2024

CVE-2024-0162 13 Mar 2024
Dell PowerEdge Server BIOS and Dell Precision Rack BIOS contain an Improper SMM communication buffer verification vulnerability. A local low privileged attacker could potentially exploit this vulnerability leading to out-of-bound read/writes to SMRAM.
CVE-2024-22041 12 Mar 2024
A vulnerability has been identified in Cerberus PRO EN Engineering Tool (All versions), Cerberus PRO EN Fire Panel FC72x (All versions < IP8 SR4), Cerberus PRO EN X200 Cloud Distribution (All versions < V4.3.5618), Cerberus PRO EN X300 Cloud Distribution (All versions < V4.3.5617), Sinteso FS20 EN Engineering Tool (All versions), Sinteso FS20 EN Fire Panel FC20 (All versions < MP8 SR4), Sinteso FS20 EN X200 Cloud Distribution (All versions < V4.3.5618), Sinteso FS20 EN X300 Cloud Distribution (All versions < V4.3.5617), Sinteso Mobile (All versions). The network communication library in affected systems improperly handles memory buffers when parsing X.509 certificates. This could allow an unauthenticated remote attacker to crash the network service.
CVE-2024-24921 (v3: 7.8) 13 Feb 2024
A vulnerability has been identified in Simcenter Femap (All versions < V2401.0000). The affected application is vulnerable to memory corruption while parsing specially crafted Catia MODEL files. This could allow an attacker to execute code in the context of the current process. (ZDI-CAN-21712)
CVE-2024-24560 2 Feb 2024
Vyper is a Pythonic Smart Contract Language for the Ethereum Virtual Machine. When calls to external contracts are made, we write the input buffer starting at byte 28, and allocate the return buffer to start at byte 0 (overlapping with the input buffer). When checking RETURNDATASIZE for dynamic types, the size is compared only to the minimum allowed size for that type, and not to the returned value's length. As a result, malformed return data can cause the contract to mistake data from the input buffer for returndata. When the called contract returns invalid ABIv2 encoded data, the calling contract can read different invalid data (from the dirty buffer) than the called contract returned.
CVE-2024-0338 2 Feb 2024
A buffer overflow vulnerability has been found in XAMPP affecting version 8.2.4 and earlier. An attacker could execute arbitrary code through a long file debug argument that controls the Structured Exception Handler (SEH).
CVE-2024-24561 1 Feb 2024
Vyper is a pythonic Smart Contract Language for the ethereum virtual machine. In versions 0.3.10 and earlier, the bounds check for slices does not account for the ability for start + length to overflow when the values aren't literals. If a slice() function uses a non-literal argument for the start or length variable, this creates the ability for an attacker to overflow the bounds check. This issue can be used to do OOB access to storage, memory or calldata addresses. It can also be used to corrupt the length slot of the respective array.
CVE-2024-1112 31 Jan 2024
Heap-based buffer overflow vulnerability in Resource Hacker, developed by Angus Johnson, affecting version 3.6.0.92. This vulnerability could allow an attacker to execute arbitrary code via a long filename argument.
CVE-2024-0774 22 Jan 2024
A vulnerability was found in Any-Capture Any Sound Recorder 2.93. It has been declared as problematic. This vulnerability affects unknown code of the component Registration Handler. The manipulation of the argument User Name/Key Code leads to memory corruption. It is possible to launch the attack on the local host. The exploit has been disclosed to the public and may be used. VDB-251674 is the identifier assigned to this vulnerability.
CVE-2024-0772 22 Jan 2024
A vulnerability was found in Nsasoft ShareAlarmPro 2.1.4 and classified as problematic. Affected by this issue is some unknown functionality of the component Registration Handler. The manipulation of the argument Name/Key leads to memory corruption. Local access is required to approach this attack. The exploit has been disclosed to the public and may be used. The identifier of this vulnerability is VDB-251672. NOTE: The vendor was contacted early about this disclosure but did not respond in any way.
CVE-2024-0771 21 Jan 2024
A vulnerability has been found in Nsasoft Product Key Explorer 4.0.9 and classified as problematic. Affected by this vulnerability is an unknown functionality of the component Registration Handler. The manipulation of the argument Name/Key leads to memory corruption. An attack has to be approached locally. The exploit has been disclosed to the public and may be used. The associated identifier of this vulnerability is VDB-251671. NOTE: The vendor was contacted early about this disclosure but did not respond in any way.
CVE-2024-0645 17 Jan 2024
Buffer overflow vulnerability in Explorer++ affecting version 1.3.5.531. A local attacker could execute arbitrary code via a long filename argument by monitoring Structured Exception Handler (SEH) records.
CVE-2024-0429 11 Jan 2024
A denial service vulnerability has been found on  Hex Workshop affecting version 6.7, an attacker could send a command line file arguments and control the Structured Exception Handler (SEH) records resulting in a service shutdown.

2023

CVE-2023-32331 4 Mar 2024
IBM Connect:Express for UNIX 1.5.0 is vulnerable to a buffer overflow that could allow a remote attacker to cause a denial of service through its browser UI. IBM X-Force ID: 254979.
CVE-2023-4590 27 Nov 2023
Buffer overflow vulnerability in Frhed hex editor, affecting version 1.6.0. This vulnerability could allow an attacker to execute arbitrary code via a long filename argument through the Structured Exception Handler (SEH) registers.
CVE-2023-44023 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the ssid parameter in the form_fast_setting_wifi_set function.
CVE-2023-44022 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the speed_dir parameter in the formSetSpeedWan function.
CVE-2023-44021 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the formSetClientState function.
CVE-2023-44020 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the security parameter in the formWifiBasicSet function.
CVE-2023-44019 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the mac parameter in the GetParentControlInfo function.
CVE-2023-44018 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the domain parameter in the add_white_node function.
CVE-2023-44017 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the timeZone parameter in the fromSetSysTime function.
CVE-2023-44016 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the deviceId parameter in the addWifiMacFilter function.
CVE-2023-44015 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the schedEndTime parameter in the setSchedWifi function.
CVE-2023-44014 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain multiple stack overflows in the formSetMacFilterCfg function via the macFilterType and deviceList parameters.
CVE-2023-44013 (v3: 9.8) 27 Sep 2023
Tenda AC10U v1.0 US_AC10UV1.0RTL_V15.03.06.49_multi_TDE01 was discovered to contain a stack overflow via the list parameter in the fromSetIpMacBind function.
CVE-2023-41846 (v3: 7.8) 12 Sep 2023
A vulnerability has been identified in Tecnomatix Plant Simulation V2201 (All versions < V2201.0008), Tecnomatix Plant Simulation V2302 (All versions < V2302.0002). The affected application is vulnerable to memory corruption while parsing specially crafted SPP files. This could allow an attacker to execute code in the context of the current process.
CVE-2023-3953 9 Aug 2023
A CWE-119: Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability exists that could cause memory corruption when an authenticated user opens a tampered log file from GP-Pro EX.
CVE-2023-35871 11 Jul 2023
The SAP Web Dispatcher - versions WEBDISP 7.53, WEBDISP 7.54, WEBDISP 7.77, WEBDISP 7.85, WEBDISP 7.89, WEBDISP 7.91, WEBDISP 7.92, WEBDISP 7.93, KERNEL 7.53, KERNEL 7.54 KERNEL 7.77, KERNEL 7.85, KERNEL 7.89, KERNEL 7.91, KERNEL 7.92, KERNEL 7.93, KRNL64UC 7.53, HDB 2.00, XS_ADVANCED_RUNTIME 1.00, SAP_EXTENDED_APP_SERVICES 1, has a vulnerability that can be exploited by an unauthenticated attacker to cause memory corruption through logical errors in memory management this may leads to information disclosure or system crashes, which can have low impact on confidentiality and high impact on the integrity and availability of the system.
CVE-2023-32203 6 Jun 2023
The affected application lacks proper validation of user-supplied data when parsing project files (e.g., HMI). This could lead to an out-of-bounds write at CScape_EnvisionRV+0x2e374b. An attacker could leverage this vulnerability to execute arbitrary code in the context of the current process.
CVE-2023-31278 6 Jun 2023
The affected application lacks proper validation of user-supplied data when parsing project files (e.g., HMI). This could lead to an out-of-bounds read. An attacker could leverage this vulnerability to potentially execute arbitrary code in the context of the current process.
CVE-2023-32539 6 Jun 2023
The affected application lacks proper validation of user-supplied data when parsing project files (e.g., HMI). This could lead to an out-of-bounds write at CScape_EnvisionRV+0x2e3c04. An attacker could leverage this vulnerability to potentially execute arbitrary code in the context of the current process.
CVE-2023-2970 30 May 2023
A vulnerability classified as problematic was found in MindSpore 2.0.0-alpha/2.0.0-rc1. This vulnerability affects the function JsonHelper::UpdateArray of the file mindspore/ccsrc/minddata/dataset/util/json_helper.cc. The manipulation leads to memory corruption. The name of the patch is 30f4729ea2c01e1ed437ba92a81e2fc098d608a9. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-230176.
CVE-2023-2873 24 May 2023
A vulnerability classified as critical was found in Twister Antivirus 8. This vulnerability affects the function 0x804f2143/0x804f217f/0x804f214b/0x80800043 in the library filppd.sys of the component IoControlCode Handler. The manipulation leads to memory corruption. Local access is required to approach this attack. The exploit has been disclosed to the public and may be used. The identifier of this vulnerability is VDB-229852. NOTE: The vendor was contacted early about this disclosure but did not respond in any way.
CVE-2023-30986 (v3: 7.8) 9 May 2023
A vulnerability has been identified in Solid Edge SE2023 (All versions < VX.223.0 Update 3), Solid Edge SE2023 (All versions < VX.223.0 Update 2). Affected applications contain a memory corruption vulnerability while parsing specially crafted STP files. This could allow an attacker to execute code in the context of the current process. (ZDI-CAN-19561)
CVE-2023-26285 5 May 2023
IBM MQ 9.2 CD, 9.2 LTS, 9.3 CD, and 9.3 LTS could allow a remote attacker to cause a denial of service due to an error processing invalid data. IBM X-Force ID: 248418.
CVE-2023-27286 2 Apr 2023
IBM Aspera Cargo 4.2.5 and IBM Aspera Connect 4.2.5 are vulnerable to a buffer overflow, caused by improper bounds checking. An attacker could overflow a buffer and execute arbitrary code on the system. IBM X-Force ID: 248616.
CVE-2023-27284 2 Apr 2023
IBM Aspera Cargo 4.2.5 and IBM Aspera Connect 4.2.5 are vulnerable to a buffer overflow, caused by improper bounds checking. An attacker could overflow a buffer and execute arbitrary code on the system. IBM X-Force ID: 248616.
CVE-2023-28638 27 Mar 2023
Snappier is a high performance C# implementation of the Snappy compression algorithm. This is a buffer overrun vulnerability that can affect any user of Snappier 1.1.0. In this release, much of the code was rewritten to use byte references rather than pointers to pinned buffers. This change generally improves performance and reduces workload on the garbage collector. However, when the garbage collector performs compaction and rearranges memory, it must update any byte references on the stack to refer to the updated location. The .NET garbage collector can only update these byte references if they still point within the buffer or to a point one byte past the end of the buffer. If they point outside this area, the buffer itself may be moved while the byte reference stays the same. There are several places in 1.1.0 where byte references very briefly point outside the valid areas of buffers. These are at locations in the code being used for buffer range checks. While the invalid references are never dereferenced directly, if a GC compaction were to occur during the brief window when they are on the stack then it could invalidate the buffer range check and allow other operations to overrun the buffer. This should be very difficult for an attacker to trigger intentionally. It would require a repetitive bulk attack with the hope that a GC compaction would occur at precisely the right moment during one of the requests. However, one of the range checks with this problem is a check based on input data in the decompression buffer, meaning malformed input data could be used to increase the chance of success. Note that any resulting buffer overrun is likely to cause access to protected memory, which will then cause an exception and the process to be terminated. Therefore, the most likely result of an attack is a denial of service. This issue has been patched in release 1.1.1. Users are advised to upgrade. Users unable to upgrade may pin buffers to a fixed location before using them for compression or decompression to mitigate some, but not all, of these cases. At least one temporary decompression buffer is internal to the library and never pinned.
CVE-2023-27403 (v3: 7.8) 14 Mar 2023
A vulnerability has been identified in Tecnomatix Plant Simulation (All versions < V2201.0006). The affected application contains a memory corruption vulnerability while parsing specially crafted SPP files. This could allow an attacker to execute code in the context of the current process. (ZDI-CAN-20303, ZDI-CAN-20348)
CVE-2023-24564 14 Feb 2023
A vulnerability has been identified in Solid Edge SE2022 (All versions < V2210Update12), Solid Edge SE2022 (All versions), Solid Edge SE2023 (All versions < V2023Update2). The affected application contains a memory corruption vulnerability while parsing specially crafted DWG files. This could allow an attacker to execute code in the context of the current process. (ZDI-CAN-19069)
CVE-2023-0251 8 Feb 2023
Delta Electronics DIAScreen versions 1.2.1.23 and prior are vulnerable to a buffer overflow through improper restrictions of operations within memory, which could allow an attacker to remotely execute arbitrary code.
CVE-2023-0637 2 Feb 2023
A vulnerability, which was classified as critical, was found in TRENDnet TEW-811DRU 1.0.10.0. This affects an unknown part of the file wan.asp of the component Web Management Interface. The manipulation leads to memory corruption. It is possible to initiate the attack remotely. The exploit has been disclosed to the public and may be used. The identifier VDB-220017 was assigned to this vulnerability.
CVE-2023-0618 1 Feb 2023
A vulnerability was found in TRENDnet TEW-652BRP 3.04B01. It has been declared as critical. This vulnerability affects unknown code of the file cfg_op.ccp of the component Web Service. The manipulation leads to memory corruption. The attack can be initiated remotely. The exploit has been disclosed to the public and may be used. VDB-219958 is the identifier assigned to this vulnerability.
CVE-2023-0613 1 Feb 2023
A vulnerability has been found in TRENDnet TEW-811DRU 1.0.10.0 and classified as critical. Affected by this vulnerability is an unknown functionality of the file /wireless/security.asp of the component httpd. The manipulation leads to memory corruption. The attack can be launched remotely. The exploit has been disclosed to the public and may be used. The identifier VDB-219937 was assigned to this vulnerability.

