al., 2017), but they showed little apparent ability to send long-range axonal projections. The latter is reminiscent of our data obtained in the motor cortex transplantation and may be the result of lack of identity match between transplanted cells and the lesioned area.

Conversely, our results contrast with several studies using, for instance, human neural stem cell transplantations that led to successful patterns of connectivity with the same cells independent of the sites of transplantation (Doerr et al., 2017, Steinbeck et al., 2012, Wernig et al., 2004). The differences could be caused by distinct degrees of fate specification before transplantation. Completely fated cortical cells, such as the ones used in this study, do not display much capacity to be influenced by or adapt to the host adult brain, while other “earlier” types of neural stem cells that are less committed may be more capable of doing so, as shown with mouse embryonic cortex (Gaillard et al., 2003). These and our current data emphasize the importance of determining the underlying mechanisms regulating fate acquisition, maintenance, and plasticity following transplantation because these could have key implications for successful transplantation in clinical settings. In addition, it will be important to test further the relevance and importance of areal identity match by testing the efficiency of human motor cortex-fated cells transplanted in motor versus non-motor cortex locations, as was done with mouse embryonic cortical cells (Gaillard et al., 2007).

Conversely, it will be interesting to determine whether and how molecular cues could be involved in the apparent area specificity observed following adult transplantation. While it seems unlikely that the re-establishment of damaged axonal pathways strictly recapitulates developmental events, it is tempting to speculate that specific guidance cues normally acting during embryogenesis (Vanderhaeghen and Polleux, 2004) may be residually expressed or re-expressed following lesions (e.g., on remaining axons), thereby allowing the reassembly of lost neural circuits with the observed specificity. In addition, it will be interesting to test whether specific patterns of neural activity could contribute to the area-specific re-establishment of damaged pathways, as they do in normal development (Sur and Rubenstein, 2005).

In addition to long-range projections, the transplanted human cortical neurons displayed electrophysiological properties that indicate their functional maturation and integration with the host brain. However, much work remains to be done to determine whether these neurons can function in a physiologically relevant way, in this case to subserve visual function, as recently shown for mouse embryonic cortex-derived transplanted neurons (Falkner et al., 2016).

Moreover, while the use of visual neurons appears to be a fruitful model to study function in an experimental setting, from a therapeutic point of view, it will be important to be able to generate human cortical neurons of other areal identities, such as motor and somatosensory, and test their relevance for the restoration of motor function (e.g., as would be mostly required following stroke or brain trauma). The ability to obtain motor cortex-fated cells is still missing in the field and will require a thorough exploration of the spatial patterning cues that control cortical neuronal fate in vivo and in vitro (Sur and Rubenstein, 2005). Another exciting approach would be to try to combine the transplantation of cortical pyramidal neurons with cortical interneurons, because the latter also have been shown to be able to integrate following transplantation (Anderson and Vanderhaeghen, 2014, Cunningham et al., 2014, Maroof et al., 2013, Nicholas et al., 2013). Finally, it will be interesting to study the temporal aspect of human corticogenesis following transplantation in the adult brain. Indeed, as observed for neonatal transplantation (Qi et al., 2017, Espuny-Camacho et al., 2013, Espuny-Camacho et al., 2017), human neurons mature and integrate into adult neural circuits only following MPT. Human neurons transplanted in the neonatal or adult brain appear to mature functionally following their own time-specific pace, reminiscent of the neoteny that characterizes human cortical synaptogenesis (Defelipe, 2011, Suzuki and Vanderhaeghen, 2015). Such a slow speed may constitute a hurdle for an efficient use in clinical settings, although it may be necessary to achieve the reassembly of neural circuits in a physiologically relevant way.

In conclusion, we show here that human cortical projection neurons can be efficiently transplanted into the lesioned rodent cortex, with a specificity that may be highly relevant for the repair of defined lesions of cortical areas. These data constitute an important proof of principle to explore the possibilities and mechanisms underlying the reconstruction of lost neuronal circuits in the adult mammalian brain.

