J Neurol Surg A Cent Eur Neurosurg 2019; 80(06): 475-487
DOI: 10.1055/s-0039-1692976
Review Article
Georg Thieme Verlag KG Stuttgart · New York

Markers for Identifying and Targeting Glioblastoma Cells during Surgery

1   Department of Neurosurgery, University Hospital Münster, Münster, Germany
,
Michael Schwake
1   Department of Neurosurgery, University Hospital Münster, Münster, Germany
,
Eric Suero Molina
1   Department of Neurosurgery, University Hospital Münster, Münster, Germany
,
Walter Stummer
2   NCH, UK Münster, Münster, Germany
› Author Affiliations
Further Information

Address for correspondence

Stephanie Schipmann, MD
Department of Neurosurgery, University Hospital Münster
Albert-Schweitzer-Campus 1, Building A1, 48149 Münster
Germany   

Publication History

30 December 2018

04 March 2019

Publication Date:
29 August 2019 (online)

 

Abstract

Glioblastoma is a highly malignant tumor with a poor prognosis. A factor influencing survival that can be affected by the surgeon is the extent of resection (EOR). Due to the infiltrative nature of the tumor, delineation of tumor from normal brain parenchyma is often challenging. To improve EOR and facilitate tumor visualization, several techniques have been developed over the last few years. This literature review presents an overview of current intraoperative strategies for identifying and targeting glioma cells and discusses the benefits and limitations of each technique. Along with conventional techniques such as neuronavigation and ultrasound, fluorescence-guided surgery with different fluorescent agents such as 5-aminolevulinc acid and fluorescein have been widely used. Recently, newer techniques have emerged and are being translated into the operating room, promising delineation of glioblastoma tissue using targeted approaches or identification on a microscopic level, for instance using Raman spectroscopy or confocal microscopy.


#

Introduction

Glioblastoma is one of the most malignant tumors of the central nervous system. Despite a better understanding of tumor biology and the development of new treatment approaches during the last decades, the tumor remains incurable with a poor prognosis, characterized by a median survival of 15 months and a 2-year survival rate of 17.4%.[1] [2]

Current first-line treatment consists of surgical resection, followed by adjuvant radiation and chemotherapy.[3] Extent of resection (EOR) is an independent predictor of survival, first shown by Lacroix et al, who demonstrated that an EOR > 98% of tumor volume results in a significant survival advantage.[4] Another study indicated that the threshold might even be lower, with an EOR of 78% as a minimum associated with a survival benefit.[5] In addition, data from a large EORTC-NCIC trial, the pivotal trial for approval of temozolomide as first-line therapy for glioblastoma, revealed a greater benefit from adjuvant radiotherapy and chemotherapy in patients with gross total resection (GTR).[6]

Clearly, EOR is a critical driver of outcome in glioblastoma and the only factor that the surgeon can influence directly. Given the infiltrative nature and often eloquent tumor localization, achieving the largest possible EOR is often challenging. Tumor infiltration into the surrounding brain parenchyma is difficult to discriminate with the human eye under standard white light microscopy, and differentiation of tumor margins from normal brain often cannot be performed based on tactile features of tissues. Consequently, several techniques have been developed to improve the ability of the surgeon to identity glioblastoma tissue during surgery.

Several modalities aiming at improving EOR entered the field decades ago, such as neuronavigation, intraoperative ultrasound, and intraoperative magnetic resonance imaging (iMRI). Newer techniques, such as fluorescence-guided surgery (FGS) using various fluorescent agents, enables intraoperative real-time imaging of glioblastoma. Recently, more novel techniques have emerged and are being translated into the operating room, promising an even better delineation of glioblastoma tissue, such as Raman spectroscopy, confocal microscopy, or methods of targeted fluorescence.[7] [8] [9] [10] [11]

We performed a literature review on currently available techniques that aim at targeting and identifying glioblastoma cells during surgery, discussing their benefits, applicability, and limitations in the field of glioma surgery. This review focuses on FGS and newer techniques; conventional techniques are only mentioned for the sake of completeness but are not reviewed in further detail.


#

Conventional Techniques for Intraoperative Visualization of Glioblastoma

For most surgical glioblastoma cases, conventional techniques such as neuronavigation and ultrasound are standard of care and have been widely integrated into the operative setting.


#

Neuronavigation and Ultrasound

Neuronavigation is a basic and ubiquitously available tool for glioblastoma surgery. All relevant preoperative digital scans such as computed tomography (CT), MRI, and positron emission tomography (PET) can be incorporated into the data set for the navigational system and help the surgeon maintain a precise sense of complex three-dimensional anatomical relationships and almost real-time intraoperative localization.[12] Neuronavigation benefits from high surgical accuracy for resection of glioblastomas and can aid in planning the surgical approach.[13] [14] A study by Wirtz et al evaluated the effect of neuronavigation on the EOR in glioblastoma compared with standard use of the microscope. They showed that the amount of residual tumor was significant lower in the patients operated on using neuronavigation, without showing a clear difference regarding the number of radical resections.[15] A major limitation using neuronavigation is the loss of accuracy caused by intraoperative brain shift due to application of mannitol, drainage of cerebrospinal fluid, patient positioning, and resection of tissue.[16]

Another uncomplicated and cost-effective method for intraoperative glioblastoma localization is ultrasound. This dynamic method helps identify tumor borders and normal brain structures.[16] A retrospective analysis showed an increase of survival when surgeons used intraoperative ultrasound for identifying residual glioma.[17] However, concerns have been expressed regarding a sometimes poor differentiation of tumor from the zone of peritumoral edema, putting patients at risk for too extensive resections with neurologic sequelae.[18] In addition, ultrasound has its limitations in the delineation of normal brain tissue from high-grade glioma tumor borders after previous irradiation.[19]

Intraoperative acquired data from ultrasound can be used to update the navigation system and help overcome the limitation of brain shift.[20]


#

Intraoperative MRI

First introduced in the 1990s, iMRI has since undergone further development. It provides almost real-time images during surgery for identification of residual tumor and can also be used to detect possible intraoperative complications such as hematoma.[21]

In addition, the acquired images can be used to update the neuronavigational system to compensate impaired accuracy from brain shift.[22] [23] A prospective randomized controlled trial published by Senft et al compared the rates of GTR in glioma patients operated using conventional microsurgery and patients in whom iMRI was used. These authors showed a significant higher frequency of GTR in the iMRI group (96% iMRI group versus 68% control group; p = 0.023), providing evidence for the beneficial role of iMRI in glioma surgery.[24] Supporting results were presented by Hatiboglu et al, showing that when used by surgeons, iMRI led to an increase of the median EOR from 84% to 99% (p < 0.001) with additional tumor removal after iMRI in contrast-enhancing gliomas.[25]

However, iMRI has distinct disadvantages. This method is expensive, time consuming, and extends the duration of surgery and anesthesia. In addition, repeated application of gadolinium may result in extravasation into the tumor area and resection cavity, leading to false-positive effects.[21] [26]


#

Fluorescence-guided Surgery for Glioblastoma

The application of fluorescing compounds for differentiating tissue during brain surgery was first described in 1947 by George E. Moore, who noted a higher concentration of fluorescein in brain invaded by malignant glioma using a wood lamp.[27] After decades of dormancy, this approach was modernized in 1998 by the senior author of this review, Walter Stummer, by the introduction of 5-aminolevulinic acid (5-ALA) for FGS.[28] FGS is based on the administration of optical imaging agents to patients before surgery, leading to a (selective) accumulation in tumor cells, helpful in intraoperative real-time detection and delineation of tumor tissue. Currently, two agents are being used clinically in the field of glioblastoma surgery: 5-ALA[28] and fluorescein.[27] A third dye, indocyanine green (ICG), is under investigation.[29]


#

5-aminolevulinc Acid

5-ALA is worldwide the most intensely studied fluorescent agent for brain tumor surgery. It was approved by the European Medicines Agency (EMA) and the U.S. Food and Drug Administration (FDA) for intraoperative visualization of malignant glioma. 5-ALA is a natural metabolite in the hemoglobin pathway and within glioma cells.[30] [31] After oral administration of 20 mg/kg body weight (BW) 5-ALA (Gliolan in the European Union [Specialized Therapeutics, Melbourne, Australia]; Gleolan in the United States [NX Development Corp., Lexington, Kentucky, United States) 3 hours before the induction of anesthesia, fluorescence can be visualized by using a surgical microscope with a xenon light source that can switch between white and violet-blue light (wavelength: 370–440 nm) and is provided with an emission filter for visualization of red tumor fluorescence with peaks at between 635 and 704 nm, thus well in the red range.[30] [31] Peak fluorescence can be expected after 6 to 8 hours.[32] [33]

It was shown that 5-ALA has a high toxicologic safety with only minor side effects such as a temporary and mild elevation of liver enzymes and transient skin phototoxicity.[34] [35]

Visualization of Glioblastomas and Intensity of 5-ALA Fluorescence

The efficacy of 5-ALA for intraoperative visualization of glioblastoma cells was shown by several studies, and investigators uniformly report a high selectivity. In a series of 10 patients, with 89 tissue biopsies, sensitivity of 5-ALA–induced fluorescence for detection of malignant glioma cells was 85% and specificity was 100%.[28] In a meta-analysis, including eight studies on histopathologic analysis and intraoperative 5-ALA fluorescence with > 800 samples from malignant glioma, the specificity for glioblastoma was 88.8% and sensitivity 82.6%.[36]

Using 5-ALA, it is important to differentiate between different qualities of fluorescence because tumor fluorescence is not homogeneous. Two fluorescence qualities can be distinguished: a vivid solid red fluorescence, representing viable tumor, and a vague, less vivid pink fluorescence, indicating the tumor-infiltrating zone. These findings were supported by histologic and spectroscopic analyses ([Fig. 1]).[37] [38]

Zoom Image
Fig. 1 5-aminolevulinic fluorescence and different fluorescence qualities. (a) Cavity with area of strong (red) and weak (pink) and no fluorescence. (b) Corresponding white light image. (Reproduced with permission from Stummer et al.[38])
Zoom Image
Fig. 2 Use of fluorescein (FL) for resection of glioblastoma. (a) After administration of FL under white light: no fluorescent effect. (b) After administration of FL under YELLOW 560 nm filter: visible fluorescent effect in the tumor. (Reproduced with permission from Schebesch et al.[54])

Especially in case of solid fluorescence, a positive predictive value (PPV) of 100% was reported. PPV was lower, between 91% and 97%, in tissue with vague fluorescence in invasive areas at the tumor border.[38] [39]

Even in recurrent glioblastoma, where tissue scarring and changes induced by previous radiotherapy and chemotherapy are present, 5-ALA–guided resection was still shown to be effective with a PPV of 99.5%.[40]

Similar findings were reported by Lau et al, who analyzed 211 intraoperative high-grade glioma biopsies from different areas of fluorescence intensity graded from 0 to 3. They revealed a PPV of 100% for high-grade gliomas and 97.2% for glioblastomas in case of highest rated amount of fluorescence. However, the negative predictive value was comparably very low with 16.7% for high-grade tumors and 43.9% for glioblastoma, indicating that not all tumor-infiltrated areas may synthesize the dye in concentrations that can be visualized using the surgical microscope.[41] Consequently, 5-ALA is a very useful marker for tumor cellularity, especially in areas with solid and bright fluorescence.