2022

CVE-2022-4046 (v3: 8.8) 3 Aug 2023
In CODESYS Control in multiple versions a improper restriction of operations within the bounds of a memory buffer allow an remote attacker with user privileges to gain full access of the device.
CVE-2022-46295 21 Jul 2023
Multiple out-of-bounds write vulnerabilities exist in the translationVectors parsing functionality in multiple supported formats of Open Babel 3.1.1 and master commit 530dbfa3. A specially-crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.This vulnerability affects the Gaussian file format
CVE-2022-46294 21 Jul 2023
Multiple out-of-bounds write vulnerabilities exist in the translationVectors parsing functionality in multiple supported formats of Open Babel 3.1.1 and master commit 530dbfa3. A specially-crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.This vulnerability affects the MOPAC Cartesian file format
CVE-2022-46293 21 Jul 2023
Multiple out-of-bounds write vulnerabilities exist in the translationVectors parsing functionality in multiple supported formats of Open Babel 3.1.1 and master commit 530dbfa3. A specially-crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.This vulnerability affects the MOPAC file format, inside the Final Point and Derivatives section
CVE-2022-46292 21 Jul 2023
Multiple out-of-bounds write vulnerabilities exist in the translationVectors parsing functionality in multiple supported formats of Open Babel 3.1.1 and master commit 530dbfa3. A specially-crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.This vulnerability affects the MOPAC file format, inside the Unit Cell Translation section
CVE-2022-46291 21 Jul 2023
Multiple out-of-bounds write vulnerabilities exist in the translationVectors parsing functionality in multiple supported formats of Open Babel 3.1.1 and master commit 530dbfa3. A specially-crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.This vulnerability affects the MSI file format
CVE-2022-43607 21 Jul 2023
An out-of-bounds write vulnerability exists in the MOL2 format attribute and value functionality of Open Babel 3.1.1 and master commit 530dbfa3. A specially crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.
CVE-2022-43467 21 Jul 2023
An out-of-bounds write vulnerability exists in the PQS format coord_file functionality of Open Babel 3.1.1 and master commit 530dbfa3. A specially crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.
CVE-2022-37331 21 Jul 2023
An out-of-bounds write vulnerability exists in the Gaussian format orientation functionality of Open Babel 3.1.1 and master commit 530dbfa3. A specially crafted malformed file can lead to arbitrary code execution. An attacker can provide a malicious file to trigger this vulnerability.
CVE-2022-47393 (v3: 6.5) 15 May 2023
An authenticated, remote attacker may use a Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability in multiple versions of multiple CODESYS products to force a denial-of-service situation.
CVE-2022-37381 29 Mar 2023
This vulnerability allows remote attackers to execute arbitrary code on affected installations of Foxit PDF Reader 11.2.1.53537. User interaction is required to exploit this vulnerability in that the target must visit a malicious page or open a malicious file. The specific flaw exists within the AFSpecial_KeystrokeEx method. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to execute code in the context of the current process. Was ZDI-CAN-17110.
CVE-2022-47977 14 Feb 2023
A vulnerability has been identified in JT Open (All versions < V11.2.3.0), JT Utilities (All versions < V13.2.3.0). The affected application contains a memory corruption vulnerability while parsing specially crafted JT files. This could allow an attacker to execute code in the context of the current process.
CVE-2022-32512 30 Jan 2023
A CWE-119: Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability exists that could cause remote code execution when a command which exploits this vulnerability is utilized. Affected Products: CanBRASS (Versions prior to V7.5.1)
CVE-2022-3161 13 Jan 2023
The APDFL.dll contains a memory corruption vulnerability while parsing specially crafted PDF files. This could allow an attacker to execute code in the context of the current process.
CVE-2022-3628 12 Jan 2023
A buffer overflow flaw was found in the Linux kernel Broadcom Full MAC Wi-Fi driver. This issue occurs when a user connects to a malicious USB device. This can allow a local user to crash the system or escalate their privileges.
CVE-2022-38105 10 Jan 2023
An information disclosure vulnerability exists in the cm_processREQ_NC opcode of Asus RT-AX82U 3.0.0.4.386_49674-ge182230 router's configuration service. A specially-crafted network packets can lead to a disclosure of sensitive information. An attacker can send a network request to trigger this vulnerability.
CVE-2022-47967 10 Jan 2023
A vulnerability has been identified in Solid Edge (All versions < V2023 MP1). The DOCMGMT.DLL contains a memory corruption vulnerability that could be triggered while parsing files in different file formats such as PAR, ASM, DFT. This could allow an attacker to execute code in the context of the current process.
CVE-2022-47935 10 Jan 2023
A vulnerability has been identified in JT Open (All versions < V11.1.1.0), JT Utilities (All versions < V13.1.1.0), Solid Edge (All versions < V2023). The Jt1001.dll contains a memory corruption vulnerability while parsing specially crafted JT files. An attacker could leverage this vulnerability to execute code in the context of the current process. (ZDI-CAN-19078)
CVE-2022-3715 5 Jan 2023
A flaw was found in the bash package, where a heap-buffer overflow can occur in valid parameter_transform. This issue may lead to memory problems.
CVE-2022-4639 21 Dec 2022
A vulnerability, which was classified as critical, has been found in sslh. This issue affects the function hexdump of the file probe.c of the component Packet Dumping Handler. The manipulation of the argument msg_info leads to format string. The attack may be initiated remotely. The name of the patch is b19f8a6046b080e4c2e28354a58556bb26040c6f. It is recommended to apply a patch to fix this issue. The identifier VDB-216497 was assigned to this vulnerability.
CVE-2022-4603 18 Dec 2022
** DISPUTED ** A vulnerability classified as problematic has been found in ppp. Affected is the function dumpppp of the file pppdump/pppdump.c of the component pppdump. The manipulation of the argument spkt.buf/rpkt.buf leads to improper validation of array index. The real existence of this vulnerability is still doubted at the moment. The name of the patch is a75fb7b198eed50d769c80c36629f38346882cbf. It is recommended to apply a patch to fix this issue. VDB-216198 is the identifier assigned to this vulnerability. NOTE: pppdump is not used in normal process of setting up a PPP connection, is not installed setuid-root, and is not invoked automatically in any scenario.
CVE-2022-4584 17 Dec 2022
A vulnerability was found in Axiomatic Bento4. It has been rated as critical. Affected by this issue is some unknown functionality of the component mp42aac. The manipulation leads to heap-based buffer overflow. The attack may be launched remotely. The exploit has been disclosed to the public and may be used. VDB-216170 is the identifier assigned to this vulnerability.
CVE-2022-2947 13 Dec 2022
Altair HyperView Player versions 2021.1.0.27 and prior perform operations on a memory buffer but can read from or write to a memory location outside of the intended boundary of the buffer. This hits initially as a read access violation, leading to a memory corruption situation.
CVE-2022-23523 13 Dec 2022
In versions prior to 0.8.1, the linux-loader crate uses the offsets and sizes provided in the ELF headers to determine the offsets to read from. If those offsets point beyond the end of the file this could lead to Virtual Machine Monitors using the `linux-loader` crate entering an infinite loop if the ELF header of the kernel they are loading was modified in a malicious manner. This issue has been addressed in 0.8.1. The issue can be mitigated by ensuring that only trusted kernel images are loaded or by verifying that the headers do not point beyond the end of the file.
CVE-2022-3461 (v3: 7.8) 15 Nov 2022
In PHOENIX CONTACT Automationworx Software Suite up to version 1.89 manipulated PC Worx or Config+ files could lead to a heap buffer overflow and a read access violation. Availability, integrity, or confidentiality of an application programming workstation might be compromised by attacks using these vulnerabilities.
CVE-2022-3974 13 Nov 2022
A vulnerability classified as critical was found in Axiomatic Bento4. Affected by this vulnerability is the function AP4_StdcFileByteStream::ReadPartial of the file Ap4StdCFileByteStream.cpp of the component mp4info. The manipulation leads to heap-based buffer overflow. The attack can be launched remotely. The exploit has been disclosed to the public and may be used. The identifier VDB-213553 was assigned to this vulnerability.
CVE-2022-3965 13 Nov 2022
A vulnerability classified as problematic was found in ffmpeg. This vulnerability affects the function smc_encode_stream of the file libavcodec/smcenc.c of the component QuickTime Graphics Video Encoder. The manipulation of the argument y_size leads to out-of-bounds read. The attack can be initiated remotely. The name of the patch is 13c13109759090b7f7182480d075e13b36ed8edd. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-213544.
CVE-2022-3964 13 Nov 2022
A vulnerability classified as problematic has been found in ffmpeg. This affects an unknown part of the file libavcodec/rpzaenc.c of the component QuickTime RPZA Video Encoder. The manipulation of the argument y_size leads to out-of-bounds read. It is possible to initiate the attack remotely. The name of the patch is 92f9b28ed84a77138105475beba16c146bdaf984. It is recommended to apply a patch to fix this issue. The associated identifier of this vulnerability is VDB-213543.
CVE-2022-3785 31 Oct 2022
A vulnerability, which was classified as critical, has been found in Axiomatic Bento4. Affected by this issue is the function AP4_DataBuffer::SetDataSize of the component Avcinfo. The manipulation leads to heap-based buffer overflow. The attack may be launched remotely. The exploit has been disclosed to the public and may be used. The identifier of this vulnerability is VDB-212564.
CVE-2022-3784 31 Oct 2022
A vulnerability classified as critical was found in Axiomatic Bento4 5e7bb34. Affected by this vulnerability is the function AP4_Mp4AudioDsiParser::ReadBits of the file Ap4Mp4AudioInfo.cpp of the component mp4hls. The manipulation leads to heap-based buffer overflow. The attack can be launched remotely. The exploit has been disclosed to the public and may be used. The associated identifier of this vulnerability is VDB-212563.
CVE-2022-3757 29 Oct 2022
A vulnerability was found in Exiv2. It has been declared as critical. Affected by this vulnerability is the function QuickTimeVideo::decodeBlock of the file quicktimevideo.cpp of the component QuickTime Video Handler. The manipulation leads to buffer overflow. The attack can be launched remotely. The name of the patch is d3651fdbd352cbaf259f89abf7557da343339378. It is recommended to apply a patch to fix this issue. The identifier VDB-212497 was assigned to this vulnerability.
CVE-2022-3719 27 Oct 2022
A vulnerability has been found in Exiv2 and classified as critical. This vulnerability affects the function QuickTimeVideo::userDataDecoder of the file quicktimevideo.cpp of the component QuickTime Video Handler. The manipulation leads to heap-based buffer overflow. The attack can be initiated remotely. The name of the patch is a38e124076138e529774d5ec9890d0731058115a. It is recommended to apply a patch to fix this issue. VDB-212350 is the identifier assigned to this vulnerability.
CVE-2022-3717 27 Oct 2022
A vulnerability, which was classified as critical, has been found in Exiv2. Affected by this issue is the function BmffImage::boxHandler of the file bmffimage.cpp. The manipulation leads to memory corruption. The attack may be launched remotely. The name of the patch is a58e52ed702d3bc7b8bab7ec1d70a4849eebece3. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-212348.
CVE-2022-3705 26 Oct 2022
A vulnerability was found in vim and classified as problematic. Affected by this issue is the function qf_update_buffer of the file quickfix.c of the component autocmd Handler. The manipulation leads to use after free. The attack may be launched remotely. Upgrading to version 9.0.0805 is able to address this issue. The name of the patch is d0fab10ed2a86698937e3c3fed2f10bd9bb5e731. It is recommended to upgrade the affected component. The identifier of this vulnerability is VDB-212324.
CVE-2022-3670 26 Oct 2022
A vulnerability was found in Axiomatic Bento4. It has been classified as critical. Affected is the function WriteSample of the component mp42hevc. The manipulation leads to heap-based buffer overflow. It is possible to launch the attack remotely. The exploit has been disclosed to the public and may be used. VDB-212010 is the identifier assigned to this vulnerability.
CVE-2022-3666 26 Oct 2022
A vulnerability, which was classified as critical, has been found in Axiomatic Bento4. Affected by this issue is the function AP4_LinearReader::Advance of the file Ap4LinearReader.cpp of the component mp42ts. The manipulation leads to use after free. The attack may be launched remotely. The exploit has been disclosed to the public and may be used. VDB-212006 is the identifier assigned to this vulnerability.
CVE-2022-3665 26 Oct 2022
A vulnerability classified as critical was found in Axiomatic Bento4. Affected by this vulnerability is an unknown functionality of the file AvcInfo.cpp of the component avcinfo. The manipulation leads to heap-based buffer overflow. The attack can be launched remotely. The exploit has been disclosed to the public and may be used. The identifier VDB-212005 was assigned to this vulnerability.
CVE-2022-3664 26 Oct 2022
A vulnerability classified as critical has been found in Axiomatic Bento4. Affected is the function AP4_BitStream::WriteBytes of the file Ap4BitStream.cpp of the component avcinfo. The manipulation leads to heap-based buffer overflow. It is possible to launch the attack remotely. The exploit has been disclosed to the public and may be used. The identifier of this vulnerability is VDB-212004.
CVE-2022-3662 26 Oct 2022
A vulnerability was found in Axiomatic Bento4. It has been declared as critical. This vulnerability affects the function GetOffset of the file Ap4Sample.h of the component mp42hls. The manipulation leads to use after free. The attack can be initiated remotely. The exploit has been disclosed to the public and may be used. VDB-212002 is the identifier assigned to this vulnerability.
CVE-2022-3649 21 Oct 2022
A vulnerability was found in Linux Kernel. It has been classified as problematic. Affected is the function nilfs_new_inode of the file fs/nilfs2/inode.c of the component BPF. The manipulation leads to use after free. It is possible to launch the attack remotely. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-211992.
CVE-2022-3640 21 Oct 2022
A vulnerability, which was classified as critical, was found in Linux Kernel. Affected is the function l2cap_conn_del of the file net/bluetooth/l2cap_core.c of the component Bluetooth. The manipulation leads to use after free. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-211944.
CVE-2022-3636 21 Oct 2022
A vulnerability, which was classified as critical, was found in Linux Kernel. This affects the function __mtk_ppe_check_skb of the file drivers/net/ethernet/mediatek/mtk_ppe.c of the component Ethernet Handler. The manipulation leads to use after free. It is recommended to apply a patch to fix this issue. The associated identifier of this vulnerability is VDB-211935.
CVE-2022-3635 21 Oct 2022
A vulnerability, which was classified as critical, has been found in Linux Kernel. Affected by this issue is the function tst_timer of the file drivers/atm/idt77252.c of the component IPsec. The manipulation leads to use after free. It is recommended to apply a patch to fix this issue. VDB-211934 is the identifier assigned to this vulnerability.
CVE-2022-3625 21 Oct 2022
A vulnerability was found in Linux Kernel. It has been classified as critical. This affects the function devlink_param_set/devlink_param_get of the file net/core/devlink.c of the component IPsec. The manipulation leads to use after free. It is recommended to apply a patch to fix this issue. The identifier VDB-211929 was assigned to this vulnerability.
CVE-2022-3620 20 Oct 2022
A vulnerability was found in Exim and classified as problematic. This issue affects the function dmarc_dns_lookup of the file dmarc.c of the component DMARC Handler. The manipulation leads to use after free. The attack may be initiated remotely. The name of the patch is 12fb3842f81bcbd4a4519d5728f2d7e0e3ca1445. It is recommended to apply a patch to fix this issue. The associated identifier of this vulnerability is VDB-211919.
CVE-2022-27625 20 Oct 2022
A vulnerability regarding improper restriction of operations within the bounds of a memory buffer is found in the message processing functionality of Out-of-Band (OOB) Management. This allows remote attackers to execute arbitrary commands via unspecified vectors. The following models with Synology DiskStation Manager (DSM) versions before 7.1.1-42962-2 may be affected: DS3622xs+, FS3410, and HD6500.
CVE-2022-27624 20 Oct 2022
A vulnerability regarding improper restriction of operations within the bounds of a memory buffer is found in the packet decryption functionality of Out-of-Band (OOB) Management. This allows remote attackers to execute arbitrary commands via unspecified vectors. The following models with Synology DiskStation Manager (DSM) versions before 7.1.1-42962-2 may be affected: DS3622xs+, FS3410, and HD6500.
CVE-2022-3595 18 Oct 2022
A vulnerability was found in Linux Kernel. It has been rated as problematic. Affected by this issue is the function sess_free_buffer of the file fs/cifs/sess.c of the component CIFS Handler. The manipulation leads to double free. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-211364.
CVE-2022-3565 17 Oct 2022
A vulnerability, which was classified as critical, has been found in Linux Kernel. Affected by this issue is the function del_timer of the file drivers/isdn/mISDN/l1oip_core.c of the component Bluetooth. The manipulation leads to use after free. It is recommended to apply a patch to fix this issue. The identifier of this vulnerability is VDB-211088.
CVE-2022-3559 17 Oct 2022
A vulnerability was found in Exim and classified as critical. This issue affects some unknown processing of the component Regex Handler. The manipulation leads to use after free. The name of the patch is 4e9ed49f8f12eb331b29bd5b6dc3693c520fddc2. It is recommended to apply a patch to fix this issue. The identifier VDB-211073 was assigned to this vulnerability.