Experimental Procedures

Human ESC Differentiation into Cortical Cells

For the differentiation of human ESCs toward cortical progenitors and neurons (Espuny-Camacho et al., 2013), cells were dissociated using StemPro Accutase on day 2 (Invitrogen, catalog no. A11105) and plated on Matrigel (BD, human ESC-qualified Matrigel)-coated coverslips/dishes at low confluency (5,000–10,000 cells/cm2) in MEF-conditioned human ESC/human iPSC medium supplemented with rho-associated protein kinase (ROCK) inhibitor (Y-27632; 10 μM Calbiochem, Sigma-Aldrich, catalog no. 688000). On day 0 of the differentiation, the medium was changed to default differentiation medium (DDM) (Gaspard et al., 2008) supplemented with B27 (10 mL B27 per 500 mL DDM, GIBCO) and Noggin (100 ng/mL, R&D Systems) (Espuny-Camacho et al., 2013). At 16 days in vitro (DIV), Noggin was withdrawn and the medium was changed to DDM and supplemented with B27 (GIBCO). At 24 DIV the human cortical progenitors were manually dissociated and a dilution of 100,000 cells/μL was prepared for transplantation experiments, as described by Espuny-Camacho et al. (2013).

Mice Lesioning and Grafting

All of the animal experiments were carried out in full compliance with the Université Libre de Bruxelles Committee for animal welfare and Belgian and European regulations. Data for this study are derived from a total of 58 adult mice of both sexes (n = 37 for motor-located transplants and n = 21 for visual-located transplants). Lesioning and grafting experiments were performed in 6- to 8-week old adult NOD/SCID mice, as described by Michelsen et al. (2015). Lesioning was performed under anesthesia by focal stereotactic injections of ibotenic acid either in the visual or motor cortex (20 mg of ibotenic acid in 1 μL distilled water; see Supplemental Experimental Procedures for coordinates). Three days after lesioning, approximately 100,000 human ESC-derived cortical progenitors (in 0.5–1 μL) were injected in the visual or the motor cortex.

Quantification of Axonal Projections

GFP+ fibers were quantified using a Zeiss Axioplan upright fluorescence microscope equipped with 20x and 40x dry objectives. The GFP was enhanced by immunofluorescence staining against GFP. For the striatum, the total number of fibers was counted in one-sixth of the coronal sections collected for each individual brain. For the thalamus and midbrain, fibers were counted in half of the coronal sections collected for each individual brain. Data are represented as means ± SEMs. For the percentage of fibers, 10 and 8 animals were used for quantification following transplantation into the visual and motor cortex, respectively. For the total number of fibers in visual and motor transplants, 11 and 17 animals were used for quantification following transplantation into the visual cortex and the motor cortex, respectively.

Quantification of the Tbr1, Foxp2, and Couptf1 Cells in Transplants

The percentage of Tbr1+, Foxp2+, and Couptf1+ cells detected among the total number of GFP+ cells (human grafted cells) was counted on confocal acquired images from immunofluorescence stained brain sections. Data are represented as mean percentages ± SEMs (n = 4−3 animals grafted into the visual cortex; n = 4−3 animals grafted into the motor cortex).

Quantification of Graft Volume

The graft volume was measured from confocal images from the graft in serial coronal sections corresponding to one-sixth of the mouse brain. Mouse brain sections were immunostained with GFP antibody and z stacks of 10 μm were acquired for a total of 6 steps (60 μm thickness). The volume of the graft was calculated using Fiji software. Data are represented as mean percentages ± SEMs (n = 8 animals grafted into the visual cortex; n = 8 animals grafted into the motor cortex).

Quantification of the Proportion of Transplants with Subcortical Projections

The proportion of visual and motor-located transplants projecting GFP+ fibers into the striatum, thalamus, or midbrain/hindbrain was quantified by manual counting from immunofluorescence stained brain sections using the GFP antibody. Data are represented as mean percentages ± SE of the proportions (SEPs) (visual grafts: n = 11 animals; motor grafts: n = 17 animals).

Brain Tissue Slice Electrophysiology

Electrophysiological recordings were performed intracellularly on grafted neurons in acute brain slices from mice at 8 and 9 MPT as previously described (Espuny-Camacho et al., 2013). Briefly, 300-μm thick coronal slices were cut in a sucrose-based solution and subsequently stored in artificial cerebrospinal fluid (aCSF) before recording. Transplanted cells were identified by their EGFP fluorescence and visualized using an upright microscope equipped with infrared differential interference contrast (DM LFS, Leica Microsystems). Whole-cell patch-clamp recordings and bipolar extracellular stimulation were performed and analyzed as described in more detail in the Supplemental Experimental Procedures.