5-ALA has the potential to be used as a tool for detection of solid tumor that can be removed without risk of neurologic deficit but will also help discrimination of infiltrated brain down to a tumor cell density of ∼ 10 to 20%, enabling even larger resection volumes in noneloquent regions than those identified by contrast enhancement on MRI.[39]


#

Influence of 5-ALA on Extent of Resection and Outcome

The first prospective study evaluating the impact of 5-ALA on the EOR was published in 2000 by Stummer et al, showing that complete resection of contrast-enhancing tumor on MRI was archived in 33 (63%) of 52 patients. In most of the remaining patients, complete resection could not be performed due to concerns about neurologic safety. In addition, the improved survival was related to the completeness of resection.[37] A large phase III randomized controlled study included 322 patients with suspected malignant glioma who were randomly assigned to 5-ALA–guided or conventional microsurgical resection. Complete resection of contrast-enhancing tumor was achieved in 90 of 139 (65%) patients in the 5-ALA group compared with 47 of 131 (36%) in the conventional group (p < 0.001). Furthermore, patients from the 5-ALA group had a longer 6-month progression-free survival (PFS) (41% versus 21.1% in the control group; p = 0.0003) with a median PFS of 5.1 months.[34]

Further studies confirmed the benefit of 5-ALA regarding EOR, and since then 5-ALA has been widely used in the resection of glioblastoma. The initially reported resection rate of 65% that was achieved using 5-ALA FGR was further improved over the last few years due to confidence in the use of the method, as well as advances in intraoperative monitoring and mapping, the latter allowing safe resections in eloquent areas. Díez et al reported GTR rates of 83.3% (30 of 36 glioblastoma patients),[39] whereas a retrospective study by Schucht et al reported a GTR of 96% (51 of 53 patients).[42] In comparison, GTR under white light microscopy is only achieved in 36% of patients,[34] indicating a major benefit of 5-ALA as an intraoperative adjunct for optimizing resection. Even in eloquent areas, the use of 5-ALA FGS, combined with intraoperative mapping or awake surgery, enables GTR rates of up to 76%.[43] [44]

A retrospective analysis of 52 glioblastoma patients with optimal resections according to conventional criteria (i.e., complete resection of contrast-enhancing tumors on early postoperative MRI) compared cases with residual fluorescent tissue and complete removal of fluorescent tissue, demonstrating an improved median overall survival (OS) of 27 months (95% confidence interval [CI], 22.4–31.6) in patients without residual fluorescence compared with 17.5 months (95% CI, 12.5–22.5) with residual fluorescence.[45] It is well known that intraoperative fluorescence exceeds the contrast enhancement visible on MRI, by far, marking almost double the resection volume outlined by contrast enhancement on MRI.[46] These data again underline the potential of 5-ALA for greater EOR and increased survival.


#
#

Combination of 5-ALA FGS with Intraoperative MRI

Coburger et al evaluated the benefit of the additional use of 5-ALA to iMRI in resection of glioblastoma in a prospective cohort and demonstrated that GTR was achieved significantly (p < 0.01) more often using the combined approach of 5-ALA and iMRI (100%) compared with iMRI alone (82%), with higher mean EOR in the combined group (99.7% versus 97.4%, respectively; p < 0.004), indicating a synergistic effect of both methods.[47]

Several studies compared the diagnostic accuracy of 5-ALA and iMRI for identifying brain infiltrated by glioma. Coburger et al described a significantly higher sensitivity (91% versus 66%) and specificity (90% versus 60%) for detection of malignant glioma than iMRI at the tumor border.[48] However, so far no clinical evidence has demonstrated the superiority of one method over the other. An ongoing trial (Impact of iMRI on the Extent of Resection in Patients with Newly Diagnosed Glioblastomas: A Prospective Multicenter Parallel Group Clinical Trial [NCT02379572]) aims at providing more data on comparison of both techniques regarding EOR. However, such comparisons may be purely academic because during surgery, technologies should be combined and synergisms utilized in the best interest of patients.

Analyzing the current literature regarding safety and side effects linked to 5-ALA, our data analysis indicates only minor toxicity such as mild and transient erythema or a mild elevation of liver enzymes in single cases without clinically relevant hepatic disorders. Overall, 5-ALA can be considered a very safe and well-tolerated drug. [Table 1] presents an overview of all included studies.

Table 1

Overview of studies analyzing the use of 5-ALA for glioma surgery

Study

Study design

No. of patients

Tumor type

Eloquence

Primary end point

5-ALA dosage

Drug-related side effects

Other intraoperative tools

GTR

Sensitivity

Specificity

PPV

NPV

Impact on survival

Stummer et al, 1998

Prospective, monocentric

10

8 GBMs, 2 AA

Eloquent and noneloquent

First evaluation of 5-ALA in malignant glioma, safety

10 mg/kg BW

None

Neuronavigation (one patient)

70%

85%

100%

90%

NR

NR

Stummer et al, 2000

Prospective, monocentric

52

GBM

Eloquent and noneloquent

GTR, survival, postoperative MRI findings

20 mg/kg BW

One patient with erythema; mild elevation of liver enzymes with no signs of hepatic disorders

None

63%

NR

NR

NR

NR

NR

Stummer et al, 2006

RCT, phase III

139

135 GBMs

4 grade IIIs

Eloquent and noneloquent

GTR, PFS 6 mo, postoperative MRI findings, adverse events

20 mg/kg BW

Liver enzymes were mildly elevated 24 h after surgery

Ultrasound or neuronavigation only for planning of approach

65%

NR

NR

NR

NR

6-mo PFS

5-ALA: 5.1 mo WL: 3.6 mo

No difference regarding OS

Nabavi et al, 2009

Prospective, multicentric, single-arm phase II

36

Recurrent HGGs

Eloquent and noneloquent

To assess feasibility of 5-ALA fluorescence guidance for resection of recurrent HGG, determine PPV

20 mg/kg BW

None

NR

19.4%

NR

NR

NR

NR

NR

Díez Valle et al, 2011

Prospective, monocentric

36

28 primary GBMs,

8 recurrent GBMs

Eloquent and noneloquent

GTR; safety, diagnostic accuracy

20 mg/kg BW

None

Neuronavigation, neuro-monitoring

83.3%

NR

NR

Solid fl: 100%;

vague fl: 97%

66%

NR

Della Puppa et al, 2012

Prospective, monocentric

31

22 primary HGGs, 9 recurrent HGGs

Only eloquent

5-ALA in eloquent areas assisted with functional mapping

20 mg/kg BW

None

IOM, neuronavigation, some cases awake surgery

76% newly diagnosed malignant gliomas; 66% recurrent gliomas

NR

NR

NR

NR

NR

Schucht et al, 2012

Retrospective, monocentric

36

GBM

Eloquent and noneloquent (only if complete resection could be achieved)

CRET and GTR; residual contrast-enhancing tissue

20 mg/kg BW

None

IOM, neuronavigation

96%

NR

NR

NR

NR

NR

Coburger et al, 2014

Prospective. monocentric

34

Primary and recurrent GBMs

Eloquent and noneloquent

Provide a histopathologic correlation of tumor delineation at the border zone of iMRI and 5-ALA

20 mg/kg BW

None

Neuronavigation, iMRI

NR

5-ALA: 85%

iMRI: 41%

5-ALA: 80%

iMRI: 60%

5-ALA: 69%

iMRI: 67%

5-ALA: 43%

iMRI: 70%

NR

Stummer et al, 2014

Prospective, monocentric

33

29 GBMs, 4 AAs

Eloquent and noneloquent

Determination of fluorescence quality

20 mg/kg BW

Transient elevation of liver enzymes

Neuronavigation

NR

NR

NR

Strong fl: 96.2%; weak fl: 92.%

39.5%

NR

Schucht et al, 2014

Prospective, monocentric

67

GBM

Eloquent, adjacent to corticospinal tract

Evaluation of mapping and 5-ALA–guided surgery in eloquent regions

20 mg/kg BW

NR

Mapping

57%

NR

NR

NR

NR

NR

Coburger et al, 2015

Prospective, monocentric, retrospective matched pair

33

GBM with intended GTR

Eloquent and noneloquent

To assess impact of additional use of 5-ALA in iMRI-assisted surgery of GBMs on EOR, PFS, OS

20 mg/kg BW

One patient sunburn

NR

5-ALA and iMRI: 100%, iMRI alone: 82%

NR

NR

NR

NR

OS: 5-ALA and iMRI: 18 mo, iMRI: 17 mo;

PFS 5-ALA and IMRI: 6 mo, iMRI: 6 mo

Lau et al, 2016

Prospective, monocentric, phase II

59

47 GBMs

12 grade III gliomas

primary and recurrent

Eloquent and noneloquent

To examine correlation of intensity of 5-ALA fluorescence with degree of tumor cellularity

20 mg/kg BW

Hypotension (two patients); mild rash (one patient)

None

NR

GBM: 84.2%

GBM 62.1%

GBM: 97.2%

HGG: 100%

GBM: 43.9%, HGG: 16.7%

NR

Summary

Mainly prospective, monocentric cohort

Total: 497

Primary and recurrent HGGs

Eloquent and noneloquent

Safety, feasibility, EOR, GTR, PFS, OS, histopathologic correlation, correlation with MRI, combination with mapping and iMRI

20 mg/kg BW

Rare; erythema, mild elevation of liver enzymes in single cases

Neuronavigation, ultrasound, IOM, mapping, awake surgery

19.4–100%

∼ 85%

62–100%

> 90%

39.5–66%

Improved 6-mo PFS

Abbreviations: 5-ALA, 5-aminolevulinic acid; AA, anaplastic astrocytoma; BW, body weight; CRET, complete resection of enhancing tumor; EOR, extent of resection; fl, fluorescence; GBM, glioblastoma multiforme; GTR, gross total resection; HGG, high-grade glioma; iMRI, intraoperative magnetic resonance imaging; IOM, intraoperative monitoring; MRI, magnetic resonance imaging; NPV, negative predictive value; NR, not recorded; OS, overall survival; PFS, progression-free survival; PPV, positive predictive value; RCT, randomized controlled trial; WL, white light.



#

Fluorescein

Fluorescein sodium, originally and still widely used in ophthalmic surgery for retinal angiography, was introduced into the field of neurosurgery by George E. Moore in 1947, and it was shown to highlight areas of blood-brain barrier (BBB) disruption linked to tumor growth after intravenous application.[27] [49] Fluorescein has a characteristic yellow-green fluorescence, with a peak absorption between 465 and 480 nm and an emission peak at 500 to 530 nm (i.e., in the yellow/green range). When administered in high concentrations, fluorescein fluorescence can even be observed under white light.[50] Fluorescein is considered a safe, robust, and inexpensive fluorophore. In some cases, it leads to transient discoloration of urine and skin after administration, and anaphylactic reactions have been described in a few cases.