2021

CVE-2021-46598 18 Feb 2022
This vulnerability allows remote attackers to execute arbitrary code on affected installations of Bentley MicroStation CONNECT 10.16.0.80. User interaction is required to exploit this vulnerability in that the target must visit a malicious page or open a malicious file. The specific flaw exists within the parsing of JT files. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to execute code in the context of the current process. Was ZDI-CAN-15392.
CVE-2021-46157 (v3: 7.8) 9 Feb 2022
A vulnerability has been identified in Simcenter Femap V2020.2 (All versions), Simcenter Femap V2021.1 (All versions). Affected application contains a memory corruption vulnerability while parsing NEU files. This could allow an attacker to execute code in the context of the current process. (ZDI-CAN-14757)
CVE-2021-46153 (v3: 7.8) 9 Feb 2022
A vulnerability has been identified in Simcenter Femap V2020.2 (All versions), Simcenter Femap V2021.1 (All versions). Affected application contains a memory corruption vulnerability while parsing NEU files. This could allow an attacker to execute code in the context of the current process. (ZDI-CAN-14645, ZDI-CAN-15305, ZDI-CAN-15589, ZDI-CAN-15599)
CVE-2021-4011 17 Dec 2021
A flaw was found in xorg-x11-server in versions before 21.1.2 and before 1.20.14. An out-of-bounds access can occur in the SwapCreateRegister function. The highest threat from this vulnerability is to data confidentiality and integrity as well as system availability.
CVE-2021-4010 17 Dec 2021
A flaw was found in xorg-x11-server in versions before 21.1.2 and before 1.20.14. An out-of-bounds access can occur in the SProcScreenSaverSuspend function. The highest threat from this vulnerability is to data confidentiality and integrity as well as system availability.
CVE-2021-4009 17 Dec 2021
A flaw was found in xorg-x11-server in versions before 21.1.2 and before 1.20.14. An out-of-bounds access can occur in the SProcXFixesCreatePointerBarrier function. The highest threat from this vulnerability is to data confidentiality and integrity as well as system availability.
CVE-2021-4008 17 Dec 2021
A flaw was found in xorg-x11-server in versions before 21.1.2 and before 1.20.14. An out-of-bounds access can occur in the SProcRenderCompositeGlyphs function. The highest threat from this vulnerability is to data confidentiality and integrity as well as system availability.
CVE-2021-42267 (v3: 7.8) 18 Nov 2021
Adobe Animate version 21.0.9 (and earlier) is affected by a memory corruption vulnerability due to insecure handling of a malicious FLA file, potentially resulting in arbitrary code execution in the context of the current user. User interaction is required to exploit this vulnerability.
CVE-2021-42721 (v3: 7.8) 16 Nov 2021
Adobe Media Encoder version 15.4 (and earlier) are affected by a memory corruption vulnerability. An unauthenticated attacker could leverage this vulnerability to achieve arbitrary code execution in the context of the current user. Exploitation of this issue requires user interaction in that a victim must open a malicious M4A file.
CVE-2021-41289 (v3: 6.3) 15 Nov 2021
ASUS P453UJ contains the Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability. With a general user’s permission, local attackers can modify the BIOS by replacing or filling in the content of the designated Memory DataBuffer, which causing a failure of integrity verification and further resulting in a failure to boot.
CVE-2021-31883 9 Nov 2021
A vulnerability has been identified in APOGEE MBC (PPC) (BACnet) (All versions), APOGEE MBC (PPC) (P2 Ethernet) (All versions), APOGEE MEC (PPC) (BACnet) (All versions), APOGEE MEC (PPC) (P2 Ethernet) (All versions), APOGEE PXC Compact (BACnet) (All versions), APOGEE PXC Compact (P2 Ethernet) (All versions), APOGEE PXC Modular (BACnet) (All versions), APOGEE PXC Modular (P2 Ethernet) (All versions), Capital VSTAR (All versions), Nucleus NET (All versions), Nucleus ReadyStart V3 (All versions < V2017.02.4), Nucleus Source Code (All versions), TALON TC Compact (BACnet) (All versions), TALON TC Modular (BACnet) (All versions). When processing a DHCP ACK message, the DHCP client application does not validate the length of the Vendor option(s), leading to Denial-of-Service conditions. (FSMD-2021-0013)
CVE-2021-31882 9 Nov 2021
A vulnerability has been identified in APOGEE MBC (PPC) (BACnet) (All versions), APOGEE MBC (PPC) (P2 Ethernet) (All versions), APOGEE MEC (PPC) (BACnet) (All versions), APOGEE MEC (PPC) (P2 Ethernet) (All versions), APOGEE PXC Compact (BACnet) (All versions), APOGEE PXC Compact (P2 Ethernet) (All versions), APOGEE PXC Modular (BACnet) (All versions), APOGEE PXC Modular (P2 Ethernet) (All versions), Capital VSTAR (All versions), Nucleus NET (All versions), Nucleus ReadyStart V3 (All versions < V2017.02.4), Nucleus Source Code (All versions), TALON TC Compact (BACnet) (All versions), TALON TC Modular (BACnet) (All versions). The DHCP client application does not validate the length of the Domain Name Server IP option(s) (0x06) when processing DHCP ACK packets. This may lead to Denial-of-Service conditions. (FSMD-2021-0011)
CVE-2021-34859 25 Oct 2021
This vulnerability allows remote attackers to execute arbitrary code on affected installations of TeamViewer 15.16.8.0. User interaction is required to exploit this vulnerability in that the target must visit a malicious page or open a malicious file. The specific flaw exists within the parsing of TVS files. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to execute code in the context of the current process. Was ZDI-CAN-13697.
CVE-2021-34856 25 Oct 2021
This vulnerability allows local attackers to escalate privileges on affected installations of Parallels Desktop 16.1.3 (49160). An attacker must first obtain the ability to execute high-privileged code on the target guest system in order to exploit this vulnerability. The specific flaw exists within the virtio-gpu virtual device. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to escalate privileges and execute arbitrary code in the context of the hypervisor. Was ZDI-CAN-13581.
CVE-2021-38473 22 Oct 2021
The affected product’s code base doesn’t properly control arguments for specific functions, which could lead to a stack overflow.
CVE-2021-3746 19 Oct 2021
A flaw was found in the libtpms code that may cause access beyond the boundary of internal buffers. The vulnerability is triggered by specially-crafted TPM2 command packets that then trigger the issue when the state of the TPM2's volatile state is written. The highest threat from this vulnerability is to system availability. This issue affects libtpms versions before 0.8.5, before 0.7.9 and before 0.6.6.
CVE-2021-38442 18 Oct 2021
FATEK Automation WinProladder versions 3.30 and prior lacks proper validation of user-supplied data when parsing project files, which could result in a heap-corruption condition. An attacker could leverage this vulnerability to execute code in the context of the current process.
CVE-2021-33737 14 Sep 2021
A vulnerability has been identified in SIMATIC CP 343-1 (incl. SIPLUS variants) (All versions), SIMATIC CP 343-1 Advanced (incl. SIPLUS variants) (All versions), SIMATIC CP 343-1 ERPC (All versions), SIMATIC CP 343-1 Lean (incl. SIPLUS variants) (All versions), SIMATIC CP 443-1 (incl. SIPLUS variants) (All versions), SIMATIC CP 443-1 Advanced (incl. SIPLUS variants) (All versions). Sending a specially crafted packet to port 102/tcp of an affected device could cause a Denial-of-Service condition. A restart is needed to restore normal operations.
CVE-2021-22789 2 Sep 2021
A CWE-119: Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability that could cause a Denial of Service on the Modicon PLC controller / simulator when updating the controller application with a specially crafted project file exists in Modicon M580 CPU (part numbers BMEP* and BMEH*, all versions), Modicon M340 CPU (part numbers BMXP34*, all versions), Modicon MC80 (part numbers BMKC80*, all versions), Modicon Momentum Ethernet CPU (part numbers 171CBU*, all versions), PLC Simulator for EcoStruxureª Control Expert, including all Unity Pro versions (former name of EcoStruxureª Control Expert, all versions), PLC Simulator for EcoStruxureª Process Expert including all HDCS versions (former name of EcoStruxureª Process Expert, all versions), Modicon Quantum CPU (part numbers 140CPU*, all versions), Modicon Premium CPU (part numbers TSXP5*, all versions).
CVE-2021-3634 31 Aug 2021
A flaw has been found in libssh in versions prior to 0.9.6. The SSH protocol keeps track of two shared secrets during the lifetime of the session. One of them is called secret_hash and the other session_id. Initially, both of them are the same, but after key re-exchange, previous session_id is kept and used as an input to new secret_hash. Historically, both of these buffers had shared length variable, which worked as long as these buffers were same. But the key re-exchange operation can also change the key exchange method, which can be based on hash of different size, eventually creating "secret_hash" of different size than the session_id has. This becomes an issue when the session_id memory is zeroed or when it is used again during second key re-exchange.
CVE-2021-34306 13 Jul 2021
A vulnerability has been identified in JT2Go (All versions < V13.2), Teamcenter Visualization (All versions < V13.2). The BMP_Loader.dll library in affected applications lacks proper validation of user-supplied data when parsing BMP files. This could result in a memory corruption condition. An attacker could leverage this vulnerability to execute code in the context of the current process. (ZDI-CAN-13342)
CVE-2021-3571 9 Jul 2021
A flaw was found in the ptp4l program of the linuxptp package. When ptp4l is operating on a little-endian architecture as a PTP transparent clock, a remote attacker could send a crafted one-step sync message to cause an information leak or crash. The highest threat from this vulnerability is to data confidentiality and system availability. This flaw affects linuxptp versions before 3.1.1 and before 2.0.1.
CVE-2021-3570 9 Jul 2021
A flaw was found in the ptp4l program of the linuxptp package. A missing length check when forwarding a PTP message between ports allows a remote attacker to cause an information leak, crash, or potentially remote code execution. The highest threat from this vulnerability is to data confidentiality and integrity as well as system availability. This flaw affects linuxptp versions before 3.1.1, before 2.0.1, before 1.9.3, before 1.8.1, before 1.7.1, before 1.6.1 and before 1.5.1.
CVE-2021-32992 29 Jun 2021
FATEK Automation WinProladder Versions 3.30 and prior do not properly restrict operations within the bounds of a memory buffer, which may allow an attacker to execute arbitrary code.
CVE-2021-3500 24 Jun 2021
A flaw was found in djvulibre-3.5.28 and earlier. A Stack overflow in function DJVU::DjVuDocument::get_djvu_file() via crafted djvu file may lead to application crash and other consequences.
CVE-2021-31495 15 Jun 2021
This vulnerability allows remote attackers to execute arbitrary code on affected installations of OpenText Brava! Desktop 16.6.3.84. User interaction is required to exploit this vulnerability in that the target must visit a malicious page or open a malicious file. The specific flaw exists within the parsing of DXF files. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to execute code in the context of the current process. Was ZDI-CAN-13307.
CVE-2021-31493 15 Jun 2021