Statistics

Statistical analyses were performed using PRISM (GraphPad). Appropriate statistical tests were chosen in each case depending on sample size, data distribution, and number of comparisons. The percentage of axonal projections in different subcortical targets for visual or motor transplants was analyzed using the non-parametric Wilcoxon signed-rank test for paired data, with p < 0.05 < and ∗∗p < 0.01. The total number of fibers in visual and motor transplants was analyzed using the non-parametric Mann-Whitney test for unpaired data, with p < 0.05 <, ∗∗p < 0.01, and ∗∗∗p < 0.001. The proportion of Tbr1+, Foxp2+, and Couptf1+ cells was analyzed by Student’s t test; n.s. = non-significant; p < 0.05. The proportion of graft volume in motor and visual transplants was analyzed by Student’s t test; n.s. = non-significant. The proportion of transplants with subcortical projections was performed following the Z test for proportions; ns = non-significant; p < 0.05.

Additional Methods

An extended version of the experimental procedures is described in the Supplemental Experimental Procedures.

Acknowledgments

We thank members of the laboratory and IRIBHM for helpful feedback and support and Jean-Marie Vanderwinden of the Light Microscopy Facility (LiMiF) for support with imaging. We also thank Chiara Cordiglieri from INGM Imaging for support with 3D image analysis. This work was funded by grants from the Belgian Fund for Scientific Research (FNRS), the Belgian Queen Elizabeth Medical Foundation, the Interuniversity Attraction Poles Program (IUAP), the WELBIO Program of the Walloon Region, the European Research Council (ERC-2013-AG-340020), the AXA Research Fund, the Fondation ULB, the VIB, and KULeuven (to P.V.) and the Flanders Research Foundation (FWO) (grant no. G0F1517N to M.G.). I.E.-C. was a post-doctoral fellow of the FNRS.

Author Contributions

I.E.-C. performed all of the differentiation and transplantation experiments with the help of A.G., K.A.M., A.B., S.A.-V., and A.H. K.A.M. performed the axon fiber quantifications. D.L. performed the electrophysiological recordings. I.E.-C. and P.V. wrote the manuscript. P.V., A.G., M.G., and I.E.-C. designed and analyzed all of the experiments.

Declaration of Interests

The authors declare no competing interests.

Published: May 29, 2018

Footnotes

Supplemental Information includes Supplemental Experimental Procedures and five figures and can be found with this article online at https://doi.org/10.1016/j.celrep.2018.04.094.

Supplemental Information

Document S1. Supplemental Experimental Procedures and Figures S1–S5
mmc1.pdf (12.5MB, pdf)
Document S2. Article plus Supplemental Information
mmc2.pdf (19.4MB, pdf)