However, no severe adverse events have been described using the recommended dosage of 3 to 5 mg/kg BW. Fluorescein is administered intravenously just after induction of anesthesia.[51] [52] [53] It is distributed via the bloodstream and then extravasates through the disrupted BBB, highlighting regions of the brain with abnormal vasculature, neovascularization, or increased vascular permeability.[27] [54] In malignant gliomas that are characterized by a disruption of the BBB, fluorescein accumulates in the extracellular space of the tumor tissue. In 1998, Kuroiwa et al introduced an operative microscope equipped with emission filters to visualize fluorescein under yellow-filtered (560 nm) light.[55] Today, various fluorescent filters for visualizing fluorescence are available and incorporated into modern surgical microscopes (e.g., the FL560 System [Leica Microscopes, Wetzlar, Germany] and YELLOW 560 system [Carl Zeiss, Dublin, California, United States]) ([Fig. 2]).

Encouraged by the success of 5-ALA, several studies analyzed the efficacy and applicability of the comparably less expensive agent fluorescein for resection of malignant gliomas, indicating a propensity for improving EOR ([Table 2]).[51] [54] [56] [57] [58] [59] However, many of these studies are retrospective, and none of these studies are randomized and may be confounded by case selection.

Table 2

Overview on studies analyzing the use of fluorescein for glioma surgery

Study

Study design

No. of patients

Tumor type

Eloquence

Primary end point

Fluorescein dosage

Drug-relate side effects

Other intraoperative tools

GTR

Sensitivity

Specificity

PPV

NPV

Impact on survival

Koc et al, 2008

Prospective, monocentric, controlled not randomized

47

(control: 33)

GBM

NR

GTR

20 mg/kg BW

NR

NR

Fluorescein: 83%

Control group: 55%

NR

NR

NR

NR

Median survival fluorescein: 44 wk

Control group: 42 wk (NS)

Chen et al, 2012

Prospective, monocentric, controlled, not randomized

10

(control: 12)

3 GBMs,

3 AAs

4 grade II

Eloquent and noneloquent

Reevaluate the utility and clinical limitations of using fluorescein sodium for treatment and resection of glioma brain tumors

15–20 mg/kg BW

Yellow staining of sclera, skin, and urine disappeared within 24 h

NR

Fluorescein group:80%

Control group: 33.3%

p = 0.047

NR

NR

NR

NR

Fluorescein: PFS 7.2 mo

Control: 5.4 mo (p = 0.033)

Schebesch et al, 2013

Retrospective, monocentric

26

17 GBMs

5 AAs

3 grade II

1 grade I

primary and recurrent

Eloquent and noneloquent

Feasibility and efficacy of fluorescein under YELLOW 560 nm, safety

3–4 mg/kg BW

None

NR

80%

NR

NR

NR

NR

NR

Acerbi et al, 2014

Prospective phase II trial

20

19 GBMs

1 AA

all amenable to complete resection

Eloquent and noneloquent

Evaluating the safety of fluorescein-guided surgery for HGGs and obtaining preliminary evidence regarding its efficacy for this purpose

5–10 mg/kg BW

None

Neuronavigation was allowed only for surgical planning, initial tumor localization, and orientation during tumor removal

Neurophysiologic monitoring

80%

94%

89.5%

NR

NR

NR

Diaz et al, 2015

Prospective, monocentric

12

9 primary GBMs; 3 recurrent GBMs

Eloquent and noneloquent

Ability of fluorescein to specifically stain glioma cells

3 mg/kg BW

NR

Neuronavigation

100%

82.2%

90.9%

NR

NR

NR

Catapano et al, 2017

Retrospective, monocentric, matched pair analysis

23

(control: 25)

Primary GBM

Eloquent and noneloquent

GTR

5 mg/kg BW

None

Neuronavigation

82.6%

(control group: 52%)

p = 0.03

84%

95%

NR

NR

NR

Francaviglia et al, 2017

Retrospective, monocentric

47

33 GBMs

14 AAs

All primary

Eloquent and noneloquent

Safety and EOR

5 mg/kg BW

Yellow staining of sclera, skin, and urine disappeared within 24 h

Neuronavigation

53.2%

NR

NR

NR

NR

NR

Summary

Both prospective and retrospective cohorts; no RCTs

185

Primary and recurrent HGGs

Eloquent and noneloquent

Safety, GTR, feasibility

3–20 mg/kg BW, mainly 5 mg/kg BW

Only minor, temporary staining of urine and sclera

Neuronavigation and IOM

53–100%

92–94%

89–95%

NR

NR

Longer PFS in fluorescein group shown in one study

Abbreviations: AA, anaplastic astrocytoma; BW, body weight; EOR, extent of resection; fl, fluorescence; GBM, glioblastoma multiforme; GTR, gross total resection; HGG, high-grade glioma; IOM, intraoperative monitoring; NPV, negative predictive value; NR, not recorded; NS, not significant; PFS, progression-free survival; PPV, positive predictive value; RCT, randomized controlled trial.


Several groups reported GTR rates of 80% using the YELLOW 560 filter.[54] [56] Diaz et al reported GTR in 100% when using fluorescein in their cohort of 12 glioblastoma patients and demonstrated a good correlation between intraoperative fluorescence and contrast enhancement on MRI.[51] However, the authors emphasized that the accumulation of fluorescein in malignant glioma is related to the passage through the disrupted BBB and cannot be attributed to a specific uptake by the tumor itself, as is the case for 5-ALA.[51] [56]

In addition, an analysis of the literature shows no clear consensus about dosage and timing of administration of fluorescein before surgery, although timing seems to be critical because extravasation and distribution of fluorescein follow a certain time course. Intravascular fluorescein will be extravasated after a half-life of 264 minutes and might stain edema in peritumoral normal brain parenchyma as well, increasing the danger of resection of nontumorous tissue.[60] Timing of administration should be planned carefully to minimize these confounders. Furthermore, surgical manipulation of brain tissue will per se disrupt the BBB, leading to unselective extravasation of fluorescein from the bloodstream along the cutting margins, also jeopardizing confident delineation between tumor and normal tissue. Therefore, fluorescein is rather a marker of BBB integrity than a specific tumor-targeting tool.[61] This aspect has to be kept in mind when using this agent.

So far, no studies have revealed reliable data on the effects of fluorescein-guided resection on outcome and survival. Two prospective controlled studies evaluated the effect on survival. One small study described an improved PFS when using fluorescein (7.2 months versus 5.4 months; p = 0.033). However, the study lacked randomization and did not use special microscope filters to visualize fluorescence, using only white light.[57] [59] A phase II trial (FLUGLIO) evaluated the safety and efficacy of fluorescein in glioma surgery and showed that fluorescein is feasible and safe, allowing complete tumor resection in a high percentage of cases.[62] Nevertheless, further prospective randomized controlled studies are warranted to investigate the benefit of fluorescein for EOR and outcome in glioma patients.

The simultaneous use of 5-ALA and fluorescein was shown to be feasible in glioblastoma surgery. 5-ALA was used to stain the tumor and fluorescein to provide tissue fluorescence of adjacent brain, leading to highly specific tumor visualization as well as enhanced background brightness at the same time.[63]


#

Indocyanine Green

ICG is a tricarbocyanine with fluorescence in the near-infrared range (NIR) and was approved by the FDA in 1959 for the diagnosis of liver function. ICG has been widely used in ophthalmology. It has a peak emission at 780 nm and excitation at 810 nm.[64] [65] ICG is considered safe with a low incidence of adverse side effects such as hypotension, arrhythmia, and anaphylactic shock in 0.05%, and mild symptoms such as nausea or skin eruptions in 0.2%.[29]

The use of ICG in neurosurgery was first described by Raabe et al for visualization of blood flow in cerebral vessels under the surgical microscope, and it is now a frequently used technique in the surgery of aneurysms and other vascular malformations.[29] [66] [67]

Recently, ICG was used for visualization of malignant gliomas using a technique referred to as second window ICG (SWIG). Twenty-four hours before surgery, 5 mg/kg BW ICG are administered to the patient, leading to the accumulation in tumor tissue mainly due to enhanced permeability and retention effects.[68] [69] A NIR camera (NIR light range: 700–850 nm), integrated into the surgical microscope, is used to visualize the tumor at an emission of 780 to 950 nm.

Compared with 5-ALA and fluorescein, which both emit fluorescence within the visible spectrum, ICG has excitation and emission in the NIR region of the spectrum. This advantage enables visualization of ICG fluorescence even in deeper regions, up to 3 cm, and also through the dura. This circumstance helps in planning a precise durotomy and corticotomy.[70]

A pilot study evaluating SWIG in 15 patients with gliomas revealed strong tumor-to-background fluorescence ratios, and a good correlation of contrast enhancement on MRI with intraoperative fluorescence. However, the specificity was very low, 45%, indicating possible illumination of adjacent edema.[68] Up to now, no studies have evaluated the benefit of ICG regarding improvement of EOR or outcome in treatment of gliomas, and further research is warranted to assess the usefulness of ICG.


#

Novel Techniques for Targeting Glioma Cells

Fluorescence-guided surgery has to date been widely implemented in the daily routine for glioblastoma surgery. However, in its present form there are some limitations regarding the sensitivity for visualization of tumor cells. Consequently, these techniques are being further improved, and other methods, some of them still in the fledgling stages, are undergoing intense research.


#

Tumor-Targeting Alkylphosphocholine Analogs

Alkylphosphocholine analogs (APCs) are small synthetic phospholipid ether molecules with a purported broad tumor-targeting potential because they are known to be taken up by malignant cells thorough overexpressed lipid rafts. Due to decreased catabolism in cancer cells, APCs undergo prolonged retention.[71] [72] In a glioblastoma xenograft mouse model, Swanson et al showed that two fluorescent APCs (CLR1501 green fluorescence and CLR1502 near-infrared fluorescence) are capable of labeling glioblastoma cells with high tumor-to-normal parenchyma.[69] Further research aims at developing dual-labeled APCs enabling fluorescence-guided visualization and PET imaging with the same agent. Despite still being under investigation in a preclinical status, this technique offers the possibility of targeting and treating glioblastoma at different phases of the disease: resection, staging, and possibly localized radiotherapy.[11] [73]


#

BLZ-100 Fluorescence-Guided Brain Tumor Surgery

BLZ-100 (tozuleristide) consists of the tumor-targeting peptide chlorotoxin, extracted from the venom of scorpions, with assumed specific binding to gliomas, conjugated with the near-infrared fluorophore ICG.[9] [74] The agent is administered 24 hours before surgery, and fluorescence is visualized using a NIR camera. Butte et al demonstrated a high affinity of BLZ-100 toward glioblastomas in a mouse model.[75] Further studies are needed to determine the role of this technique as a further step toward using fluorescent-labeled probes with tumor-specific molecular targets to visualize glioma cells with higher accuracy in the clinical setting of glioblastoma surgery, and early-phase clinical studies are underway.