This vulnerability allows remote attackers to execute arbitrary code on affected installations of OpenText Brava! Desktop 16.6.3.84. User interaction is required to exploit this vulnerability in that the target must visit a malicious page or open a malicious file. The specific flaw exists within the parsing of DXF files. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to execute code in the context of the current process. Was ZDI-CAN-13304.
CVE-2021-23169 8 Jun 2021
A heap-buffer overflow was found in the copyIntoFrameBuffer function of OpenEXR in versions before 3.0.1. An attacker could use this flaw to execute arbitrary code with the permissions of the user running the application compiled against OpenEXR.
CVE-2021-29579 14 May 2021
TensorFlow is an end-to-end open source platform for machine learning. The implementation of `tf.raw_ops.MaxPoolGrad` is vulnerable to a heap buffer overflow. The implementation(https://github.com/tensorflow/tensorflow/blob/ab1e644b48c82cb71493f4362b4dd38f4577a1cf/tensorflow/core/kernels/maxpooling_op.cc#L194-L203) fails to validate that indices used to access elements of input/output arrays are valid. Whereas accesses to `input_backprop_flat` are guarded by `FastBoundsCheck`, the indexing in `out_backprop_flat` can result in OOB access. The fix will be included in TensorFlow 2.5.0. We will also cherrypick this commit on TensorFlow 2.4.2, TensorFlow 2.3.3, TensorFlow 2.2.3 and TensorFlow 2.1.4, as these are also affected and still in supported range.
CVE-2021-29578 14 May 2021
TensorFlow is an end-to-end open source platform for machine learning. The implementation of `tf.raw_ops.FractionalAvgPoolGrad` is vulnerable to a heap buffer overflow. The implementation(https://github.com/tensorflow/tensorflow/blob/dcba796a28364d6d7f003f6fe733d82726dda713/tensorflow/core/kernels/fractional_avg_pool_op.cc#L216) fails to validate that the pooling sequence arguments have enough elements as required by the `out_backprop` tensor shape. The fix will be included in TensorFlow 2.5.0. We will also cherrypick this commit on TensorFlow 2.4.2, TensorFlow 2.3.3, TensorFlow 2.2.3 and TensorFlow 2.1.4, as these are also affected and still in supported range.
CVE-2021-29577 14 May 2021
TensorFlow is an end-to-end open source platform for machine learning. The implementation of `tf.raw_ops.AvgPool3DGrad` is vulnerable to a heap buffer overflow. The implementation(https://github.com/tensorflow/tensorflow/blob/d80ffba9702dc19d1fac74fc4b766b3fa1ee976b/tensorflow/core/kernels/pooling_ops_3d.cc#L376-L450) assumes that the `orig_input_shape` and `grad` tensors have similar first and last dimensions but does not check that this assumption is validated. The fix will be included in TensorFlow 2.5.0. We will also cherrypick this commit on TensorFlow 2.4.2, TensorFlow 2.3.3, TensorFlow 2.2.3 and TensorFlow 2.1.4, as these are also affected and still in supported range.
CVE-2021-29576 14 May 2021
TensorFlow is an end-to-end open source platform for machine learning. The implementation of `tf.raw_ops.MaxPool3DGradGrad` is vulnerable to a heap buffer overflow. The implementation(https://github.com/tensorflow/tensorflow/blob/596c05a159b6fbb9e39ca10b3f7753b7244fa1e9/tensorflow/core/kernels/pooling_ops_3d.cc#L694-L696) does not check that the initialization of `Pool3dParameters` completes successfully. Since the constructor(https://github.com/tensorflow/tensorflow/blob/596c05a159b6fbb9e39ca10b3f7753b7244fa1e9/tensorflow/core/kernels/pooling_ops_3d.cc#L48-L88) uses `OP_REQUIRES` to validate conditions, the first assertion that fails interrupts the initialization of `params`, making it contain invalid data. In turn, this might cause a heap buffer overflow, depending on default initialized values. The fix will be included in TensorFlow 2.5.0. We will also cherrypick this commit on TensorFlow 2.4.2, TensorFlow 2.3.3, TensorFlow 2.2.3 and TensorFlow 2.1.4, as these are also affected and still in supported range.
CVE-2021-29575 14 May 2021
TensorFlow is an end-to-end open source platform for machine learning. The implementation of `tf.raw_ops.ReverseSequence` allows for stack overflow and/or `CHECK`-fail based denial of service. The implementation(https://github.com/tensorflow/tensorflow/blob/5b3b071975e01f0d250c928b2a8f901cd53b90a7/tensorflow/core/kernels/reverse_sequence_op.cc#L114-L118) fails to validate that `seq_dim` and `batch_dim` arguments are valid. Negative values for `seq_dim` can result in stack overflow or `CHECK`-failure, depending on the version of Eigen code used to implement the operation. Similar behavior can be exhibited by invalid values of `batch_dim`. The fix will be included in TensorFlow 2.5.0. We will also cherrypick this commit on TensorFlow 2.4.2, TensorFlow 2.3.3, TensorFlow 2.2.3 and TensorFlow 2.1.4, as these are also affected and still in supported range.
CVE-2021-3507 6 May 2021
A heap buffer overflow was found in the floppy disk emulator of QEMU up to 6.0.0 (including). It could occur in fdctrl_transfer_handler() in hw/block/fdc.c while processing DMA read data transfers from the floppy drive to the guest system. A privileged guest user could use this flaw to crash the QEMU process on the host resulting in DoS scenario, or potential information leakage from the host memory.
CVE-2021-20204 6 May 2021
A heap memory corruption problem (use after free) can be triggered in libgetdata v0.10.0 when processing maliciously crafted dirfile databases. This degrades the confidentiality, integrity and availability of third-party software that uses libgetdata as a library. This vulnerability may lead to arbitrary code execution or privilege escalation depending on input/skills of attacker.
CVE-2021-1521 6 May 2021
A vulnerability in the Cisco Discovery Protocol implementation for Cisco Video Surveillance 8000 Series IP Cameras could allow an unauthenticated, adjacent attacker to cause an affected IP camera to reload. This vulnerability is due to missing checks when processing Cisco Discovery Protocol messages. An attacker could exploit this vulnerability by sending a malicious Cisco Discovery Protocol packet to an affected IP camera. A successful exploit could allow the attacker to cause the affected IP camera to reload unexpectedly, resulting in a denial of service (DoS) condition. Note: Cisco Discovery Protocol is a Layer 2 protocol. To exploit this vulnerability, an attacker must be in the same broadcast domain as the affected device (Layer 2 adjacent).
CVE-2021-1511 6 May 2021
Multiple vulnerabilities in Cisco SD-WAN vEdge Software could allow an attacker to execute arbitrary code as the root user or cause a denial of service (DoS) condition on an affected device. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1510 6 May 2021
Multiple vulnerabilities in Cisco SD-WAN vEdge Software could allow an attacker to execute arbitrary code as the root user or cause a denial of service (DoS) condition on an affected device. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1509 6 May 2021
Multiple vulnerabilities in Cisco SD-WAN vEdge Software could allow an attacker to execute arbitrary code as the root user or cause a denial of service (DoS) condition on an affected device. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1402 29 Apr 2021
A vulnerability in the software-based SSL/TLS message handler of Cisco Firepower Threat Defense (FTD) Software could allow an unauthenticated, remote attacker to trigger a reload of an affected device, resulting in a denial of service (DoS) condition. The vulnerability is due to insufficient validation of SSL/TLS messages when the device performs software-based SSL decryption. An attacker could exploit this vulnerability by sending a crafted SSL/TLS message through an affected device. SSL/TLS messages sent to an affected device do not trigger this vulnerability. A successful exploit could allow the attacker to cause a process to crash. This crash would then trigger a reload of the device. No manual intervention is needed to recover the device after the reload.
CVE-2021-20294 29 Apr 2021
A flaw was found in binutils readelf 2.35 program. An attacker who is able to convince a victim using readelf to read a crafted file could trigger a stack buffer overflow, out-of-bounds write of arbitrary data supplied by the attacker. The highest impact of this flaw is to confidentiality, integrity, and availability.
CVE-2021-21220 (v3: 8.8) 26 Apr 2021
Insufficient validation of untrusted input in V8 in Google Chrome prior to 89.0.4389.128 allowed a remote attacker to potentially exploit heap corruption via a crafted HTML page.
CVE-2021-3498 19 Apr 2021
GStreamer before 1.18.4 might cause heap corruption when parsing certain malformed Matroska files.
CVE-2021-1459 8 Apr 2021
A vulnerability in the web-based management interface of Cisco Small Business RV110W, RV130, RV130W, and RV215W Routers could allow an unauthenticated, remote attacker to execute arbitrary code on an affected device. The vulnerability is due to improper validation of user-supplied input in the web-based management interface. An attacker could exploit this vulnerability by sending crafted HTTP requests to a targeted device. A successful exploit could allow the attacker to execute arbitrary code as the root user on the underlying operating system of the affected device. Cisco has not released software updates that address this vulnerability.
CVE-2021-1472 8 Apr 2021
Multiple vulnerabilities exist in the web-based management interface of Cisco Small Business RV Series Routers. A remote attacker could execute arbitrary commands or bypass authentication and upload files on an affected device. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1473 8 Apr 2021
Multiple vulnerabilities exist in the web-based management interface of Cisco Small Business RV Series Routers. A remote attacker could execute arbitrary commands or bypass authentication and upload files on an affected device. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1480 8 Apr 2021
Multiple vulnerabilities in Cisco SD-WAN vManage Software could allow an unauthenticated, remote attacker to execute arbitrary code or allow an authenticated, local attacker to gain escalated privileges on an affected system. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1137 8 Apr 2021
Multiple vulnerabilities in Cisco SD-WAN vManage Software could allow an unauthenticated, remote attacker to execute arbitrary code or allow an authenticated, local attacker to gain escalated privileges on an affected system. For more information about these vulnerabilities, see the Details section of this advisory.
CVE-2021-1251 8 Apr 2021
Multiple vulnerabilities exist in the Link Layer Discovery Protocol (LLDP) implementation for Cisco Small Business RV Series Routers. An unauthenticated, adjacent attacker could execute arbitrary code or cause an affected router to leak system memory or reload. A memory leak or device reload would cause a denial of service (DoS) condition on an affected device. For more information about these vulnerabilities, see the Details section of this advisory. Note: LLDP is a Layer 2 protocol. To exploit these vulnerabilities, an attacker must be in the same broadcast domain as the affected device (Layer 2 adjacent).
CVE-2021-1308 8 Apr 2021
Multiple vulnerabilities exist in the Link Layer Discovery Protocol (LLDP) implementation for Cisco Small Business RV Series Routers. An unauthenticated, adjacent attacker could execute arbitrary code or cause an affected router to leak system memory or reload. A memory leak or device reload would cause a denial of service (DoS) condition on an affected device. For more information about these vulnerabilities, see the Details section of this advisory. Note: LLDP is a Layer 2 protocol. To exploit these vulnerabilities, an attacker must be in the same broadcast domain as the affected device (Layer 2 adjacent).