References

  1. Aboody K., Capela A., Niazi N., Stern J.H., Temple S. Translating stem cell studies to the clinic for CNS repair: current state of the art and the need for a Rosetta stone. Neuron. 2011;70:597–613. doi: 10.1016/j.neuron.2011.05.007. [DOI] [PubMed] [Google Scholar]
  2. Anderson S., Vanderhaeghen P. Cortical neurogenesis from pluripotent stem cells: complexity emerging from simplicity. Curr. Opin. Neurobiol. 2014;27:151–157. doi: 10.1016/j.conb.2014.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Arlotta P., Berninger B. Brains in metamorphosis: reprogramming cell identity within the central nervous system. Curr. Opin. Neurobiol. 2014;27:208–214. doi: 10.1016/j.conb.2014.04.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Armentano M., Chou S.-J., Tomassy G.S., Leingärtner A., O’Leary D.D.M., Studer M. COUP-TFI regulates the balance of cortical patterning between frontal/motor and sensory areas. Nat. Neurosci. 2007;10:1277–1286. doi: 10.1038/nn1958. [DOI] [PubMed] [Google Scholar]
  5. Barker R.A., Drouin-Ouellet J., Parmar M. Cell-based therapies for Parkinson disease—past insights and future potential. Nat. Rev. Neurol. 2015;11:492–503. doi: 10.1038/nrneurol.2015.123. [DOI] [PubMed] [Google Scholar]
  6. Cunningham M., Cho J.H., Leung A., Savvidis G., Ahn S., Moon M., Lee P.K.J., Han J.J., Azimi N., Kim K.S. hPSC-derived maturing GABAergic interneurons ameliorate seizures and abnormal behavior in epileptic mice. Cell Stem Cell. 2014;15:559–573. doi: 10.1016/j.stem.2014.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Defelipe J. The evolution of the brain, the human nature of cortical circuits, and intellectual creativity. Front. Neuroanat. 2011;5:29. doi: 10.3389/fnana.2011.00029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Doerr J., Schwarz M.K., Wiedermann D., Leinhaas A., Jakobs A., Schloen F., Schwarz I., Diedenhofen M., Braun N.C., Koch P. Whole-brain 3D mapping of human neural transplant innervation. Nat. Commun. 2017;8:14162. doi: 10.1038/ncomms14162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Eiraku M., Watanabe K., Matsuo-Takasaki M., Kawada M., Yonemura S., Matsumura M., Wataya T., Nishiyama A., Muguruma K., Sasai Y. Self-organized formation of polarized cortical tissues from ESCs and its active manipulation by extrinsic signals. Cell Stem Cell. 2008;3:519–532. doi: 10.1016/j.stem.2008.09.002. [DOI] [PubMed] [Google Scholar]
  10. Espuny-Camacho I., Michelsen K.A., Gall D., Linaro D., Hasche A., Bonnefont J., Bali C., Orduz D., Bilheu A., Herpoel A. Pyramidal neurons derived from human pluripotent stem cells integrate efficiently into mouse brain circuits in vivo. Neuron. 2013;77:440–456. doi: 10.1016/j.neuron.2012.12.011. [DOI] [PubMed] [Google Scholar]
  11. Espuny-Camacho I., Arranz A.M., Fiers M., Snellinx A., Ando K., Munck S., Bonnefont J., Lambot L., Corthout N., Omodho L. Hallmarks of Alzheimer’s disease in stem-cell-derived human neurons transplanted into mouse brain. Neuron. 2017;93:1066–1081.e8. doi: 10.1016/j.neuron.2017.02.001. [DOI] [PubMed] [Google Scholar]
  12. Falkner S., Grade S., Dimou L., Conzelmann K.-K., Bonhoeffer T., Götz M., Hübener M. Transplanted embryonic neurons integrate into adult neocortical circuits. Nature. 2016;539:248–253. doi: 10.1038/nature20113. [DOI] [PubMed] [Google Scholar]
  13. Gage F.H., Temple S. Neural stem cells: generating and regenerating the brain. Neuron. 2013;80:588–601. doi: 10.1016/j.neuron.2013.10.037. [DOI] [PubMed] [Google Scholar]
  14. Gaillard A., Nasarre C., Roger M. Early (E12) cortical progenitors can change their fate upon heterotopic transplantation. Eur. J. Neurosci. 2003;17:1375–1383. doi: 10.1046/j.1460-9568.2003.02576.x. [DOI] [PubMed] [Google Scholar]
  15. Gaillard A., Prestoz L., Dumartin B., Cantereau A., Morel F., Roger M., Jaber M. Reestablishment of damaged adult motor pathways by grafted embryonic cortical neurons. Nat. Neurosci. 2007;10:1294–1299. doi: 10.1038/nn1970. [DOI] [PubMed] [Google Scholar]
  16. Gascón S., Masserdotti G., Russo G.L., Götz M. Direct neuronal reprogramming: Achievements, hurdles and new roads to success. Cell Stem Cell. 2017;21:18–34. doi: 10.1016/j.stem.2017.06.011. [DOI] [PubMed] [Google Scholar]
  17. Gaspard N., Bouschet T., Hourez R., Dimidschstein J., Naeije G., van den Ameele J., Espuny-Camacho I., Herpoel A., Passante L., Schiffmann S.N. An intrinsic mechanism of corticogenesis from embryonic stem cells. Nature. 2008;455:351–357. doi: 10.1038/nature07287. [DOI] [PubMed] [Google Scholar]