#

Confocal Endomicroscopy

Major limitations of FGS are the lack of high resolution and the subjective interpretation of fluorescence qualities. Especially at the tumor margin, delineation of tumor tissue from normal brain is often challenging, and prediction of histologic tumor grading from preoperative imaging is often not possible. Intraoperative frozen sections are frequently performed to acquire immediate diagnosis. However, this procedure is time consuming and can be nondiagnostic or even misleading in certain cases.[76] Confocal endomicroscopy is a technique that was recently introduced into the field of neurosurgery. Images are acquired using a handheld probe that has a single optical fiber for illumination and detection. The images are displayed in high resolution in up to 1,000-fold magnification to an LCD workstation. To provide tissue contrast, fluorescent agents like fluorescein are administered.[10] [77] [78] Confocal endomicroscopy allows real-time visualization of malignant cells and is particularly useful for scanning the tumor margin for residual tumor tissue with high accuracy to enhance EOR and at the same time lower the risk of resection of nontumorous tissue in eloquent areas leading to possible neurologic deficits. For interpretation of acquired images, profound histopathologic knowledge or the presence of a neuropathologist is required.[79]


#

Raman Spectroscopy

Raman spectroscopy is based on the Raman effect, first described by C.V. Raman in 1928, and refers to the scattering of monochromatic light in tissue. Most photons in the visible spectrum are scattered elastically, implying they have the same level of energy when interacting with a tissue or object. However, some photons transfer or absorb energy to or from the object being imaged, resulting in a transmission of energy. This phenomenon is called inelastic scattering and known as the Raman effect.[80] With the help of a spectrometer (Raman spectroscopy), information regarding the chemical composition of different tissues, for example, the amount and ratios of lipid and protein, can be obtained. These data provide a unique biochemical signature of the tissue and enable delineation between different tissues. In comparison with other techniques, Raman spectroscopy is a label-free visualization method that depends on intrinsic biochemical properties of different tissues to provide image contrast.[81] This technique was shown to be effective in delineation of glioblastoma, necrosis, and normal brain parenchyma as well.[8] [82] Normal, necrotic, and glioblastoma tissue was distinguished by Raman spectroscopy in frozen sections with 99.5% accuracy.[8] Jermyn et al used a Raman spectroscopy handheld probe system intraoperatively and found an accuracy of 92% for glioma detection.[83] Similar to confocal endomicroscopy, this technique enables intraoperative tissue analysis before resection and is a promising guide for surgical resection and decision making.[83] [84]


#

Conclusion and Future Perspective

In summary, several intraoperative imaging methods aiming at improvement of intraoperative glioma targeting and visualization are presently available. Neurosurgeons have started to integrate these techniques into their daily routine for glioma surgery. The ultimate purpose of these methods is to increase the EOR while keeping the risk for postoperative neurologic deterioration low.

Still, there are limitations, as discussed earlier and listed in [Table 3], that have to be considered when applying one of these techniques. To overcome these limitations, further research is being performed. One approach is the combination of different techniques, such as neuronavigation and FGS, allowing the generation of comprehensive information on tumor extent, anatomy, and metabolism. Adding newer techniques, like Raman spectroscopy or targeted fluorescence, further information regarding chemical and metabolic composition of the tissue will be provided.

Table 3

Overview of current techniques for intraoperative visualization of glioblastoma cells with their advantages and disadvantages

Technique

Publications

Principle

Application/ development

Advantages

Disadvantages

Neuronavigation

Maciunas et al, 1996

Jung et al, 2006

Wirtz et al, 2000

Orringer et al, 2012

Preoperative images, intraoperative orientation

Widespread use in clinical setting

– Maintaining orientation

– Visualization of anatomy

– Planning surgical approach

– Combination with other tools

– Access to various preoperative images including PET, CT, MRI, data for fiber tracking

– Brain shift, loss of accuracy

– Relies on preoperative imaging, not real time

– Interruption if surgical workflow

Ultrasound

Mercier et al, 2011

Saether et al, 2012

Intraoperative imaging

Widespread use in clinical setting

– Dynamic, cheap, and easy to use

– Provides intraoperative real-time images

– May be used to update navigation system

– Low resolution

iMRI

Hatiboglu et al, 2009

Senft et al, 2011

Liang et al, 2012

Ozduman et al, 2014

Coburger et al, 2014

Coburger et al, 2015

Intraoperative imaging

Widespread use in clinical setting

– Almost real-time images during surgery

– Identification of residual tumor

– Expensive

– Time consuming

– Extends duration of surgery and anesthesia

5-ALA

Stummer et al, 1998

Stummer et al, 2000

Stummer et al, 2006

Nabavi et al, 2009

Díez Valle et al, 2011

Della Puppa et al, 2012

Schucht et al, 2012

Della Puppa et al, 2012

Stummer et al, 2014

Schucht et al, 2014

Lau et al, 2016

Metabolic

Widespread use in clinical setting, FDA and EMA approval

– Selectively absorbed by tumor cells

– Low toxicity, high safety

– Intraoperative real-time imaging

– Full integration into the surgical microscope and view of full surgical field

– Use without interruption to the surgical workflow

– Reliable correlation with preoperative contrast enhancement on MRI

– Correlation with histopathology

– Brain shift is no concern

– Low background illumination

 o Alternating between white light and fluorescence mode

– Imaging surface tool, depth can limit visualization

– Requires special microscope

– Expensive

– Bleaching effect

– Time dependency

– Subjective interpretation of fluorescence intensities

Fluorescein

Koc et al, 2008

Chen et al, 2012

Schebesch et al, 2013

Acerbi et al, 2014

Diaz et al, 2015

Francaviglia et al, 2017

Catapano et al, 2017

Permeability of BBB

Human use, off-label

– Robust, safe, cheap

– Can be visualized under white light (using higher concentrations)

– Intraoperative real-time imaging

– Full integration into the surgical microscope and view of the full surgical field

– Use without interruption to the surgical workflow

– Brain shift is no concern

– Not tumor cell specific

 o Marker of BBB breakdown

– Unselective extravasation during surgery

– Time dependency

– Subjective interpretation of fluorescence intensities

ICG

Lee et al, 2016

Permeability of BBB

Human use, off-label

– Excitation and emission in the near-infrared region

 O Enables visualization of fluorescence situated deeper in the tissue

– Low toxicity, high safety

– Intraoperative real-time imaging

– Brain shift does not interfere with this technique

– Full integration into the surgical microscope and view of the full surgical field

– Use without interruption to the surgical workflow

– Brain shift is no concern

– Requires special cameras to visualize fluorescence

– Not tumor specific

– Accumulates due to an enhanced permeability of the BBB

– Time dependency

– Subjective interpretation of fluorescence intensities

Tumor-targeted alkylphosphocholine analogs

Swanson et al, 2015

Tumor-targeted

Animal model

– Specific detection of tumor cells

BLZ-100 (tozuleristide)

Butte et al, 2013

Tumor-targeted

Animal model

– Tumor-specific molecular targets

Confocal endomicroscopy

Hoffman et al, 2006

Foersch et al, 2012

Intraoperative microscopy, fluorescence labeling

Human use, clinical trials

– Intraoperative neuropathologic diagnostic

– High resolution in up to 1,000-fold magnification

– High accuracy

– Only small field can be analyzed at the same time

– Time consuming

– Presence of a neuropathologist required to interpret images

Raman spectroscopy

Krafft et al, 2004

Kalkanis et al, 2014

Jermyn et al, 2015

Intrinsic biochemical properties of different tissues

Human use, clinical trials

– Unique biochemical signature

– High accuracy

– Only small field can be analyzed at the same time

– Time consuming

Abbreviations: 5-ALA, 5-aminolevulinic acid; BBB, blood-brain barrier; CT, computed tomography; EMA, European Medicines Agency; FDA, Food and Drug Administration; ICG, indocyanine green; iMRI, intraoperative magnetic resonance imaging; MRI, magnetic resonance imaging; PET, positron emission tomography.


5-ALA appears to be the only available intraoperative tool for direct identification of glioblastoma cells. It has further shown a good correlation with regions of higher metabolic activity in tumor, similar to FET (fluoroethyltyrosine)-PET, although these PET hot spots often cannot be matched on MRI.[85] [86] In addition, a higher Ki-67/MIB-1 index and other features of malignancy correlate with the amount of 5-ALA fluorescence observed.[87] [88] Most randomized controlled trials are based on the gadolinium-based assessment of residual tumor and EOR. For the future, the EOR based on 5-ALA–induced fluorescence might be a more accurate marker.

Currently, the intensity of fluorescence relies on the subjective interpretation of the surgeon. To quantify fluorescence, further attempts have been undertaken, for example, using spectroscopic techniques to determine intraoperative protoporphyrin (Pp) IX concentration in tumor tissue via a handheld device, even in cases with no visible fluorescence under the surgical microscope.[89] For low-grade glioma, where fluorescence is often not visible using standard surgical microscopy, a 100-fold increase in sensitivity of fluorescence detection using handheld spectroscopy can be achieved, resulting in detection of PpIX fluorescence in these slowly growing tumors also.[90]

Targeted fluorescence imaging will soon be available, together with innovations in neurosurgical microscope technology, to help detect optical features in gliomas presently invisible to the human eye.[70] Such technologies will help overcome the limitations of the sensitivity and specificity of the present methods.


#
#

Conflict of Interest

None declared.