2020

CVE-2020-36619 19 Dec 2022
A vulnerability was found in multimon-ng. It has been rated as critical. This issue affects the function add_ch of the file demod_flex.c. The manipulation of the argument ch leads to format string. Upgrading to version 1.2.0 is able to address this issue. The name of the patch is e5a51c508ef952e81a6da25b43034dd1ed023c07. It is recommended to upgrade the affected component. The identifier VDB-216269 was assigned to this vulnerability.
CVE-2020-15782 28 May 2021
A vulnerability has been identified in SIMATIC Drive Controller family (All versions < V2.9.2), SIMATIC ET 200SP Open Controller CPU 1515SP PC (incl. SIPLUS variants) (All versions), SIMATIC ET 200SP Open Controller CPU 1515SP PC2 (incl. SIPLUS variants) (All versions), SIMATIC S7-1200 CPU family (incl. SIPLUS variants) (All versions < V4.5.0), SIMATIC S7-1500 CPU family (incl. related ET200 CPUs and SIPLUS variants) (All versions < V2.9.2), SIMATIC S7-1500 Software Controller (All versions), SIMATIC S7-PLCSIM Advanced (All versions < V4.0). Affected devices are vulnerable to a memory protection bypass through a specific operation. A remote unauthenticated attacker with network access to port 102/tcp could potentially write arbitrary data and code to protected memory areas or read sensitive data to launch further attacks.
CVE-2020-35521 9 Mar 2021
A flaw was found in libtiff. Due to a memory allocation failure in tif_read.c, a crafted TIFF file can lead to an abort, resulting in denial of service.
CVE-2020-35522 9 Mar 2021
In LibTIFF, there is a memory malloc failure in tif_pixarlog.c. A crafted TIFF document can lead to an abort, resulting in a remote denial of service attack.
CVE-2020-35524 9 Mar 2021
A heap-based buffer overflow flaw was found in libtiff in the handling of TIFF images in libtiff's TIFF2PDF tool. A specially crafted TIFF file can lead to arbitrary code execution. The highest threat from this vulnerability is to confidentiality, integrity, as well as system availability.
CVE-2020-25690 23 Feb 2021
An out-of-bounds write flaw was found in FontForge in versions before 20200314 while parsing SFD files containing certain LayerCount tokens. This flaw allows an attacker to manipulate the memory allocated on the heap, causing the application to crash or execute arbitrary code. The highest threat from this vulnerability is to confidentiality, integrity, as well as system availability.
CVE-2020-14360 20 Jan 2021
A flaw was found in the X.Org Server before version 1.20.10. An out-of-bounds access in the XkbSetMap function may lead to a privilege escalation vulnerability. The highest threat from this vulnerability is to data confidentiality and integrity as well as system availability.
CVE-2020-28220 11 Dec 2020
A CWE-119: Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability exists in Modicon M258 Firmware (All versions prior to V5.0.4.11) and SoMachine/SoMachine Motion software (All versions), that could cause a buffer overflow when the length of a file transferred to the webserver is not verified.
CVE-2020-7551 19 Nov 2020
A CWE-119 Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability exists in IGSS Definition (Def.exe) version 14.0.0.20247 that could cause Remote Code Execution when malicious CGF (Configuration Group File) file is imported to IGSS Definition.
CVE-2020-7552 19 Nov 2020
A CWE-119 Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability exists in IGSS Definition (Def.exe) version 14.0.0.20247 that could cause Remote Code Execution when malicious CGF (Configuration Group File) file is imported to IGSS Definition.
CVE-2020-7554 19 Nov 2020
A CWE-119 Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability exists in IGSS Definition (Def.exe) version 14.0.0.20247 that could cause Remote Code Execution when malicious CGF (Configuration Group File) file is imported to IGSS Definition.
CVE-2020-15254 16 Oct 2020
Crossbeam is a set of tools for concurrent programming. In crossbeam-channel before version 0.4.4, the bounded channel incorrectly assumes that `Vec::from_iter` has allocated capacity that same as the number of iterator elements. `Vec::from_iter` does not actually guarantee that and may allocate extra memory. The destructor of the `bounded` channel reconstructs `Vec` from the raw pointer based on the incorrect assumes described above. This is unsound and causing deallocation with the incorrect capacity when `Vec::from_iter` has allocated different sizes with the number of iterator elements. This has been fixed in crossbeam-channel 0.4.4.
CVE-2020-15195 25 Sep 2020
In Tensorflow before versions 1.15.4, 2.0.3, 2.1.2, 2.2.1 and 2.3.1, the implementation of `SparseFillEmptyRowsGrad` uses a double indexing pattern. It is possible for `reverse_index_map(i)` to be an index outside of bounds of `grad_values`, thus resulting in a heap buffer overflow. The issue is patched in commit 390611e0d45c5793c7066110af37c8514e6a6c54, and is released in TensorFlow versions 1.15.4, 2.0.3, 2.1.2, 2.2.1, or 2.3.1.
CVE-2020-15196 25 Sep 2020
In Tensorflow version 2.3.0, the `SparseCountSparseOutput` and `RaggedCountSparseOutput` implementations don't validate that the `weights` tensor has the same shape as the data. The check exists for `DenseCountSparseOutput`, where both tensors are fully specified. In the sparse and ragged count weights are still accessed in parallel with the data. But, since there is no validation, a user passing fewer weights than the values for the tensors can generate a read from outside the bounds of the heap buffer allocated for the weights. The issue is patched in commit 3cbb917b4714766030b28eba9fb41bb97ce9ee02 and is released in TensorFlow version 2.3.1.
CVE-2020-15198 25 Sep 2020
In Tensorflow before version 2.3.1, the `SparseCountSparseOutput` implementation does not validate that the input arguments form a valid sparse tensor. In particular, there is no validation that the `indices` tensor has the same shape as the `values` one. The values in these tensors are always accessed in parallel. Thus, a shape mismatch can result in accesses outside the bounds of heap allocated buffers. The issue is patched in commit 3cbb917b4714766030b28eba9fb41bb97ce9ee02 and is released in TensorFlow version 2.3.1.
CVE-2020-15205 25 Sep 2020
In Tensorflow before versions 1.15.4, 2.0.3, 2.1.2, 2.2.1 and 2.3.1, the `data_splits` argument of `tf.raw_ops.StringNGrams` lacks validation. This allows a user to pass values that can cause heap overflow errors and even leak contents of memory In the linked code snippet, all the binary strings after `ee ff` are contents from the memory stack. Since these can contain return addresses, this data leak can be used to defeat ASLR. The issue is patched in commit 0462de5b544ed4731aa2fb23946ac22c01856b80, and is released in TensorFlow versions 1.15.4, 2.0.3, 2.1.2, 2.2.1, or 2.3.1.
CVE-2020-15207 25 Sep 2020
In tensorflow-lite before versions 1.15.4, 2.0.3, 2.1.2, 2.2.1 and 2.3.1, to mimic Python's indexing with negative values, TFLite uses `ResolveAxis` to convert negative values to positive indices. However, the only check that the converted index is now valid is only present in debug builds. If the `DCHECK` does not trigger, then code execution moves ahead with a negative index. This, in turn, results in accessing data out of bounds which results in segfaults and/or data corruption. The issue is patched in commit 2d88f470dea2671b430884260f3626b1fe99830a, and is released in TensorFlow versions 1.15.4, 2.0.3, 2.1.2, 2.2.1, or 2.3.1.
CVE-2020-15213 25 Sep 2020
In TensorFlow Lite before versions 2.2.1 and 2.3.1, models using segment sum can trigger a denial of service by causing an out of memory allocation in the implementation of segment sum. Since code uses the last element of the tensor holding them to determine the dimensionality of output tensor, attackers can use a very large value to trigger a large allocation. The issue is patched in commit 204945b19e44b57906c9344c0d00120eeeae178a and is released in TensorFlow versions 2.2.1, or 2.3.1. A potential workaround would be to add a custom `Verifier` to limit the maximum value in the segment ids tensor. This only handles the case when the segment ids are stored statically in the model, but a similar validation could be done if the segment ids are generated at runtime, between inference steps. However, if the segment ids are generated as outputs of a tensor during inference steps, then there are no possible workaround and users are advised to upgrade to patched code.
CVE-2020-15173 10 Sep 2020
In ACCEL-PPP (an implementation of PPTP/PPPoE/L2TP/SSTP), there is a buffer overflow when receiving an l2tp control packet ith an AVP which type is a string and no hidden flags, length set to less than 6. If your application is used in open networks or there are untrusted nodes in the network it is highly recommended to apply the patch. The problem was patched with commit 2324bcd5ba12cf28f47357a8f03cd41b7c04c52b As a workaround changes of commit 2324bcd5ba12cf28f47357a8f03cd41b7c04c52b can be applied to older versions.
CVE-2020-15158 26 Aug 2020
In libIEC61850 before version 1.4.3, when a message with COTP message length field with value < 4 is received an integer underflow will happen leading to heap buffer overflow. This can cause an application crash or on some platforms even the execution of remote code. If your application is used in open networks or there are untrusted nodes in the network it is highly recommend to apply the patch. This was patched with commit 033ab5b. Users of version 1.4.x should upgrade to version 1.4.3 when available. As a workaround changes of commit 033ab5b can be applied to older versions.
CVE-2020-17397 25 Aug 2020
This vulnerability allows local attackers to escalate privileges on affected installations of Parallels Desktop 15.1.4. An attacker must first obtain the ability to execute high-privileged code on the target guest system in order to exploit this vulnerability. The specific flaw exists within the handling of network packets. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to escalate privileges and execute code in the context of the hypervisor. Was ZDI-CAN-11253.
CVE-2020-4549 (v3: 7.8) 3 Aug 2020
IBM i2 Analyst Notebook 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 183317.
CVE-2020-4550 (v3: 7.8) 3 Aug 2020
IBM i2 Analyst Notebook 9.2.1 and 9.2.2 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 183318.
CVE-2020-4551 (v3: 7.8) 3 Aug 2020
IBM i2 Analyst Notebook 9.2.1 and 9.2.2 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 183319.
CVE-2020-4552 (v3: 7.8) 3 Aug 2020
IBM i2 Analyst Notebook 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 183320.
CVE-2020-4553 (v3: 7.8) 3 Aug 2020
IBM i2 Analyst Notebook 9.2.1 and 9.2.2 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 183321.
CVE-2020-4554 (v3: 7.8) 3 Aug 2020
IBM i2 Analyst Notebook 9.2.1 and 9.2.2 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 183322.
CVE-2020-0160 (v3: 8.8) 11 Jun 2020
In setSyncSampleParams of SampleTable.cpp, there is possible resource exhaustion due to a missing bounds check. This could lead to remote denial of service with no additional execution privileges needed. User interaction is needed for exploitation.Product: AndroidVersions: Android-10Android ID: A-124771364
CVE-2020-13831 (v3: 9.8) 4 Jun 2020
An issue was discovered on Samsung mobile devices with O(8.x) and P(9.0) (Exynos 7570 chipsets) software. The Trustonic Kinibi component allows arbitrary memory mapping. The Samsung ID is SVE-2019-16665 (June 2020).
CVE-2020-13832 (v3: 9.8) 4 Jun 2020
An issue was discovered on Samsung mobile devices with Q(10.0) (with TEEGRIS on Exynos chipsets) software. The Widevine Trustlet allows arbitrary code execution because of memory disclosure, The Samsung IDs are SVE-2020-17117, SVE-2020-17118, SVE-2020-17119, and SVE-2020-17161 (June 2020).
CVE-2020-6453 (v3: 8.8) 4 Jun 2020
Inappropriate implementation in V8 in Google Chrome prior to 80.0.3987.162 allowed a remote attacker to potentially exploit heap corruption via a crafted HTML page.
CVE-2020-13759 (v3: 7.5) 2 Jun 2020
rust-vmm vm-memory before 0.1.1 and 0.2.x before 0.2.1 allows attackers to cause a denial of service (loss of IP networking) because read_obj and write_obj do not properly access memory. This affects aarch64 (with musl or glibc) and x86_64 (with musl).
CVE-2020-3630 (v3: 7.8) 2 Jun 2020
Possibility of out of bound access while processing the responses from video firmware in Snapdragon Auto, Snapdragon Compute, Snapdragon Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wearables in APQ8009, APQ8053, APQ8096AU, APQ8098, Kamorta, MDM9150, MDM9206, MDM9207C, MDM9607, MDM9640, MDM9650, MSM8909W, MSM8917, MSM8953, MSM8996, MSM8996AU, MSM8998, Nicobar, QCM2150, QCS405, QCS605, QM215, Rennell, SA415M, SA6155P, Saipan, SC8180X, SDA660, SDM429, SDM429W, SDM439, SDM450, SDM630, SDM632, SDM636, SDM660, SDM670, SDM710, SDM845, SDX20, SDX24, SDX55, SM6150, SM7150, SM8150, SM8250, SXR1130, SXR2130
CVE-2020-13754 (v3: 6.7) 2 Jun 2020
hw/pci/msix.c in QEMU 4.2.0 allows guest OS users to trigger an out-of-bounds access via a crafted address in an msi-x mmio operation.
CVE-2020-4285 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by a memory corruption error. By persuading a victim to open a specially-crafted document, a remote attacker could exploit this vulnerability to execute arbitrary code on the system with the privileges of the victim or cause the application to crash. IBM X-Force ID: 176266
CVE-2020-4287 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by a memory corruption error. By persuading a victim to open a specially-crafted document, a remote attacker could exploit this vulnerability to execute arbitrary code on the system with the privileges of the victim or cause the application to crash. IBM X-Force ID: 176269.
CVE-2020-4288 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by a memory corruption error. By persuading a victim to open a specially-crafted document, a remote attacker could exploit this vulnerability to execute arbitrary code on the system with the privileges of the victim or cause the application to crash. IBM X-Force ID: 176270.
CVE-2020-4343 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially crafted file, a remote attacker could exploit this vulnerability to execute arbitrary code on the system or cause the application to crash. IBM X-Force ID: 178244.
CVE-2020-4422 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially crafted file, a remote attacker could exploit this vulnerability to execute arbitrary code on the system or cause the application to crash. IBM X-Force ID: 180167.
CVE-2020-4467 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by memory corruption. By persuading a victim to open a specially-crafted document, a remote attacker could exploit this vulnerability to execute arbitrary code on the system with the privileges of the victim or cause the application to crash. IBM X-Force ID: 181721.
CVE-2020-4468 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a remote attacker to execute arbitrary code on the system, caused by memory corruption. By persuading a victim to open a specially-crafted document, a remote attacker could exploit this vulnerability to execute arbitrary code on the system with the privileges of the victim or cause the application to crash. IBM X-Force ID: 181723.
CVE-2020-4257 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175635.
CVE-2020-4258 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175637.
CVE-2020-4261 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175644.
CVE-2020-4262 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175645.
CVE-2020-4263 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175646.
CVE-2020-4264 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175647.
CVE-2020-4265 (v3: 7.3) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175648.
CVE-2020-4266 (v3: 7.8) 14 May 2020
IBM i2 Intelligent Analyis Platform 9.2.1 could allow a local attacker to execute arbitrary code on the system, caused by a memory corruption. By persuading a victim to open a specially-crafted file, an attacker could exploit this vulnerability to execute arbitrary code on the system. IBM X-Force ID: 175649.
CVE-2020-12751 (v3: 7.8) 11 May 2020
An issue was discovered on Samsung mobile devices with O(8.X), P(9.0), and Q(10.0) software. The Quram image codec library allows attackers to overwrite memory and execute arbitrary code via crafted JPEG data that is mishandled during decoding. The Samsung ID is SVE-2020-16943 (May 2020).