  18. Goldman S.A. Stem and progenitor cell-based therapy of the central nervous system: hopes, hype, and wishful thinking. Cell Stem Cell. 2016;18:174–188. doi: 10.1016/j.stem.2016.01.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Maroof A.M., Keros S., Tyson J.A., Ying S.-W., Ganat Y.M., Merkle F.T., Liu B., Goulburn A., Stanley E.G., Elefanty A.G. Directed differentiation and functional maturation of cortical interneurons from human embryonic stem cells. Cell Stem Cell. 2013;12:559–572. doi: 10.1016/j.stem.2013.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Michelsen K.A., Acosta-Verdugo S., Benoit-Marand M., Espuny-Camacho I., Gaspard N., Saha B., Gaillard A., Vanderhaeghen P. Area-specific reestablishment of damaged circuits in the adult cerebral cortex by cortical neurons derived from mouse embryonic stem cells. Neuron. 2015;85:982–997. doi: 10.1016/j.neuron.2015.02.001. [DOI] [PubMed] [Google Scholar]
  21. Nicholas C.R., Chen J., Tang Y., Southwell D.G., Chalmers N., Vogt D., Arnold C.M., Chen Y.J.J., Stanley E.G., Elefanty A.G. Functional maturation of hPSC-derived forebrain interneurons requires an extended timeline and mimics human neural development. Cell Stem Cell. 2013;12:573–586. doi: 10.1016/j.stem.2013.04.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Oh S.W., Harris J.A., Ng L., Winslow B., Cain N., Mihalas S., Wang Q., Lau C., Kuan L., Henry A.M. A mesoscale connectome of the mouse brain. Nature. 2014;508:207–214. doi: 10.1038/nature13186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Péron S., Droguerre M., Debarbieux F., Ballout N., Benoit-Marand M., Francheteau M., Brot S., Rougon G., Jaber M., Gaillard A. A delay between motor cortex lesions and neuronal transplantation enhances graft integration and improves repair and recovery. J. Neurosci. 2017;37:1820–1834. doi: 10.1523/JNEUROSCI.2936-16.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Qi Y., Zhang X.J., Renier N., Wu Z., Atkin T., Sun Z., Ozair M.Z., Tchieu J., Zimmer B., Fattahi F. Combined small-molecule inhibition accelerates the derivation of functional cortical neurons from human pluripotent stem cells. Nat. Biotechnol. 2017;35:154–163. doi: 10.1038/nbt.3777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Shi Y., Kirwan P., Smith J., Robinson H.P.C., Livesey F.J. Human cerebral cortex development from pluripotent stem cells to functional excitatory synapses. Nat. Neurosci. 2012;15:477–486. doi: 10.1038/nn.3041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Steinbeck J.A., Koch P., Derouiche A., Brüstle O. Human embryonic stem cell-derived neurons establish region-specific, long-range projections in the adult brain. Cell. Mol. Life Sci. 2012;69:461–470. doi: 10.1007/s00018-011-0759-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Sur M., Rubenstein J.L.R. Patterning and plasticity of the cerebral cortex. Science. 2005;310:805–810. doi: 10.1126/science.1112070. [DOI] [PubMed] [Google Scholar]
  28. Suzuki I.K., Vanderhaeghen P. Is this a brain which I see before me? Modeling human neural development with pluripotent stem cells. Development. 2015;142:3138–3150. doi: 10.1242/dev.120568. [DOI] [PubMed] [Google Scholar]
  29. Tabar V., Studer L. Pluripotent stem cells in regenerative medicine: challenges and recent progress. Nat. Rev. Genet. 2014;15:82–92. doi: 10.1038/nrg3563. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Tornero D., Wattananit S., Grønning Madsen M., Koch P., Wood J., Tatarishvili J., Mine Y., Ge R., Monni E., Devaraju K. Human induced pluripotent stem cell-derived cortical neurons integrate in stroke-injured cortex and improve functional recovery. Brain. 2013;136:3561–3577. doi: 10.1093/brain/awt278. [DOI] [PubMed] [Google Scholar]
  31. Tornero D., Tsupykov O., Granmo M., Rodriguez C., Grønning-Hansen M., Thelin J., Smozhanik E., Laterza C., Wattananit S., Ge R. Synaptic inputs from stroke-injured brain to grafted human stem cell-derived neurons activated by sensory stimuli. Brain. 2017;140:692–706. doi: 10.1093/brain/aww347. [DOI] [PubMed] [Google Scholar]
  32. van den Ameele J., Tiberi L., Vanderhaeghen P., Espuny-Camacho I. Thinking out of the dish: what to learn about cortical development using pluripotent stem cells. Trends Neurosci. 2014;37:334–342. doi: 10.1016/j.tins.2014.03.005. [DOI] [PubMed] [Google Scholar]
  33. Vanderhaeghen P., Polleux F. Developmental mechanisms patterning thalamocortical projections: intrinsic, extrinsic and in between. Trends Neurosci. 2004;27:384–391. doi: 10.1016/j.tins.2004.05.009. [DOI] [PubMed] [Google Scholar]
  34. Wernig M., Benninger F., Schmandt T., Rade M., Tucker K.L., Büssow H., Beck H., Brüstle O. Functional integration of embryonic stem cell-derived neurons in vivo. J. Neurosci. 2004;24:5258–5268. doi: 10.1523/JNEUROSCI.0428-04.200. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Document S1. Supplemental Experimental Procedures and Figures S1–S5
mmc1.pdf (12.5MB, pdf)
Document S2. Article plus Supplemental Information
mmc2.pdf (19.4MB, pdf)

RESOURCES