  • References

  • 1 Koshy M, Villano JL, Dolecek TA. , et al. Improved survival time trends for glioblastoma using the SEER 17 population-based registries. J Neurooncol 2012; 107 (01) 207-212
  • 2 Ostrom QT, Gittleman H, Truitt G, Boscia A, Kruchko C, Barnholtz-Sloan JS. CBTRUS statistical report: primary brain and other central nervous system tumors diagnosed in the United States in 2011-2015. Neuro Oncol 2018; 20 (04) (Suppl. 04) iv1-iv86
  • 3 Stupp R, Mason WP, van den Bent MJ. , et al; European Organisation for Research and Treatment of Cancer Brain Tumor and Radiotherapy Groups; National Cancer Institute of Canada Clinical Trials Group. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med 2005; 352 (10) 987-996
  • 4 Lacroix M, Abi-Said D, Fourney DR. , et al. A multivariate analysis of 416 patients with glioblastoma multiforme: prognosis, extent of resection, and survival. J Neurosurg 2001; 95 (02) 190-198
  • 5 Sanai N, Polley MY, McDermott MW, Parsa AT, Berger MS. An extent of resection threshold for newly diagnosed glioblastomas. J Neurosurg 2011; 115 (01) 3-8
  • 6 Stupp R, Hegi ME, Mason WP. , et al; European Organisation for Research and Treatment of Cancer Brain Tumour and Radiation Oncology Groups; National Cancer Institute of Canada Clinical Trials Group. Effects of radiotherapy with concomitant and adjuvant temozolomide versus radiotherapy alone on survival in glioblastoma in a randomised phase III study: 5-year analysis of the EORTC-NCIC trial. Lancet Oncol 2009; 10 (05) 459-466
  • 7 Desroches J, Jermyn M, Mok K. , et al. Characterization of a Raman spectroscopy probe system for intraoperative brain tissue classification. Biomed Opt Express 2015; 6 (07) 2380-2397
  • 8 Kalkanis SN, Kast RE, Rosenblum ML. , et al. Raman spectroscopy to distinguish grey matter, necrosis, and glioblastoma multiforme in frozen tissue sections. J Neurooncol 2014; 116 (03) 477-485
  • 9 Lyons SA, O'Neal J, Sontheimer H. Chlorotoxin, a scorpion-derived peptide, specifically binds to gliomas and tumors of neuroectodermal origin. Glia 2002; 39 (02) 162-173
  • 10 Martirosyan NL, Cavalcanti DD, Eschbacher JM. , et al. Use of in vivo near-infrared laser confocal endomicroscopy with indocyanine green to detect the boundary of infiltrative tumor. J Neurosurg 2011; 115 (06) 1131-1138
  • 11 Zhang RR, Swanson KI, Hall LT, Weichert JP, Kuo JS. Diapeutic cancer-targeting alkylphosphocholine analogs may advance management of brain malignancies. CNS Oncol 2016; 5 (04) 223-231
  • 12 Maciunas RJ, Berger MS, Copeland B, Mayberg MR, Selker R, Allen GS. A technique for interactive image-guided neurosurgical intervention in primary brain tumors. Neurosurg Clin N Am 1996; 7 (02) 245-266
  • 13 Jung TY, Jung S, Kim IY. , et al. Application of neuronavigation system to brain tumor surgery with clinical experience of 420 cases. Minim Invasive Neurosurg 2006; 49 (04) 210-215
  • 14 Schulz C, Waldeck S, Mauer UM. Intraoperative image guidance in neurosurgery: development, current indications, and future trends. Radiol Res Pract 2012; 2012: 197364
  • 15 Wirtz CR, Albert FK, Schwaderer M. , et al. The benefit of neuronavigation for neurosurgery analyzed by its impact on glioblastoma surgery. Neurol Res 2000; 22 (04) 354-360
  • 16 Orringer DA, Golby A, Jolesz F. Neuronavigation in the surgical management of brain tumors: current and future trends. Expert Rev Med Devices 2012; 9 (05) 491-500
  • 17 Sæther CA, Torsteinsen M, Torp SH, Sundstrøm S, Unsgård G, Solheim O. Did survival improve after the implementation of intraoperative neuronavigation and 3D ultrasound in glioblastoma surgery? A retrospective analysis of 192 primary operations. J Neurol Surg A Cent Eur Neurosurg 2012; 73 (02) 73-78
  • 18 Rygh OM, Selbekk T, Torp SH, Lydersen S, Hernes TA, Unsgaard G. Comparison of navigated 3D ultrasound findings with histopathology in subsequent phases of glioblastoma resection. Acta Neurochir (Wien) 2008; 150 (10) 1033-1041 ; discussion 1042
  • 19 Mursch K, Scholz M, Brück W, Behnke-Mursch J. The value of intraoperative ultrasonography during the resection of relapsed irradiated malignant gliomas in the brain. Ultrasonography 2017; 36 (01) 60-65
  • 20 Mercier L, Del Maestro RF, Petrecca K. , et al. New prototype neuronavigation system based on preoperative imaging and intraoperative freehand ultrasound: system description and validation. Int J CARS 2011; 6 (04) 507-522
  • 21 Liang D, Schulder M. The role of intraoperative magnetic resonance imaging in glioma surgery. Surg Neurol Int 2012; 3 (Suppl. 04) S320-S327
  • 22 Jolesz FA. Intraoperative imaging in neurosurgery: where will the future take us?. Acta Neurochir Suppl (Wien) 2011; 109: 21-25
  • 23 Sutherland GR, Kaibara T, Louw D, Hoult DI, Tomanek B, Saunders J. A mobile high-field magnetic resonance system for neurosurgery. J Neurosurg 1999; 91 (05) 804-813
  • 24 Senft C, Bink A, Franz K, Vatter H, Gasser T, Seifert V. Intraoperative MRI guidance and extent of resection in glioma surgery: a randomised, controlled trial. Lancet Oncol 2011; 12 (11) 997-1003
  • 25 Hatiboglu MA, Weinberg JS, Suki D. , et al. Impact of intraoperative high-field magnetic resonance imaging guidance on glioma surgery: a prospective volumetric analysis. Neurosurgery 2009; 64 (06) 1073-1081 ; discussion 1081
  • 26 Özduman K, Yıldız E, Dinçer A, Sav A, Pamir MN. Using intraoperative dynamic contrast-enhanced T1-weighted MRI to identify residual tumor in glioblastoma surgery. J Neurosurg 2014; 120 (01) 60-66
  • 27 Moore GE. Fluorescein as an agent in the differentiation of normal and malignant tissues. Science 1947; 106 (2745): 130-131
  • 28 Stummer W, Stocker S, Wagner S. , et al. Intraoperative detection of malignant gliomas by 5-aminolevulinic acid-induced porphyrin fluorescence. Neurosurgery 1998; 42 (03) 518-525 ; discussion 525–526
  • 29 Raabe A, Beck J, Gerlach R, Zimmermann M, Seifert V. Near-infrared indocyanine green video angiography: a new method for intraoperative assessment of vascular flow. Neurosurgery 2003; 52 (01) 132-139 ; discussion 139
  • 30 Colditz MJ, Leyen Kv, Jeffree RL. Aminolevulinic acid (ALA)-protoporphyrin IX fluorescence guided tumour resection. Part 2: theoretical, biochemical and practical aspects. J Clin Neurosci 2012; 19 (12) 1611-1616
  • 31 Stummer W, Stepp H, Möller G, Ehrhardt A, Leonhard M, Reulen HJ. Technical principles for protoporphyrin-IX-fluorescence guided microsurgical resection of malignant glioma tissue. Acta Neurochir (Wien) 1998; 140 (10) 995-1000
  • 32 Stummer W, Stocker S, Novotny A. , et al. In vitro and in vivo porphyrin accumulation by C6 glioma cells after exposure to 5-aminolevulinic acid. J Photochem Photobiol B 1998; 45 (2–3): 160-169
  • 33 Kaneko S, Suero Molina E, Ewelt C, Warneke N, Stummer W. Fluorescence-based measurement of real-time kinetics of protoporphyrin IX after 5-Aminolevulinic acid administration in human in situ malignant gliomas. Neurosurgery. In press
  • 34 Stummer W, Pichlmeier U, Meinel T, Wiestler OD, Zanella F, Reulen HJ. ; ALA-Glioma Study Group. Fluorescence-guided surgery with 5-aminolevulinic acid for resection of malignant glioma: a randomised controlled multicentre phase III trial. Lancet Oncol 2006; 7 (05) 392-401
  • 35 Teixidor P, Arráez MA, Villalba G. , et al. Safety and efficacy of 5-aminolevulinic acid for high grade glioma in usual clinical practice: a prospective cohort study. PLoS One 2016; 11 (02) e0149244
  • 36 Eljamel S. 5-ALA fluorescence image guided resection of glioblastoma multiforme: a meta-analysis of the literature. Int J Mol Sci 2015; 16 (05) 10443-10456
  • 37 Stummer W, Novotny A, Stepp H, Goetz C, Bise K, Reulen HJ. Fluorescence-guided resection of glioblastoma multiforme by using 5-aminolevulinic acid-induced porphyrins: a prospective study in 52 consecutive patients. J Neurosurg 2000; 93 (06) 1003-1013
  • 38 Stummer W, Tonn JC, Goetz C. , et al. 5-Aminolevulinic acid-derived tumor fluorescence: the diagnostic accuracy of visible fluorescence qualities as corroborated by spectrometry and histology and postoperative imaging. Neurosurgery 2014; 74 (03) 310-319 ; discussion 319–320
  • 39 Díez Valle R, Tejada Solis S, Idoate Gastearena MA, García de Eulate R, Domínguez Echávarri P, Aristu Mendiroz J. Surgery guided by 5-aminolevulinic fluorescence in glioblastoma: volumetric analysis of extent of resection in single-center experience. J Neurooncol 2011; 102 (01) 105-113
  • 40 Nabavi A, Thurm H, Zountsas B. , et al; 5-ALA Recurrent Glioma Study Group. Five-aminolevulinic acid for fluorescence-guided resection of recurrent malignant gliomas: a phase ii study. Neurosurgery 2009; 65 (06) 1070-1076 ; discussion 1076–1077
  • 41 Lau D, Hervey-Jumper SL, Chang S. , et al. A prospective Phase II clinical trial of 5-aminolevulinic acid to assess the correlation of intraoperative fluorescence intensity and degree of histologic cellularity during resection of high-grade gliomas. J Neurosurg 2016; 124 (05) 1300-1309
  • 42 Schucht P, Beck J, Abu-Isa J. , et al. Gross total resection rates in contemporary glioblastoma surgery: results of an institutional protocol combining 5-aminolevulinic acid intraoperative fluorescence imaging and brain mapping. Neurosurgery 2012; 71 (05) 927-935 ; discussion 935–936
  • 43 Della Puppa A, De Pellegrin S, d'Avella E. , et al. 5-aminolevulinic acid (5-ALA) fluorescence guided surgery of high-grade gliomas in eloquent areas assisted by functional mapping. Our experience and review of the literature. Acta Neurochir (Wien) 2013; 155 (06) 965-972 ; discussion 972