2019

CVE-2019-25078 13 Dec 2022
A vulnerability classified as problematic was found in pacparser up to 1.3.x. Affected by this vulnerability is the function pacparser_find_proxy of the file src/pacparser.c. The manipulation of the argument url leads to buffer overflow. Attacking locally is a requirement. Upgrading to version 1.4.0 is able to address this issue. The name of the patch is 853e8f45607cb07b877ffd270c63dbcdd5201ad9. It is recommended to upgrade the affected component. The associated identifier of this vulnerability is VDB-215443.
CVE-2019-14077 (v3: 7.8) 2 Jun 2020
Out of bound memory access while processing ese transmit command due to passing Response buffer received from user in Snapdragon Auto, Snapdragon Compute, Snapdragon Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wired Infrastructure and Networking in APQ8009, APQ8098, IPQ6018, Kamorta, MDM9150, MDM9205, MDM9607, MDM9650, MSM8909, MSM8998, Nicobar, QCS404, QCS405, QCS605, Rennell, SA415M, SA6155P, SC7180, SC8180X, SDA660, SDA845, SDM630, SDM636, SDM660, SDM670, SDM710, SDM845, SDM850, SDX24, SDX55, SM6150, SM7150, SM8150, SM8250, SXR1130, SXR2130
CVE-2019-14078 (v3: 7.8) 2 Jun 2020
Out of bound memory access while processing qpay due to not validating length of the response buffer provided by User. in Snapdragon Auto, Snapdragon Compute, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wired Infrastructure and Networking in APQ8009, APQ8098, MSM8909, MSM8998, SDA660, SDA845, SDM630, SDM636, SDM660, SDM845
CVE-2019-15880 (v3: 9.8) 13 May 2020
In FreeBSD 12.1-STABLE before r356911, and 12.1-RELEASE before p5, insufficient checking in the cryptodev module allocated the size of a kernel buffer based on a user-supplied length allowing an unprivileged process to trigger a kernel panic.
CVE-2019-14009 (v3: 7.8) 16 Apr 2020
Out of bound memory access while processing TZ command handler due to improper input validation on response length received from user in Snapdragon Auto, Snapdragon Compute, Snapdragon Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wired Infrastructure and Networking in APQ8009, APQ8098, MDM9150, MDM9607, MDM9650, MSM8905, MSM8909, MSM8998, SDA660, SDA845, SDM630, SDM636, SDM660, SDM845, SDM850, SXR2130
CVE-2019-5105 (v3: 7.5) 26 Mar 2020
An exploitable memory corruption vulnerability exists in the Name Service Client functionality of 3S-Smart Software Solutions CODESYS GatewayService 3.5.13.20. A specially crafted packet can cause a large memcpy, resulting in an access violation and termination of the process. An attacker can send a packet to a device running the GatewayService.exe to trigger this vulnerability.
CVE-2019-20601 (v3: 7.5) 24 Mar 2020
An issue was discovered on Samsung mobile devices with N(7.x), O(8.x), and P(9.0) (Exynos7570, 7580, 7870, 7880, and 8890 chipsets) software. RKP memory corruption causes an arbitrary write to protected memory. The Samsung ID is SVE-2019-13921-2 (May 2019).
CVE-2019-20549 (v3: 9.8) 24 Mar 2020
An issue was discovered on Samsung mobile devices with N(7.x), O(8.x), and P(9.0) (Broadcom chipsets) software. A heap out-of-bounds access can occur during LE Packet reception in Broadcom Bluetooth. The Samsung ID is SVE-2019-15724 (November 2019).
CVE-2019-20553 (v3: 9.8) 24 Mar 2020
An issue was discovered on Samsung mobile devices with P(9.0) (SM6150, SM8150, SM8150_FUSION, exynos7885, exynos9610, and exynos9820 chipsets) software. Arbitrary memory read and write operations can occur in RKP. The Samsung ID is SVE-2019-15143 (October 2019).
CVE-2019-20556 (v3: 9.8) 24 Mar 2020
An issue was discovered on Samsung mobile devices with P(9.0) (SM6150, SM8150, SM8150_FUSION, exynos7885, exynos9610, and exynos9820 chipsets) software. RKP memory corruption allows attackers to control the effective address in EL2. The Samsung ID is SVE-2019-15221 (October 2019).
CVE-2019-20566 (v3: 9.8) 24 Mar 2020
An issue was discovered on Samsung mobile devices with any (before September 2019 for SMP1300 Exynos modem chipsets) software. Attackers can trigger stack corruption in the Shannon modem via a crafted RP-Originator/Destination address. The Samsung ID is SVE-2019-14858 (September 2019).
CVE-2019-20577 (v3: 7.5) 24 Mar 2020
An issue was discovered on Samsung mobile devices with P(9.0) (Exynos chipsets) software. The MALI GPU Driver allows a kernel panic. The Samsung ID is SVE-2019-14372 (August 2019).
CVE-2019-19945 (v3: 7.5) 16 Mar 2020
uhttpd in OpenWrt through 18.06.5 and 19.x through 19.07.0-rc2 has an integer signedness error. This leads to out-of-bounds access to a heap buffer and a subsequent crash. It can be triggered with an HTTP POST request to a CGI script, specifying both "Transfer-Encoding: chunked" and a large negative Content-Length value.
CVE-2019-14310 (v3: 9.8) 13 Mar 2020
Ricoh SP C250DN 1.05 devices allow denial of service (issue 2 of 3). Unauthenticated crafted packets to the IPP service will cause a vulnerable device to crash. A memory corruption has been identified in the way of how the embedded device parsed the IPP packets
CVE-2019-17135 (v3: 7.8) 8 Feb 2020
This vulnerability allows remote attackers to execute arbitrary code on affected installations of Foxit PhantomPDF 9.5.0.20723. User interaction is required to exploit this vulnerability in that the target must visit a malicious page or open a malicious file. The specific flaw exists within the parsing of DXF files. The issue results from the lack of proper validation of user-supplied data, which can result in a memory corruption condition. An attacker can leverage this vulnerability to execute code in the context of the current process. Was ZDI-CAN-8775.
CVE-2019-20391 (v3: 6.5) 22 Jan 2020
An invalid memory access flaw is present in libyang before v1.0-r3 in the function resolve_feature_value() when an if-feature statement is used inside a bit. Applications that use libyang to parse untrusted input yang files may crash.
CVE-2019-20392 (v3: 6.5) 22 Jan 2020
An invalid memory access flaw is present in libyang before v1.0-r1 in the function resolve_feature_value() when an if-feature statement is used inside a list key node, and the feature used is not defined. Applications that use libyang to parse untrusted input yang files may crash.
CVE-2019-19339 (v3: 6.5) 17 Jan 2020
It was found that the Red Hat Enterprise Linux 8 kpatch update did not include the complete fix for CVE-2018-12207. A flaw was found in the way Intel CPUs handle inconsistency between, virtual to physical memory address translations in CPU's local cache and system software's Paging structure entries. A privileged guest user may use this flaw to induce a hardware Machine Check Error on the host processor, resulting in a severe DoS scenario by halting the processor. System software like OS OR Virtual Machine Monitor (VMM) use virtual memory system for storing program instructions and data in memory. Virtual Memory system uses Paging structures like Page Tables and Page Directories to manage system memory. The processor's Memory Management Unit (MMU) uses Paging structure entries to translate program's virtual memory addresses to physical memory addresses. The processor stores these address translations into its local cache buffer called - Translation Lookaside Buffer (TLB). TLB has two parts, one for instructions and other for data addresses. System software can modify its Paging structure entries to change address mappings OR certain attributes like page size etc. Upon such Paging structure alterations in memory, system software must invalidate the corresponding address translations in the processor's TLB cache. But before this TLB invalidation takes place, a privileged guest user may trigger an instruction fetch operation, which could use an already cached, but now invalid, virtual to physical address translation from Instruction TLB (ITLB). Thus accessing an invalid physical memory address and resulting in halting the processor due to the Machine Check Error (MCE) on Page Size Change.
CVE-2019-17025 (v3: 8.8) 8 Jan 2020
Mozilla developers reported memory safety bugs present in Firefox 71. Some of these bugs showed evidence of memory corruption and we presume that with enough effort some of these could have been exploited to run arbitrary code. This vulnerability affects Firefox < 72.
CVE-2019-11758 (v3: 8.8) 8 Jan 2020
Mozilla community member Philipp reported a memory safety bug present in Firefox 68 when 360 Total Security was installed. This bug showed evidence of memory corruption in the accessibility engine and we presume that with enough effort that it could be exploited to run arbitrary code. This vulnerability affects Firefox < 69, Thunderbird < 68.2, and Firefox ESR < 68.2.
CVE-2019-13465 (v3: 8.6) 30 Dec 2019
An issue was discovered in the ROS communications-related packages (aka ros_comm or ros-melodic-ros-comm) through 1.14.3. ROS_ASSERT_MSG only works when ROS_ASSERT_ENABLED is defined. This leads to a problem in the remove() function in clients/roscpp/src/libros/spinner.cpp. When ROS_ASSERT_ENABLED is not defined, the iterator loop will run out of the scope of the array, and cause denial of service for other components (that depend on the communication-related functions of this package). NOTE: The reporter of this issue now believes it was a false alarm.
CVE-2019-18236 (v3: 7.8) 23 Dec 2019
Multiple buffer overflow vulnerabilities exist when the PLC Editor Version 1.3.5_20190129 processes project files. An attacker could use a specially crafted project file to exploit and execute code under the privileges of the application.
CVE-2019-8253 (v3: 7.8) 19 Dec 2019
Adobe Photoshop CC versions before 20.0.8 and 21.0.x before 21.0.2 have a memory corruption vulnerability. Successful exploitation could lead to arbitrary code execution.
CVE-2019-8254 (v3: 7.8) 19 Dec 2019
Adobe Photoshop CC versions before 20.0.8 and 21.0.x before 21.0.2 have a memory corruption vulnerability. Successful exploitation could lead to arbitrary code execution.
CVE-2019-16462 (v3: 9.8) 19 Dec 2019
Adobe Acrobat and Reader versions , 2019.021.20056 and earlier, 2017.011.30152 and earlier, 2017.011.30155 and earlier version, 2017.011.30152 and earlier, and 2015.006.30505 and earlier have a buffer error vulnerability. Successful exploitation could lead to arbitrary code execution .
CVE-2019-5075 (v3: 9.8) 18 Dec 2019
An exploitable stack buffer overflow vulnerability exists in the command line utility getcouplerdetails of WAGO PFC200 Firmware versions 03.01.07(13) and 03.00.39(12), and WAGO PFC100 Firmware version 03.00.39(12). A specially crafted set of packets sent to the iocheckd service "I/O-Check" can cause a stack buffer overflow in the sub-process getcouplerdetails, resulting in code execution. An attacker can send unauthenticated packets to trigger this vulnerability.
CVE-2019-8504 (v3: 5.5) 18 Dec 2019
A memory initialization issue was addressed with improved memory handling. This issue is fixed in iOS 12.2, macOS Mojave 10.14.4. A local user may be able to read kernel memory.
CVE-2019-6237 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, Safari 12.1.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-6239 (v3: 7.8) 18 Dec 2019
This issue was addressed with improved handling of file metadata. This issue is fixed in macOS Mojave 10.14.4. A malicious application may bypass Gatekeeper checks.
CVE-2019-8518 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8523 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8535 (v3: 8.8) 18 Dec 2019
A memory corruption issue was addressed with improved state management. This issue is fixed in iOS 12.2, tvOS 12.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8536 (v3: 8.8) 18 Dec 2019
A memory corruption issue was addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8544 (v3: 8.8) 18 Dec 2019
A memory corruption issue was addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8545 (v3: 7.1) 18 Dec 2019
A memory corruption issue was addressed with improved state management. This issue is fixed in iOS 12.2, macOS Mojave 10.14.4, tvOS 12.2, watchOS 5.2. A local user may be able to cause unexpected system termination or read kernel memory.
CVE-2019-8552 (v3: 7.8) 18 Dec 2019
A memory initialization issue was addressed with improved memory handling. This issue is fixed in iOS 12.2, macOS Mojave 10.14.4, tvOS 12.2, watchOS 5.2. A malicious application may be able to elevate privileges.
CVE-2019-8553 (v3: 8.8) 18 Dec 2019
A memory corruption issue was addressed with improved validation. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2. Clicking a malicious SMS link may lead to arbitrary code execution.
CVE-2019-8555 (v3: 7.8) 18 Dec 2019
A buffer overflow was addressed with improved size validation. This issue is fixed in macOS Mojave 10.14.4. A malicious application may be able to execute arbitrary code with kernel privileges.
CVE-2019-8558 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8559 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8562 (v3: 9.6) 18 Dec 2019
A memory corruption issue was addressed with improved validation. This issue is fixed in iOS 12.2, tvOS 12.2, Safari 12.1, iTunes 12.9.4 for Windows. A sandboxed process may be able to circumvent sandbox restrictions.
CVE-2019-8563 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.2, tvOS 12.2, watchOS 5.2, Safari 12.1, iTunes 12.9.4 for Windows, iCloud for Windows 7.11. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8574 (v3: 7.8) 18 Dec 2019
A memory corruption issue was addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, watchOS 5.2.1. An application may be able to execute arbitrary code with system privileges.
CVE-2019-8583 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, watchOS 5.2.1, Safari 12.1.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8586 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, Safari 12.1.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8593 (v3: 7.8) 18 Dec 2019
A memory corruption issue was addressed with improved memory handling. This issue is fixed in iOS 12.3, tvOS 12.3, watchOS 5.2.1. An application may be able to execute arbitrary code with system privileges.
CVE-2019-8594 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, Safari 12.1.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8595 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, Safari 12.1.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8596 (v3: 8.8) 18 Dec 2019
Multiple memory corruption issues were addressed with improved memory handling. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, Safari 12.1.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. Processing maliciously crafted web content may lead to arbitrary code execution.
CVE-2019-8602 (v3: 7.8) 18 Dec 2019
A memory corruption issue was addressed by removing the vulnerable code. This issue is fixed in iOS 12.3, macOS Mojave 10.14.5, tvOS 12.3, watchOS 5.2.1, iTunes for Windows 12.9.5, iCloud for Windows 7.12. A malicious application may be able to elevate privileges.

2018

CVE-2018-21052 (v3: 9.8) 8 Apr 2020
An issue was discovered on Samsung mobile devices with N(7.x) and O(8.X) (Exynos chipsets) software. There is incorrect usage of shared memory in the vaultkeeper Trustlet, leading to arbitrary code execution. The Samsung ID is SVE-2018-12855 (October 2018).
CVE-2018-21072 (v3: 9.8) 8 Apr 2020
An issue was discovered on Samsung mobile devices with M(6.0), N(7.x), and O(8.0) (Exynos chipsets) software. A kernel driver allows out-of-bounds Read/Write operations and possibly arbitrary code execution. The Samsung ID is SVE-2018-11358 (May 2018).
CVE-2018-21027 (v3: 9.8) 11 Oct 2019
Boa through 0.94.14rc21 allows remote attackers to trigger an out-of-memory (OOM) condition because malloc is mishandled.
CVE-2018-5732 (v3: 7.5) 9 Oct 2019
Failure to properly bounds-check a buffer used for processing DHCP options allows a malicious server (or an entity masquerading as a server) to cause a buffer overflow (and resulting crash) in dhclient by sending a response containing a specially constructed options section. Affects ISC DHCP versions 4.1.0 -> 4.1-ESV-R15, 4.2.0 -> 4.2.8, 4.3.0 -> 4.3.6, 4.4.0
CVE-2018-11768 (v3: 7.5) 4 Oct 2019
In Apache Hadoop 3.1.0 to 3.1.1, 3.0.0-alpha1 to 3.0.3, 2.9.0 to 2.9.1, and 2.0.0-alpha to 2.8.4, the user/group information can be corrupted across storing in fsimage and reading back from fsimage.
CVE-2018-21010 (v3: 8.8) 5 Sep 2019
OpenJPEG before 2.3.1 has a heap buffer overflow in color_apply_icc_profile in bin/common/color.c.
CVE-2018-20995 (v3: 9.8) 26 Aug 2019
An issue was discovered in the slice-deque crate before 0.1.16 for Rust. move_head_unchecked allows memory corruption because deque updates are mishandled.
CVE-2018-20998 (v3: 9.8) 26 Aug 2019
An issue was discovered in the arrayfire crate before 3.6.0 for Rust. Addition of the repr() attribute to an enum is mishandled, leading to memory corruption.
CVE-2018-21000 (v3: 9.8) 26 Aug 2019
An issue was discovered in the safe-transmute crate before 0.10.1 for Rust. A constructor's arguments are in the wrong order, causing heap memory corruption.
CVE-2018-20992 (v3: 6.5) 26 Aug 2019
An issue was discovered in the claxon crate before 0.4.1 for Rust. Uninitialized memory can be exposed because certain decode buffer sizes are mishandled.
CVE-2018-20994 (v3: 7.5) 26 Aug 2019
An issue was discovered in the trust-dns-proto crate before 0.5.0-alpha.3 for Rust. There is infinite recursion because DNS message compression is mishandled.
CVE-2018-1635 (v3: 6.7) 20 Aug 2019
Stack-based buffer overflow in oninit in IBM Informix Dynamic Server Enterprise Edition 12.1 allows an authenticated user to execute predefined code with root privileges, such as escalating to a root shell. IBM X-Force ID: 144439.
CVE-2018-1636 (v3: 6.7) 20 Aug 2019
Stack-based buffer overflow in oninit in IBM Informix Dynamic Server Enterprise Edition 12.1 allows an authenticated user to execute predefined code with root privileges, such as escalating to a root shell. IBM X-Force ID: 144441.
CVE-2018-20855 (v3: 3.3) 26 Jul 2019
An issue was discovered in the Linux kernel before 4.18.7. In create_qp_common in drivers/infiniband/hw/mlx5/qp.c, mlx5_ib_create_qp_resp was never initialized, resulting in a leak of stack memory to userspace.
CVE-2018-13924 (v3: 9.8) 22 Jul 2019
Lack of check to prevent the buffer length taking negative values can lead to stack overflow. in Snapdragon Auto, Snapdragon Compute, Snapdragon Consumer Electronics Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon IoT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wearables, Snapdragon Wired Infrastructure and Networking in IPQ8074, MDM9150, MDM9206, MDM9607, MDM9615, MDM9625, MDM9635M, MDM9640, MDM9650, MDM9655, MSM8909W, MSM8996AU, QCA6174A, QCA8081, QCS404, QCS405, QCS605, Qualcomm 215, SD 210/SD 212/SD 205, SD 425, SD 427, SD 430, SD 435, SD 439 / SD 429, SD 450, SD 615/16/SD 415, SD 625, SD 632, SD 636, SD 650/52, SD 665, SD 675, SD 712 / SD 710 / SD 670, SD 730, SD 820, SD 820A, SD 835, SD 845 / SD 850, SD 855, SD 8CX, SDA660, SDM439, SDM630, SDM660, SDX20, Snapdragon_High_Med_2016, SXR1130
CVE-2018-7838 (v3: 7.5) 15 Jul 2019
A CWE-119 Buffer Errors vulnerability exists in Modicon M580 CPU - BMEP582040, all versions before V2.90, and Modicon Ethernet Module BMENOC0301, all versions before V2.16, which could cause denial of service on the FTP service of the controller or the Ethernet BMENOC module when it receives a FTP CWD command with a data length greater than 1020 bytes. A power cycle is then needed to reactivate the FTP service.
CVE-2018-14496 (v3: 9.8) 10 Jul 2019
** DISPUTED ** Vivotek FD8136 devices allow remote memory corruption and remote code execution because of a stack-based buffer overflow, related to sprintf, vlocal_buff_4326, and set_getparam.cgi. NOTE: The vendor has disputed this as a vulnerability and states that the issue does not cause a web server crash or have any other affect on it's performance.
CVE-2018-14550 (v3: 8.8) 10 Jul 2019
An issue has been found in third-party PNM decoding associated with libpng 1.6.35. It is a stack-based buffer overflow in the function get_token in pnm2png.c in pnm2png.
CVE-2018-11423 (v3: 7.5) 3 Jul 2019
There is Memory corruption in the web interface Moxa OnCell G3100-HSPA Series version 1.6 Build 17100315 and prior, different vulnerability than CVE-2018-11420.
CVE-2018-11425 (v3: 9.8) 3 Jul 2019
Memory corruption issue was discovered in Moxa OnCell G3470A-LTE Series version 1.6 Build 18021314 and prior, a different vulnerability than CVE-2018-11424.
CVE-2018-15519 (v3: 9.8) 28 Jun 2019
Various Lexmark devices have a Buffer Overflow (issue 1 of 2).
CVE-2018-15520 (v3: 9.8) 28 Jun 2019
Various Lexmark devices have a Buffer Overflow (issue 2 of 2).
CVE-2018-6154 (v3: 8.8) 27 Jun 2019
Insufficient data validation in WebGL in Google Chrome prior to 68.0.3440.75 allowed a remote attacker to potentially exploit heap corruption via a crafted HTML page.
CVE-2018-6156 (v3: 8.8) 27 Jun 2019
Incorect derivation of a packet length in WebRTC in Google Chrome prior to 68.0.3440.75 allowed a remote attacker to potentially exploit heap corruption via a crafted video file.
CVE-2018-16119 (v3: 7.2) 20 Jun 2019
Stack-based buffer overflow in the httpd server of TP-Link WR1043nd (Firmware Version 3) allows remote attackers to execute arbitrary code via a malicious MediaServer request to /userRpm/MediaServerFoldersCfgRpm.htm.
CVE-2018-16595 (v3: 6.5) 19 Jun 2019
The Photo Sharing Plus component on Sony Bravia TV through 8.587 devices has a Buffer Overflow.
CVE-2018-18944 (v3: 7.5) 18 Jun 2019
Artha ~ The Open Thesaurus 1.0.3.0 has a Buffer Overflow.
CVE-2018-19447 (v3: 7.8) 17 Jun 2019
A stack-based buffer overflow can occur for specially crafted PDF files in Foxit Reader SDK (ActiveX) 5.4.0.1031 when parsing the URI string. An attacker can leverage this to gain remote code execution.
CVE-2018-20655 (v3: 9.8) 14 Jun 2019
When receiving calls using WhatsApp for iOS, a missing size check when parsing a sender-provided packet allowed for a stack-based overflow. This issue affects WhatsApp for iOS prior to v2.18.90.24 and WhatsApp Business for iOS prior to v2.18.90.24.
CVE-2018-5911 (v3: 7.8) 14 Jun 2019
Buffer overflow in WLAN function due to improper check of buffer size before copying in Snapdragon Auto, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile in MDM9150, MDM9206, MDM9607, MDM9640, MDM9650, MSM8996AU, QCS605, SD 625, SD 636, SD 675, SD 712 / SD 710 / SD 670, SD 730, SD 820A, SD 855, SDM630, SDM660, SDX20, SDX24
CVE-2018-11929 (v3: 7.8) 14 Jun 2019
Lack of input validation in WLAN function can lead to potential heap overflow in Snapdragon Auto, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music in MDM9150, MDM9206, MDM9607, MDM9640, MDM9650, MSM8996AU, QCS405, QCS605, SD 425, SD 427, SD 430, SD 435, SD 450, SD 625, SD 636, SD 675, SD 712 / SD 710 / SD 670, SD 730, SD 820A, SD 835, SD 845 / SD 850, SD 855, SDM630, SDM660, SDX20, SDX24
CVE-2018-6339 (v3: 9.8) 14 Jun 2019
When receiving calls using WhatsApp on Android, a stack allocation failed to properly account for the amount of data being passed in. An off-by-one error meant that data was written beyond the allocated space on the stack. This issue affects WhatsApp for Android starting in version 2.18.180 and was fixed in version 2.18.295. It also affects WhatsApp Business for Android starting in version v2.18.103 and was fixed in version v2.18.150.
CVE-2018-6349 (v3: 9.8) 14 Jun 2019
When receiving calls using WhatsApp for Android, a missing size check when parsing a sender-provided packet allowed for a stack-based overflow. This issue affects WhatsApp for Android prior to 2.18.248 and WhatsApp Business for Android prior to 2.18.132.
CVE-2018-3583 (v3: 7.8) 14 Jun 2019
A buffer overflow can occur while processing an extscan hotlist event in Snapdragon Auto, Snapdragon Consumer Electronics Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Wearables in MDM9206, MDM9607, MDM9640, MDM9650, MSM8909W, MSM8996AU, QCA9379, QCS605, SD 625, SD 636, SD 820, SD 820A, SD 835, SD 855, SDA660, SDM630, SDM660, SDX20
CVE-2018-13907 (v3: 5.3) 14 Jun 2019
While deserializing any key blob during key operations, buffer overflow could occur, exposing partial key information if any key operations are invoked in Snapdragon Auto, Snapdragon Compute, Snapdragon Connectivity, Snapdragon Consumer Electronics Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon IoT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wearables, Snapdragon Wired Infrastructure and Networking in IPQ4019, IPQ8074, MDM9150, MDM9206, MDM9607, MDM9635M, MDM9640, MDM9650, MDM9655, MSM8909W, MSM8996AU, QCA8081, QCS405, QCS605, Qualcomm 215, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 427, SD 430, SD 435, SD 439 / SD 429, SD 450, SD 615/16/SD 415, SD 625, SD 632, SD 636, SD 650/52, SD 712 / SD 710 / SD 670, SD 820, SD 820A, SD 835, SD 845 / SD 850, SD 855, SD 8CX, SDA660, SDM439, SDM630, SDM660, SDX20, Snapdragon_High_Med_2016, SXR1130
CVE-2018-10693 (v3: 8.8) 7 Jun 2019
An issue was discovered on Moxa AWK-3121 1.14 devices. It provides ping functionality so that an administrator can execute ICMP calls to check if the network is working correctly. However, the same functionality allows an attacker to execute commands on the device. The POST parameter "srvName" is susceptible to a buffer overflow. By crafting a packet that contains a string of 516 characters, it is possible for an attacker to execute the attack.
CVE-2018-10695 (v3: 8.8) 7 Jun 2019
An issue was discovered on Moxa AWK-3121 1.14 devices. It provides alert functionality so that an administrator can send emails to his/her account when there are changes to the device's network. However, the same functionality allows an attacker to execute commands on the device. The POST parameters "to1,to2,to3,to4" are all susceptible to buffer overflow. By crafting a packet that contains a string of 678 characters, it is possible for an attacker to execute the attack.
CVE-2018-10701 (v3: 8.8) 7 Jun 2019
An issue was discovered on Moxa AWK-3121 1.14 devices. It provides functionality so that an administrator can run scripts on the device to troubleshoot any issues. However, the same functionality allows an attacker to execute commands on the device. The POST parameter "iw_filename" is susceptible to buffer overflow. By crafting a packet that contains a string of 162 characters, it is possible for an attacker to execute the attack.
CVE-2018-10703 (v3: 8.8) 7 Jun 2019
An issue was discovered on Moxa AWK-3121 1.14 devices. It provides functionality so that an administrator can run scripts on the device to troubleshoot any issues. However, the same functionality allows an attacker to execute commands on the device. The POST parameter "iw_serverip" is susceptible to buffer overflow. By crafting a packet that contains a string of 480 characters, it is possible for an attacker to execute the attack.
CVE-2018-19800 (v3: 9.8) 7 Jun 2019
aubio v0.4.0 to v0.4.8 has a Buffer Overflow in new_aubio_tempo.
CVE-2018-13381 (v3: 7.5) 4 Jun 2019
A buffer overflow vulnerability in Fortinet FortiOS 6.0.0 to 6.0.4, 5.6.0 to 5.6.7, 5.4 and below versions under SSL VPN web portal allows a non-authenticated attacker to perform a Denial-of-service attack via special craft message payloads.
CVE-2018-20815 (v3: 9.8) 31 May 2019
In QEMU 3.1.0, load_device_tree in device_tree.c calls the deprecated load_image function, which has a buffer overflow risk.
CVE-2018-19978 (v3: 8) 29 May 2019
A buffer overflow vulnerability in the DHCP and PPPOE configuration interface of the Auerswald COMfort 1200 IP phone 3.4.4.1-10589 allows a remote attacker (authenticated as simple user in the same network as the device) to trigger remote code execution via a POST request (ManufacturerName parameter) to the web server on the device. The web server is running with root privileges and the injected code will also run with root privileges.
CVE-2018-13383 (v3: 6.5) 29 May 2019
A heap buffer overflow in Fortinet FortiOS 6.0.0 to 6.0.4, 5.6.0 to 5.6.10, 5.4.0 to 5.4.12, 5.2.14 and below in the SSL VPN web portal may cause the SSL VPN web service termination for logged in users due to a failure to properly handle javascript href data when proxying webpages.
CVE-2018-11928 (v3: 7.8) 24 May 2019
Lack of check on length parameter may cause buffer overflow while processing WMI commands in Snapdragon Auto, Snapdragon Compute, Snapdragon Connectivity, Snapdragon Consumer Electronics Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wired Infrastructure and Networking in IPQ8074, MDM9206, MDM9607, MDM9640, MDM9650, MSM8996AU, QCA6174A, QCA6564, QCA6574, QCA6574AU, QCA6584, QCA6584AU, QCA8081, QCA9377, QCA9379, QCA9886, QCS605, SD 210/SD 212/SD 205, SD 425, SD 600, SD 625, SD 636, SD 675, SD 712 / SD 710 / SD 670, SD 820, SD 820A, SD 835, SD 845 / SD 850, SD 855, SD 8CX, SDA660, SDM630, SDM660, SDX20, SDX24, SM7150, SXR1130
CVE-2018-11940 (v3: 9.8) 24 May 2019
Lack of check in length before using memcpy in WLAN function can lead to OOB access in Snapdragon Auto, Snapdragon Compute, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile in MDM9150, MDM9206, MDM9607, MDM9640, MDM9650, MSM8996AU, QCS605, SD 625, SD 636, SD 712 / SD 710 / SD 670, SD 820, SD 820A, SD 835, SD 845 / SD 850, SD 855, SDA660, SDM630, SDM660, SDX20, SDX24, SXR1130
CVE-2018-12012 (v3: 7.8) 24 May 2019
While updating blacklisting region shared buffered memory region is not validated against newly updated black list, causing boot-up to be compromised in Snapdragon Auto, Snapdragon Compute, Snapdragon Consumer Electronics Connectivity, Snapdragon Consumer IOT, Snapdragon Industrial IOT, Snapdragon Mobile, Snapdragon Voice & Music, Snapdragon Wearables in MDM9206, MDM9607, MDM9650, MDM9655, QCS605, SD 210/SD 212/SD 205, SD 410/12, SD 615/16/SD 415, SD 712 / SD 710 / SD 670, SD 835, SD 845 / SD 850, SD 8CX, SXR1130
CVE-2018-7846 (v3: 9.8) 22 May 2019
A CWE-501: Trust Boundary Violation vulnerability on connection to the Controller exists in all versions of the Modicon M580, Modicon M340, Modicon Quantum and Modicon Premium which could cause unauthorized access by conducting a brute force attack on Modbus protocol to the controller.
CVE-2018-7851 (v3: 6.5) 22 May 2019
CWE-119: Buffer errors vulnerability exists in Modicon M580 with firmware prior to V2.50, Modicon M340 with firmware prior to V3.01, BMxCRA312xx with firmware prior to V2.40, All firmware versions of Modicon Premium and 140CRA312xxx when sending a specially crafted Modbus packet, which could cause a denial of service to the device that would force a restart to restore availability.
CVE-2018-12886 (v3: 8.1) 22 May 2019
stack_protect_prologue in cfgexpand.c and stack_protect_epilogue in function.c in GNU Compiler Collection (GCC) 4.1 through 8 (under certain circumstances) generate instruction sequences when targeting ARM targets that spill the address of the stack protector guard, which allows an attacker to bypass the protection of -fstack-protector, -fstack-protector-all, -fstack-protector-strong, and -fstack-protector-explicit against stack overflow by controlling what the stack canary is compared against.