  • 44 Schucht P, Seidel K, Beck J. , et al. Intraoperative monopolar mapping during 5-ALA-guided resections of glioblastomas adjacent to motor eloquent areas: evaluation of resection rates and neurological outcome. Neurosurg Focus 2014; 37 (06) E16
  • 45 Aldave G, Tejada S, Pay E. , et al. Prognostic value of residual fluorescent tissue in glioblastoma patients after gross total resection in 5-aminolevulinic Acid-guided surgery. Neurosurgery 2013; 72 (06) 915-920 ; discussion 920–921
  • 46 Schucht P, Knittel S, Slotboom J. , et al. 5-ALA complete resections go beyond MR contrast enhancement: shift corrected volumetric analysis of the extent of resection in surgery for glioblastoma. Acta Neurochir (Wien) 2014; 156 (02) 305-312 ; discussion 312
  • 47 Coburger J, Hagel V, Wirtz CR, König R. Surgery for glioblastoma: impact of the combined use of 5-aminolevulinic acid and intraoperative MRI on extent of resection and survival. PLoS One 2015; 10 (06) e0131872
  • 48 Coburger J, Engelke J, Scheuerle A. , et al. Tumor detection with 5-aminolevulinic acid fluorescence and Gd-DTPA-enhanced intraoperative MRI at the border of contrast-enhancing lesions: a prospective study based on histopathological assessment. Neurosurg Focus 2014; 36 (02) E3
  • 49 Rabb MF, Burton TC, Schatz H, Yannuzzi LA. Fluorescein angiography of the fundus: a schematic approach to interpretation. Surv Ophthalmol 1978; 22 (06) 387-403
  • 50 Shinoda J, Yano H, Yoshimura S. , et al. Fluorescence-guided resection of glioblastoma multiforme by using high-dose fluorescein sodium. Technical note. J Neurosurg 2003; 99 (03) 597-603
  • 51 Diaz RJ, Dios RR, Hattab EM. , et al. Study of the biodistribution of fluorescein in glioma-infiltrated mouse brain and histopathological correlation of intraoperative findings in high-grade gliomas resected under fluorescein fluorescence guidance. J Neurosurg 2015; 122 (06) 1360-1369
  • 52 Hamamcıoğlu MK, Akçakaya MO, Göker B, Kasımcan MO, Kırış T. The use of the YELLOW 560 nm surgical microscope filter for sodium fluorescein-guided resection of brain tumors: our preliminary results in a series of 28 patients. Clin Neurol Neurosurg 2016; 143: 39-45
  • 53 Schebesch KM, Brawanski A, Hohenberger C, Hohne J. Fluorescein sodium-guided surgery of malignant brain tumors: history, current concepts, and future project. Turk Neurosurg 2016; 26 (02) 185-194
  • 54 Schebesch KM, Proescholdt M, Höhne J. , et al. Sodium fluorescein-guided resection under the YELLOW 560 nm surgical microscope filter in malignant brain tumor surgery—a feasibility study. Acta Neurochir (Wien) 2013; 155 (04) 693-699
  • 55 Kuroiwa T, Kajimoto Y, Ohta T. Development of a fluorescein operative microscope for use during malignant glioma surgery: a technical note and preliminary report. Surg Neurol 1998; 50 (01) 41-48 ; discussion 48–49
  • 56 Acerbi F, Broggi M, Eoli M. , et al. Is fluorescein-guided technique able to help in resection of high-grade gliomas?. Neurosurg Focus 2014; 36 (02) E5
  • 57 Chen B, Wang H, Ge P. , et al. Gross total resection of glioma with the intraoperative fluorescence-guidance of fluorescein sodium. Int J Med Sci 2012; 9 (08) 708-714
  • 58 Francaviglia N, Iacopino DG, Costantino G. , et al. Fluorescein for resection of high-grade gliomas: a safety study control in a single center and review of the literature. Surg Neurol Int 2017; 8: 145
  • 59 Koc K, Anik I, Cabuk B, Ceylan S. Fluorescein sodium-guided surgery in glioblastoma multiforme: a prospective evaluation. Br J Neurosurg 2008; 22 (01) 99-103
  • 60 Stummer W. Poor man's fluorescence?. Acta Neurochir (Wien) 2015; 157 (08) 1379-1381
  • 61 Schwake M, Stummer W, Suero Molina EJ, Wölfer J. Simultaneous fluorescein sodium and 5-ALA in fluorescence-guided glioma surgery. Acta Neurochir (Wien) 2015; 157 (05) 877-879
  • 62 Acerbi F, Broggi M, Schebesch KM. , et al. Fluorescein-guided surgery for resection of high-grade gliomas: a multicentric prospective phase II study (FLUOGLIO). Clin Cancer Res 2018; 24 (01) 52-61
  • 63 Suero Molina E, Wölfer J, Ewelt C, Ehrhardt A, Brokinkel B, Stummer W. Dual-labeling with 5-aminolevulinic acid and fluorescein for fluorescence-guided resection of high-grade gliomas: technical note. J Neurosurg 2018; 128 (02) 399-405
  • 64 Cherrick GR, Stein SW, Leevy CM, Davidson CS. Indocyanine green: observations on its physical properties, plasma decay, and hepatic extraction. J Clin Invest 1960; 39: 592-600
  • 65 Reinhart MB, Huntington CR, Blair LJ, Heniford BT, Augenstein VA. Indocyanine green: historical context, current applications, and future considerations. Surg Innov 2016; 23 (02) 166-175
  • 66 Hänggi D, Etminan N, Steiger HJ. The impact of microscope-integrated intraoperative near-infrared indocyanine green videoangiography on surgery of arteriovenous malformations and dural arteriovenous fistulae. Neurosurgery 2010; 67 (04) 1094-1103 ; discussion 1103–1104
  • 67 Raabe A, Nakaji P, Beck J. , et al. Prospective evaluation of surgical microscope-integrated intraoperative near-infrared indocyanine green videoangiography during aneurysm surgery. J Neurosurg 2005; 103 (06) 982-989
  • 68 Lee JY, Thawani JP, Pierce J. , et al. Intraoperative near-infrared optical imaging can localize gadolinium-enhancing gliomas during surgery. Neurosurgery 2016; 79 (06) 856-871
  • 69 Swanson KI, Clark PA, Zhang RR. , et al. Fluorescent cancer-selective alkylphosphocholine analogs for intraoperative glioma detection. Neurosurgery 2015; 76 (02) 115-123 ; discussion 123–124
  • 70 Valdés PA, Roberts DW, Lu FK, Golby A. Optical technologies for intraoperative neurosurgical guidance. Neurosurg Focus 2016; 40 (03) E8
  • 71 Hilgard P, Klenner T, Stekar J, Unger C. Alkylphosphocholines: a new class of membrane-active anticancer agents. Cancer Chemother Pharmacol 1993; 32 (02) 90-95
  • 72 Weichert JP, Clark PA, Kandela IK. , et al. Alkylphosphocholine analogs for broad-spectrum cancer imaging and therapy. Sci Transl Med 2014; 6 (240) 240ra75
  • 73 Kuo JS, Zhang RR, Pinchuk AN, Clark PA, Weichert JP. Creation of a dual-labeled cancer-targeting alkylphosphocholine analog for dual modality quantitative positron emission tomography and intraoperative tumor visualization. Neurosurgery 2016; 63 (01) 208
  • 74 Parrish-Novak J, Byrnes-Blake K, Lalayeva N. , et al. Nonclinical profile of BLZ-100, a tumor-targeting fluorescent imaging agent. Int J Toxicol 2017; 36 (02) 104-112
  • 75 Butte PV, Mamelak A, Parrish-Novak J. , et al. Near-infrared imaging of brain tumors using the Tumor Paint BLZ-100 to achieve near-complete resection of brain tumors. Neurosurg Focus 2014; 36 (02) E1
  • 76 Uematsu Y, Owai Y, Okita R, Tanaka Y, Itakura T. The usefulness and problem of intraoperative rapid diagnosis in surgical neuropathology. Brain Tumor Pathol 2007; 24 (02) 47-52
  • 77 Foersch S, Heimann A, Ayyad A. , et al. Confocal laser endomicroscopy for diagnosis and histomorphologic imaging of brain tumors in vivo. PLoS One 2012; 7 (07) e41760
  • 78 Hoffman A, Goetz M, Vieth M, Galle PR, Neurath MF, Kiesslich R. Confocal laser endomicroscopy: technical status and current indications. Endoscopy 2006; 38 (12) 1275-1283
  • 79 Belykh E, Martirosyan NL, Yagmurlu K. , et al. Intraoperative fluorescence imaging for personalized brain tumor resection: current state and future directions. Front Surg 2016; 3: 55
  • 80 Raman C. A new type of secondary radiation. Nature 1928; 121: 501-502
  • 81 Hollon T, Lewis S, Freudiger CW, Sunney Xie X, Orringer DA. Improving the accuracy of brain tumor surgery via Raman-based technology. Neurosurg Focus 2016; 40 (03) E9
  • 82 Krafft C, Neudert L, Simat T, Salzer R. Near infrared Raman spectra of human brain lipids. Spectrochim Acta A Mol Biomol Spectrosc 2005; 61 (07) 1529-1535
  • 83 Jermyn M, Mok K, Mercier J. , et al. Intraoperative brain cancer detection with Raman spectroscopy in humans. Sci Transl Med 2015; 7 (274) 274ra19
  • 84 Desroches J, Jermyn M, Pinto M. , et al. A new method using Raman spectroscopy for in vivo targeted brain cancer tissue biopsy. Sci Rep 2018; 8 (01) 1792
  • 85 Arbizu J, Tejada S, Marti-Climent JM. , et al. Quantitative volumetric analysis of gliomas with sequential MRI and 11C-methionine PET assessment: patterns of integration in therapy planning. Eur J Nucl Med Mol Imaging 2012; 39 (05) 771-781
  • 86 Jaber M, Wölfer J, Ewelt C. , et al. The value of 5-aminolevulinic acid in low-grade gliomas and high-grade gliomas lacking glioblastoma imaging features: an analysis based on fluorescence, magnetic resonance imaging, 18F-fluoroethyl tyrosine positron emission tomography, and tumor molecular factors. Neurosurgery 2016; 78 (03) 401-411 ; discussion 411
  • 87 Idoate MA, Díez Valle R, Echeveste J, Tejada S. Pathological characterization of the glioblastoma border as shown during surgery using 5-aminolevulinic acid-induced fluorescence. Neuropathology 2011; 31 (06) 575-582
  • 88 Roberts DW, Valdés PA, Harris BT. , et al. Coregistered fluorescence-enhanced tumor resection of malignant glioma: relationships between δ-aminolevulinic acid-induced protoporphyrin IX fluorescence, magnetic resonance imaging enhancement, and neuropathological parameters. Clinical article. J Neurosurg 2011; 114 (03) 595-603
  • 89 Valdés PA, Leblond F, Kim A. , et al. Quantitative fluorescence in intracranial tumor: implications for ALA-induced PpIX as an intraoperative biomarker. J Neurosurg 2011; 115 (01) 11-17
  • 90 Valdés PA, Jacobs V, Harris BT. , et al. Quantitative fluorescence using 5-aminolevulinic acid-induced protoporphyrin IX biomarker as a surgical adjunct in low-grade glioma surgery. J Neurosurg 2015; 123 (03) 771-780

Address for correspondence

Stephanie Schipmann, MD
Department of Neurosurgery, University Hospital Münster
Albert-Schweitzer-Campus 1, Building A1, 48149 Münster
Germany   

  • References

  • 1 Koshy M, Villano JL, Dolecek TA. , et al. Improved survival time trends for glioblastoma using the SEER 17 population-based registries. J Neurooncol 2012; 107 (01) 207-212
  • 2 Ostrom QT, Gittleman H, Truitt G, Boscia A, Kruchko C, Barnholtz-Sloan JS. CBTRUS statistical report: primary brain and other central nervous system tumors diagnosed in the United States in 2011-2015. Neuro Oncol 2018; 20 (04) (Suppl. 04) iv1-iv86
  • 3 Stupp R, Mason WP, van den Bent MJ. , et al; European Organisation for Research and Treatment of Cancer Brain Tumor and Radiotherapy Groups; National Cancer Institute of Canada Clinical Trials Group. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med 2005; 352 (10) 987-996
  • 4 Lacroix M, Abi-Said D, Fourney DR. , et al. A multivariate analysis of 416 patients with glioblastoma multiforme: prognosis, extent of resection, and survival. J Neurosurg 2001; 95 (02) 190-198
  • 5 Sanai N, Polley MY, McDermott MW, Parsa AT, Berger MS. An extent of resection threshold for newly diagnosed glioblastomas. J Neurosurg 2011; 115 (01) 3-8
  • 6 Stupp R, Hegi ME, Mason WP. , et al; European Organisation for Research and Treatment of Cancer Brain Tumour and Radiation Oncology Groups; National Cancer Institute of Canada Clinical Trials Group. Effects of radiotherapy with concomitant and adjuvant temozolomide versus radiotherapy alone on survival in glioblastoma in a randomised phase III study: 5-year analysis of the EORTC-NCIC trial. Lancet Oncol 2009; 10 (05) 459-466
  • 7 Desroches J, Jermyn M, Mok K. , et al. Characterization of a Raman spectroscopy probe system for intraoperative brain tissue classification. Biomed Opt Express 2015; 6 (07) 2380-2397
  • 8 Kalkanis SN, Kast RE, Rosenblum ML. , et al. Raman spectroscopy to distinguish grey matter, necrosis, and glioblastoma multiforme in frozen tissue sections. J Neurooncol 2014; 116 (03) 477-485
  • 9 Lyons SA, O'Neal J, Sontheimer H. Chlorotoxin, a scorpion-derived peptide, specifically binds to gliomas and tumors of neuroectodermal origin. Glia 2002; 39 (02) 162-173
  • 10 Martirosyan NL, Cavalcanti DD, Eschbacher JM. , et al. Use of in vivo near-infrared laser confocal endomicroscopy with indocyanine green to detect the boundary of infiltrative tumor. J Neurosurg 2011; 115 (06) 1131-1138
  • 11 Zhang RR, Swanson KI, Hall LT, Weichert JP, Kuo JS. Diapeutic cancer-targeting alkylphosphocholine analogs may advance management of brain malignancies. CNS Oncol 2016; 5 (04) 223-231
  • 12 Maciunas RJ, Berger MS, Copeland B, Mayberg MR, Selker R, Allen GS. A technique for interactive image-guided neurosurgical intervention in primary brain tumors. Neurosurg Clin N Am 1996; 7 (02) 245-266
  • 13 Jung TY, Jung S, Kim IY. , et al. Application of neuronavigation system to brain tumor surgery with clinical experience of 420 cases. Minim Invasive Neurosurg 2006; 49 (04) 210-215
  • 14 Schulz C, Waldeck S, Mauer UM. Intraoperative image guidance in neurosurgery: development, current indications, and future trends. Radiol Res Pract 2012; 2012: 197364
  • 15 Wirtz CR, Albert FK, Schwaderer M. , et al. The benefit of neuronavigation for neurosurgery analyzed by its impact on glioblastoma surgery. Neurol Res 2000; 22 (04) 354-360
  • 16 Orringer DA, Golby A, Jolesz F. Neuronavigation in the surgical management of brain tumors: current and future trends. Expert Rev Med Devices 2012; 9 (05) 491-500
  • 17 Sæther CA, Torsteinsen M, Torp SH, Sundstrøm S, Unsgård G, Solheim O. Did survival improve after the implementation of intraoperative neuronavigation and 3D ultrasound in glioblastoma surgery? A retrospective analysis of 192 primary operations. J Neurol Surg A Cent Eur Neurosurg 2012; 73 (02) 73-78
  • 18 Rygh OM, Selbekk T, Torp SH, Lydersen S, Hernes TA, Unsgaard G. Comparison of navigated 3D ultrasound findings with histopathology in subsequent phases of glioblastoma resection. Acta Neurochir (Wien) 2008; 150 (10) 1033-1041 ; discussion 1042
  • 19 Mursch K, Scholz M, Brück W, Behnke-Mursch J. The value of intraoperative ultrasonography during the resection of relapsed irradiated malignant gliomas in the brain. Ultrasonography 2017; 36 (01) 60-65
  • 20 Mercier L, Del Maestro RF, Petrecca K. , et al. New prototype neuronavigation system based on preoperative imaging and intraoperative freehand ultrasound: system description and validation. Int J CARS 2011; 6 (04) 507-522
  • 21 Liang D, Schulder M. The role of intraoperative magnetic resonance imaging in glioma surgery. Surg Neurol Int 2012; 3 (Suppl. 04) S320-S327
  • 22 Jolesz FA. Intraoperative imaging in neurosurgery: where will the future take us?. Acta Neurochir Suppl (Wien) 2011; 109: 21-25
  • 23 Sutherland GR, Kaibara T, Louw D, Hoult DI, Tomanek B, Saunders J. A mobile high-field magnetic resonance system for neurosurgery. J Neurosurg 1999; 91 (05) 804-813
  • 24 Senft C, Bink A, Franz K, Vatter H, Gasser T, Seifert V. Intraoperative MRI guidance and extent of resection in glioma surgery: a randomised, controlled trial. Lancet Oncol 2011; 12 (11) 997-1003
  • 25 Hatiboglu MA, Weinberg JS, Suki D. , et al. Impact of intraoperative high-field magnetic resonance imaging guidance on glioma surgery: a prospective volumetric analysis. Neurosurgery 2009; 64 (06) 1073-1081 ; discussion 1081
  • 26 Özduman K, Yıldız E, Dinçer A, Sav A, Pamir MN. Using intraoperative dynamic contrast-enhanced T1-weighted MRI to identify residual tumor in glioblastoma surgery. J Neurosurg 2014; 120 (01) 60-66
  • 27 Moore GE. Fluorescein as an agent in the differentiation of normal and malignant tissues. Science 1947; 106 (2745): 130-131
  • 28 Stummer W, Stocker S, Wagner S. , et al. Intraoperative detection of malignant gliomas by 5-aminolevulinic acid-induced porphyrin fluorescence. Neurosurgery 1998; 42 (03) 518-525 ; discussion 525–526
  • 29 Raabe A, Beck J, Gerlach R, Zimmermann M, Seifert V. Near-infrared indocyanine green video angiography: a new method for intraoperative assessment of vascular flow. Neurosurgery 2003; 52 (01) 132-139 ; discussion 139
  • 30 Colditz MJ, Leyen Kv, Jeffree RL. Aminolevulinic acid (ALA)-protoporphyrin IX fluorescence guided tumour resection. Part 2: theoretical, biochemical and practical aspects. J Clin Neurosci 2012; 19 (12) 1611-1616
  • 31 Stummer W, Stepp H, Möller G, Ehrhardt A, Leonhard M, Reulen HJ. Technical principles for protoporphyrin-IX-fluorescence guided microsurgical resection of malignant glioma tissue. Acta Neurochir (Wien) 1998; 140 (10) 995-1000
  • 32 Stummer W, Stocker S, Novotny A. , et al. In vitro and in vivo porphyrin accumulation by C6 glioma cells after exposure to 5-aminolevulinic acid. J Photochem Photobiol B 1998; 45 (2–3): 160-169
  • 33 Kaneko S, Suero Molina E, Ewelt C, Warneke N, Stummer W. Fluorescence-based measurement of real-time kinetics of protoporphyrin IX after 5-Aminolevulinic acid administration in human in situ malignant gliomas. Neurosurgery. In press
  • 34 Stummer W, Pichlmeier U, Meinel T, Wiestler OD, Zanella F, Reulen HJ. ; ALA-Glioma Study Group. Fluorescence-guided surgery with 5-aminolevulinic acid for resection of malignant glioma: a randomised controlled multicentre phase III trial. Lancet Oncol 2006; 7 (05) 392-401
  • 35 Teixidor P, Arráez MA, Villalba G. , et al. Safety and efficacy of 5-aminolevulinic acid for high grade glioma in usual clinical practice: a prospective cohort study. PLoS One 2016; 11 (02) e0149244
  • 36 Eljamel S. 5-ALA fluorescence image guided resection of glioblastoma multiforme: a meta-analysis of the literature. Int J Mol Sci 2015; 16 (05) 10443-10456
  • 37 Stummer W, Novotny A, Stepp H, Goetz C, Bise K, Reulen HJ. Fluorescence-guided resection of glioblastoma multiforme by using 5-aminolevulinic acid-induced porphyrins: a prospective study in 52 consecutive patients. J Neurosurg 2000; 93 (06) 1003-1013
  • 38 Stummer W, Tonn JC, Goetz C. , et al. 5-Aminolevulinic acid-derived tumor fluorescence: the diagnostic accuracy of visible fluorescence qualities as corroborated by spectrometry and histology and postoperative imaging. Neurosurgery 2014; 74 (03) 310-319 ; discussion 319–320
  • 39 Díez Valle R, Tejada Solis S, Idoate Gastearena MA, García de Eulate R, Domínguez Echávarri P, Aristu Mendiroz J. Surgery guided by 5-aminolevulinic fluorescence in glioblastoma: volumetric analysis of extent of resection in single-center experience. J Neurooncol 2011; 102 (01) 105-113
  • 40 Nabavi A, Thurm H, Zountsas B. , et al; 5-ALA Recurrent Glioma Study Group. Five-aminolevulinic acid for fluorescence-guided resection of recurrent malignant gliomas: a phase ii study. Neurosurgery 2009; 65 (06) 1070-1076 ; discussion 1076–1077
  • 41 Lau D, Hervey-Jumper SL, Chang S. , et al. A prospective Phase II clinical trial of 5-aminolevulinic acid to assess the correlation of intraoperative fluorescence intensity and degree of histologic cellularity during resection of high-grade gliomas. J Neurosurg 2016; 124 (05) 1300-1309
  • 42 Schucht P, Beck J, Abu-Isa J. , et al. Gross total resection rates in contemporary glioblastoma surgery: results of an institutional protocol combining 5-aminolevulinic acid intraoperative fluorescence imaging and brain mapping. Neurosurgery 2012; 71 (05) 927-935 ; discussion 935–936
  • 43 Della Puppa A, De Pellegrin S, d'Avella E. , et al. 5-aminolevulinic acid (5-ALA) fluorescence guided surgery of high-grade gliomas in eloquent areas assisted by functional mapping. Our experience and review of the literature. Acta Neurochir (Wien) 2013; 155 (06) 965-972 ; discussion 972
  • 44 Schucht P, Seidel K, Beck J. , et al. Intraoperative monopolar mapping during 5-ALA-guided resections of glioblastomas adjacent to motor eloquent areas: evaluation of resection rates and neurological outcome. Neurosurg Focus 2014; 37 (06) E16
  • 45 Aldave G, Tejada S, Pay E. , et al. Prognostic value of residual fluorescent tissue in glioblastoma patients after gross total resection in 5-aminolevulinic Acid-guided surgery. Neurosurgery 2013; 72 (06) 915-920 ; discussion 920–921
  • 46 Schucht P, Knittel S, Slotboom J. , et al. 5-ALA complete resections go beyond MR contrast enhancement: shift corrected volumetric analysis of the extent of resection in surgery for glioblastoma. Acta Neurochir (Wien) 2014; 156 (02) 305-312 ; discussion 312
  • 47 Coburger J, Hagel V, Wirtz CR, König R. Surgery for glioblastoma: impact of the combined use of 5-aminolevulinic acid and intraoperative MRI on extent of resection and survival. PLoS One 2015; 10 (06) e0131872
  • 48 Coburger J, Engelke J, Scheuerle A. , et al. Tumor detection with 5-aminolevulinic acid fluorescence and Gd-DTPA-enhanced intraoperative MRI at the border of contrast-enhancing lesions: a prospective study based on histopathological assessment. Neurosurg Focus 2014; 36 (02) E3
  • 49 Rabb MF, Burton TC, Schatz H, Yannuzzi LA. Fluorescein angiography of the fundus: a schematic approach to interpretation. Surv Ophthalmol 1978; 22 (06) 387-403
  • 50 Shinoda J, Yano H, Yoshimura S. , et al. Fluorescence-guided resection of glioblastoma multiforme by using high-dose fluorescein sodium. Technical note. J Neurosurg 2003; 99 (03) 597-603
  • 51 Diaz RJ, Dios RR, Hattab EM. , et al. Study of the biodistribution of fluorescein in glioma-infiltrated mouse brain and histopathological correlation of intraoperative findings in high-grade gliomas resected under fluorescein fluorescence guidance. J Neurosurg 2015; 122 (06) 1360-1369
  • 52 Hamamcıoğlu MK, Akçakaya MO, Göker B, Kasımcan MO, Kırış T. The use of the YELLOW 560 nm surgical microscope filter for sodium fluorescein-guided resection of brain tumors: our preliminary results in a series of 28 patients. Clin Neurol Neurosurg 2016; 143: 39-45
  • 53 Schebesch KM, Brawanski A, Hohenberger C, Hohne J. Fluorescein sodium-guided surgery of malignant brain tumors: history, current concepts, and future project. Turk Neurosurg 2016; 26 (02) 185-194
  • 54 Schebesch KM, Proescholdt M, Höhne J. , et al. Sodium fluorescein-guided resection under the YELLOW 560 nm surgical microscope filter in malignant brain tumor surgery—a feasibility study. Acta Neurochir (Wien) 2013; 155 (04) 693-699
  • 55 Kuroiwa T, Kajimoto Y, Ohta T. Development of a fluorescein operative microscope for use during malignant glioma surgery: a technical note and preliminary report. Surg Neurol 1998; 50 (01) 41-48 ; discussion 48–49
  • 56 Acerbi F, Broggi M, Eoli M. , et al. Is fluorescein-guided technique able to help in resection of high-grade gliomas?. Neurosurg Focus 2014; 36 (02) E5
  • 57 Chen B, Wang H, Ge P. , et al. Gross total resection of glioma with the intraoperative fluorescence-guidance of fluorescein sodium. Int J Med Sci 2012; 9 (08) 708-714
  • 58 Francaviglia N, Iacopino DG, Costantino G. , et al. Fluorescein for resection of high-grade gliomas: a safety study control in a single center and review of the literature. Surg Neurol Int 2017; 8: 145
  • 59 Koc K, Anik I, Cabuk B, Ceylan S. Fluorescein sodium-guided surgery in glioblastoma multiforme: a prospective evaluation. Br J Neurosurg 2008; 22 (01) 99-103
  • 60 Stummer W. Poor man's fluorescence?. Acta Neurochir (Wien) 2015; 157 (08) 1379-1381
  • 61 Schwake M, Stummer W, Suero Molina EJ, Wölfer J. Simultaneous fluorescein sodium and 5-ALA in fluorescence-guided glioma surgery. Acta Neurochir (Wien) 2015; 157 (05) 877-879
  • 62 Acerbi F, Broggi M, Schebesch KM. , et al. Fluorescein-guided surgery for resection of high-grade gliomas: a multicentric prospective phase II study (FLUOGLIO). Clin Cancer Res 2018; 24 (01) 52-61
  • 63 Suero Molina E, Wölfer J, Ewelt C, Ehrhardt A, Brokinkel B, Stummer W. Dual-labeling with 5-aminolevulinic acid and fluorescein for fluorescence-guided resection of high-grade gliomas: technical note. J Neurosurg 2018; 128 (02) 399-405
  • 64 Cherrick GR, Stein SW, Leevy CM, Davidson CS. Indocyanine green: observations on its physical properties, plasma decay, and hepatic extraction. J Clin Invest 1960; 39: 592-600
  • 65 Reinhart MB, Huntington CR, Blair LJ, Heniford BT, Augenstein VA. Indocyanine green: historical context, current applications, and future considerations. Surg Innov 2016; 23 (02) 166-175
  • 66 Hänggi D, Etminan N, Steiger HJ. The impact of microscope-integrated intraoperative near-infrared indocyanine green videoangiography on surgery of arteriovenous malformations and dural arteriovenous fistulae. Neurosurgery 2010; 67 (04) 1094-1103 ; discussion 1103–1104
  • 67 Raabe A, Nakaji P, Beck J. , et al. Prospective evaluation of surgical microscope-integrated intraoperative near-infrared indocyanine green videoangiography during aneurysm surgery. J Neurosurg 2005; 103 (06) 982-989
  • 68 Lee JY, Thawani JP, Pierce J. , et al. Intraoperative near-infrared optical imaging can localize gadolinium-enhancing gliomas during surgery. Neurosurgery 2016; 79 (06) 856-871
  • 69 Swanson KI, Clark PA, Zhang RR. , et al. Fluorescent cancer-selective alkylphosphocholine analogs for intraoperative glioma detection. Neurosurgery 2015; 76 (02) 115-123 ; discussion 123–124
  • 70 Valdés PA, Roberts DW, Lu FK, Golby A. Optical technologies for intraoperative neurosurgical guidance. Neurosurg Focus 2016; 40 (03) E8
  • 71 Hilgard P, Klenner T, Stekar J, Unger C. Alkylphosphocholines: a new class of membrane-active anticancer agents. Cancer Chemother Pharmacol 1993; 32 (02) 90-95
  • 72 Weichert JP, Clark PA, Kandela IK. , et al. Alkylphosphocholine analogs for broad-spectrum cancer imaging and therapy. Sci Transl Med 2014; 6 (240) 240ra75
  • 73 Kuo JS, Zhang RR, Pinchuk AN, Clark PA, Weichert JP. Creation of a dual-labeled cancer-targeting alkylphosphocholine analog for dual modality quantitative positron emission tomography and intraoperative tumor visualization. Neurosurgery 2016; 63 (01) 208
  • 74 Parrish-Novak J, Byrnes-Blake K, Lalayeva N. , et al. Nonclinical profile of BLZ-100, a tumor-targeting fluorescent imaging agent. Int J Toxicol 2017; 36 (02) 104-112
  • 75 Butte PV, Mamelak A, Parrish-Novak J. , et al. Near-infrared imaging of brain tumors using the Tumor Paint BLZ-100 to achieve near-complete resection of brain tumors. Neurosurg Focus 2014; 36 (02) E1
  • 76 Uematsu Y, Owai Y, Okita R, Tanaka Y, Itakura T. The usefulness and problem of intraoperative rapid diagnosis in surgical neuropathology. Brain Tumor Pathol 2007; 24 (02) 47-52
  • 77 Foersch S, Heimann A, Ayyad A. , et al. Confocal laser endomicroscopy for diagnosis and histomorphologic imaging of brain tumors in vivo. PLoS One 2012; 7 (07) e41760
  • 78 Hoffman A, Goetz M, Vieth M, Galle PR, Neurath MF, Kiesslich R. Confocal laser endomicroscopy: technical status and current indications. Endoscopy 2006; 38 (12) 1275-1283
  • 79 Belykh E, Martirosyan NL, Yagmurlu K. , et al. Intraoperative fluorescence imaging for personalized brain tumor resection: current state and future directions. Front Surg 2016; 3: 55
  • 80 Raman C. A new type of secondary radiation. Nature 1928; 121: 501-502
  • 81 Hollon T, Lewis S, Freudiger CW, Sunney Xie X, Orringer DA. Improving the accuracy of brain tumor surgery via Raman-based technology. Neurosurg Focus 2016; 40 (03) E9
  • 82 Krafft C, Neudert L, Simat T, Salzer R. Near infrared Raman spectra of human brain lipids. Spectrochim Acta A Mol Biomol Spectrosc 2005; 61 (07) 1529-1535
  • 83 Jermyn M, Mok K, Mercier J. , et al. Intraoperative brain cancer detection with Raman spectroscopy in humans. Sci Transl Med 2015; 7 (274) 274ra19
  • 84 Desroches J, Jermyn M, Pinto M. , et al. A new method using Raman spectroscopy for in vivo targeted brain cancer tissue biopsy. Sci Rep 2018; 8 (01) 1792
  • 85 Arbizu J, Tejada S, Marti-Climent JM. , et al. Quantitative volumetric analysis of gliomas with sequential MRI and 11C-methionine PET assessment: patterns of integration in therapy planning. Eur J Nucl Med Mol Imaging 2012; 39 (05) 771-781
  • 86 Jaber M, Wölfer J, Ewelt C. , et al. The value of 5-aminolevulinic acid in low-grade gliomas and high-grade gliomas lacking glioblastoma imaging features: an analysis based on fluorescence, magnetic resonance imaging, 18F-fluoroethyl tyrosine positron emission tomography, and tumor molecular factors. Neurosurgery 2016; 78 (03) 401-411 ; discussion 411
  • 87 Idoate MA, Díez Valle R, Echeveste J, Tejada S. Pathological characterization of the glioblastoma border as shown during surgery using 5-aminolevulinic acid-induced fluorescence. Neuropathology 2011; 31 (06) 575-582
  • 88 Roberts DW, Valdés PA, Harris BT. , et al. Coregistered fluorescence-enhanced tumor resection of malignant glioma: relationships between δ-aminolevulinic acid-induced protoporphyrin IX fluorescence, magnetic resonance imaging enhancement, and neuropathological parameters. Clinical article. J Neurosurg 2011; 114 (03) 595-603
  • 89 Valdés PA, Leblond F, Kim A. , et al. Quantitative fluorescence in intracranial tumor: implications for ALA-induced PpIX as an intraoperative biomarker. J Neurosurg 2011; 115 (01) 11-17
  • 90 Valdés PA, Jacobs V, Harris BT. , et al. Quantitative fluorescence using 5-aminolevulinic acid-induced protoporphyrin IX biomarker as a surgical adjunct in low-grade glioma surgery. J Neurosurg 2015; 123 (03) 771-780

Zoom Image
Fig. 1 5-aminolevulinic fluorescence and different fluorescence qualities. (a) Cavity with area of strong (red) and weak (pink) and no fluorescence. (b) Corresponding white light image. (Reproduced with permission from Stummer et al.[38])
Zoom Image
Fig. 2 Use of fluorescein (FL) for resection of glioblastoma. (a) After administration of FL under white light: no fluorescent effect. (b) After administration of FL under YELLOW 560 nm filter: visible fluorescent effect in the tumor. (Reproduced with permission from Schebesch et al.[54])