2017

CVE-2017-20144 24 Jul 2022
A vulnerability has been found in Anvsoft PDFMate PDF Converter Pro 1.7.5.0 and classified as critical. The manipulation leads to memory corruption. The attack can be launched remotely. The exploit has been disclosed to the public and may be used.
CVE-2017-18662 (v3: 7.5) 7 Apr 2020
An issue was discovered on Samsung mobile devices with M(6.0) and N(7.x) software. Data outside of the rkp log buffer boundary is read, causing an information leak. The Samsung ID is SVE-2017-9109 (July 2017).
CVE-2017-18696 (v3: 9.8) 7 Apr 2020
An issue was discovered on Samsung mobile devices with M(6.0) and N(7.0) (Exynos7420, Exynos8890, or MSM8996 chipsets) software. RKP allows memory corruption. The Samsung ID is SVE-2016-7897 (January 2017).
CVE-2017-5332 (v3: 7.8) 4 Nov 2019
The extract_group_icon_cursor_resource in wrestool/extract.c in icoutils before 0.31.1 can access unallocated memory, which allows local users to cause a denial of service (process crash) and execute arbitrary code via a crafted executable.
CVE-2017-5731 (v3: 7.8) 28 Oct 2019
Bounds checking in Tianocompress before November 7, 2017 may allow an authenticated user to potentially enable an escalation of privilege via local access.
CVE-2017-14742 (v3: 9.8) 25 Oct 2019
Buffer overflow in LabF nfsAxe FTP client 3.7 allows an attacker to execute code remotely.
CVE-2017-14202 (v3: 7.8) 29 Aug 2019
Improper Restriction of Operations within the Bounds of a Memory Buffer vulnerability in the shell component of Zephyr allows a serial or telnet connected user to cause a crash, possibly with arbitrary code execution. This issue affects: Zephyr shell versions prior to 1.14.0 on all.
CVE-2017-18379 (v3: 9.8) 27 Jul 2019
In the Linux kernel before 4.14, an out of boundary access happened in drivers/nvme/target/fc.c.
CVE-2017-13719 (v3: 9.8) 3 Jul 2019
The Amcrest IPM-721S Amcrest_IPC-AWXX_Eng_N_V2.420.AC00.17.R.20170322 allows HTTP requests that permit enabling various functionalities of the camera by using HTTP APIs, instead of the web management interface that is provided by the application. This HTTP API receives the credentials as base64 encoded in the Authorization HTTP header. However, a missing length check in the code allows an attacker to send a string of 1024 characters in the password field, and allows an attacker to exploit a memory corruption issue. This can allow an attacker to circumvent the account protection mechanism and brute force the credentials. If the firmware version Amcrest_IPC-AWXX_Eng_N_V2.420.AC00.17.R.20170322 is dissected using the binwalk tool, one obtains a _user-x.squashfs.img.extracted archive which contains the filesystem set up on the device that has many of the binaries in the /usr folder. The binary "sonia" is the one that has the vulnerable function that performs the credential check in the binary for the HTTP API specification. If we open this binary in IDA Pro we will notice that this follows an ARM little-endian format. The function at address 00415364 in IDA Pro starts the HTTP authentication process. This function calls another function at sub_ 0042CCA0 at address 0041549C. This function performs a strchr operation after base64 decoding the credentials, and stores the result on the stack, which results in a stack-based buffer overflow.
CVE-2017-8412 (v3: 8.8) 2 Jul 2019
An issue was discovered on D-Link DCS-1100 and DCS-1130 devices. The device has a custom binary called mp4ts under the /var/www/video folder. It seems that this binary dumps the HTTP VERB in the system logs. As a part of doing that it retrieves the HTTP VERB sent by the user and uses a vulnerable sprintf function at address 0x0000C3D4 in the function sub_C210 to copy the value into a string and then into a log file. Since there is no bounds check being performed on the environment variable at address 0x0000C360 this results in a stack overflow and overwrites the PC register allowing an attacker to execute buffer overflow or even a command injection attack.
CVE-2017-8416 (v3: 8.8) 2 Jul 2019
An issue was discovered on D-Link DCS-1100 and DCS-1130 devices. The device runs a custom daemon on UDP port 5978 which is called "dldps2121" and listens for broadcast packets sent on 255.255.255.255. This daemon handles custom D-Link UDP based protocol that allows D-Link mobile applications and desktop applications to discover D-Link devices on the local network. The binary processes the received UDP packets sent from any device in "main" function. One path in the function traverses towards a block of code that processing of packets which does an unbounded copy operation which allows to overflow the buffer. The custom protocol created by Dlink follows the following pattern: Packetlen, Type of packet; M=MAC address of device or broadcast; D=Device Type;C=base64 encoded command string;test=1111 We can see at address function starting at address 0x0000DBF8 handles the entire UDP packet and performs an insecure copy using strcpy function at address 0x0000DC88. This results in overflowing the stack pointer after 1060 characters and thus allows to control the PC register and results in code execution. The same form of communication can be initiated by any process including an attacker process on the mobile phone or the desktop and this allows a third-party application on the device to execute commands on the device without any authentication by sending just 1 UDP packet with custom base64 encoding.
CVE-2017-8410 (v3: 9.8) 2 Jul 2019
An issue was discovered on D-Link DCS-1100 and DCS-1130 devices. The binary rtspd in /sbin folder of the device handles all the rtsp connections received by the device. It seems that the binary performs a memcpy operation at address 0x00011E34 with the value sent in the "Authorization: Basic" RTSP header and stores it on the stack. The number of bytes to be copied are calculated based on the length of the string sent in the RTSP header by the client. As a result, memcpy copies more data then it can hold on stack and this results in corrupting the registers for the caller function sub_F6CC which results in memory corruption. The severity of this attack is enlarged by the fact that the same value is then copied on the stack in the function 0x00011378 and this allows to overflow the buffer allocated and thus control the PC register which will result in arbitrary code execution on the device.
CVE-2017-8414 (v3: 7.8) 2 Jul 2019
An issue was discovered on D-Link DCS-1100 and DCS-1130 devices. The binary orthrus in /sbin folder of the device handles all the UPnP connections received by the device. It seems that the binary performs a sprintf operation at address 0x0000A3E4 with the value in the command line parameter "-f" and stores it on the stack. Since there is no length check, this results in corrupting the registers for the function sub_A098 which results in memory corruption.
CVE-2017-8329 (v3: 6.4) 18 Jun 2019
An issue was discovered on Securifi Almond, Almond+, and Almond 2015 devices with firmware AL-R096. The device provides a user with the capability of setting a name for the wireless network. These values are stored by the device in NVRAM (Non-volatile RAM). It seems that the POST parameters passed in this request to set up names on the device do not have a string length check on them. This allows an attacker to send a large payload in the "mssid_1" POST parameter. The device also allows a user to view the name of the Wifi Network set by the user. While processing this request, the device calls a function at address 0x00412CE4 (routerSummary) in the binary "webServer" located in Almond folder, which retrieves the value set earlier by "mssid_1" parameter as SSID2 and this value then results in overflowing the stack set up for this function and allows an attacker to control $ra register value on the stack which allows an attacker to control the device by executing a payload of an attacker's choice. If the firmware version AL-R096 is dissected using binwalk tool, we obtain a cpio-root archive which contains the filesystem set up on the device that contains all the binaries. The binary "goahead" is the one that has the vulnerable function that receives the values sent by the POST request. If we open this binary in IDA-pro we will notice that this follows a MIPS little endian format. The function sub_00420F38 in IDA pro is identified to be receiving the values sent in the POST parameter "mssid_1" at address 0x0042BA00 and then sets in the NVRAM at address 0x0042C314. The value is later retrieved in the function at address 0x00412EAC and this results in overflowing the buffer as the function copies the value directly on the stack.
CVE-2017-8335 (v3: 8) 18 Jun 2019
An issue was discovered on Securifi Almond, Almond+, and Almond 2015 devices with firmware AL-R096. The device provides a user with the capability of setting name for wireless network. These values are stored by the device in NVRAM (Non-volatile RAM). It seems that the POST parameters passed in this request to set up names on the device do not have a string length check on them. This allows an attacker to send a large payload in the "mssid_1" POST parameter. The device also allows a user to view the name of the Wifi Network set by the user. While processing this request, the device calls a function named "getCfgToHTML" at address 0x004268A8 which retrieves the value set earlier by "mssid_1" parameter as SSID2 and this value then results in overflowing the stack set up for this function and allows an attacker to control $ra register value on the stack which allows an attacker to control the device by executing a payload of an attacker's choice. If the firmware version AL-R096 is dissected using binwalk tool, we obtain a cpio-root archive which contains the filesystem set up on the device that contains all the binaries. The binary "goahead" is the one that has the vulnerable function that recieves the values sent by the POST request. If we open this binary in IDA-pro we will notice that this follows a MIPS little endian format. The function sub_00420F38 in IDA pro is identified to be receiving the values sent in the POST parameter "mssid_1" at address 0x0042BA00 and then sets in the NVRAM at address 0x0042C314. The value is later retrieved in the function "getCfgToHTML" at address 0x00426924 and this results in overflowing the buffer due to "strcat" function that is utilized by this function.
CVE-2017-8336 (v3: 8.8) 18 Jun 2019
An issue was discovered on Securifi Almond, Almond+, and Almond 2015 devices with firmware AL-R096. The device provides a user with the capability of adding new routes to the device. It seems that the POST parameters passed in this request to set up routes on the device can be set in such a way that would result in overflowing the stack set up and allow an attacker to control the $ra register stored on the stack. If the firmware version AL-R096 is dissected using binwalk tool, we obtain a cpio-root archive which contains the filesystem set up on the device that contains all the binaries. The binary "goahead" is the one that has the vulnerable function that recieves the values sent by the POST request. If we open this binary in IDA-pro we will notice that this follows a MIPS little endian format. The function sub_00420F38 in IDA pro is identified to be receiving the values sent in the POST request. The POST parameter "gateway" allows to overflow the stack and control the $ra register after 1546 characters. The value from this post parameter is then copied on the stack at address 0x00421348 as shown below. This allows an attacker to provide the payload of his/her choice and finally take control of the device.
CVE-2017-10720 (v3: 7.8) 17 Jun 2019
Recently it was discovered as a part of the research on IoT devices in the most recent firmware for Shekar Endoscope that the desktop application used to connect to the device suffers from a stack overflow if more than 26 characters are passed to it as the Wi-Fi name. This application is installed on the device and an attacker who can provide the right payload can execute code on the user's system directly. Any breach of this system can allow an attacker to get access to all the data that the user has access too. The application uses a dynamic link library(DLL) called "avilib.dll" which is used by the application to send binary packets to the device that allow to control the device. One such action that the DLL provides is change password in the function "sendchangename" which allows a user to change the Wi-Fi name on the device. This function calls a sub function "sub_75876EA0" at address 0x758784F8. The function determines which action to execute based on the parameters sent to it. The "sendchangename" passes the datastring as the second argument which is the name we enter in the textbox and integer 1 as first argument. The rest of the 3 arguments are set to 0. The function "sub_75876EA0" at address 0x75876F19 uses the first argument received and to determine which block to jump to. Since the argument passed is 1, it jumps to 0x75876F20 and proceeds from there to address 0x75876F56 which calculates the length of the data string passed as the first parameter. This length and the first argument are then passed to the address 0x75877001 which calls the memmove function which uses a stack address as the destination where the password typed by us is passed as the source and length calculated above is passed as the number of bytes to copy which leads to a stack overflow.
CVE-2017-10722 (v3: 7.8) 17 Jun 2019
Recently it was discovered as a part of the research on IoT devices in the most recent firmware for Shekar Endoscope that the desktop application used to connect to the device suffers from a stack overflow if more than 26 characters are passed to it as the Wi-Fi password. This application is installed on the device and an attacker who can provide the right payload can execute code on the user's system directly. Any breach of this system can allow an attacker to get access to all the data that the user has access too. The application uses a dynamic link library(DLL) called "avilib.dll" which is used by the application to send binary packets to the device that allow to control the device. One such action that the DLL provides is change password in the function "sendchangepass" which allows a user to change the Wi-Fi password on the device. This function calls a sub function "sub_75876EA0" at address 0x7587857C. The function determines which action to execute based on the parameters sent to it. The "sendchangepass" passes the datastring as the second argument which is the password we enter in the textbox and integer 2 as first argument. The rest of the 3 arguments are set to 0. The function "sub_75876EA0" at address 0x75876F19 uses the first argument received and to determine which block to jump to. Since the argument passed is 2, it jumps to 0x7587718C and proceeds from there to address 0x758771C2 which calculates the length of the data string passed as the first parameter.This length and the first argument are then passed to the address 0x7587726F which calls a memmove function which uses a stack address as the destination where the password typed by us is passed as the source and length calculated above is passed as the number of bytes to copy which leads to a stack overflow.
CVE-2017-10723 (v3: 8.8) 17 Jun 2019
Recently it was discovered as a part of the research on IoT devices in the most recent firmware for Shekar Endoscope that an attacker connected to the device Wi-Fi SSID can exploit a memory corruption issue and execute remote code on the device. This device acts as an Endoscope camera that allows its users to use it in various industrial systems and settings, car garages, and also in some cases in the medical clinics to get access to areas that are difficult for a human being to reach. Any breach of this system can allow an attacker to get access to video feed and pictures viewed by that user and might allow them to get a foot hold in air gapped networks especially in case of nation critical infrastructure/industries. The firmware contains binary uvc_stream that is the UDP daemon which is responsible for handling all the UDP requests that the device receives. The client application sends a UDP request to change the Wi-Fi name which contains the following format: "SETCMD0001+0001+[2 byte length of wifiname]+[Wifiname]. This request is handled by "control_Dev_thread" function which at address "0x00409AE0" compares the incoming request and determines if the 10th byte is 01 and if it is then it redirects to 0x0040A74C which calls the function "setwifiname". The function "setwifiname" uses a memcpy function but uses the length of the payload obtained by using strlen function as the third parameter which is the number of bytes to copy and this allows an attacker to overflow the function and control the $PC value.
CVE-2017-10724 (v3: 8.8) 17 Jun 2019
Recently it was discovered as a part of the research on IoT devices in the most recent firmware for Shekar Endoscope that an attacker connected to the device Wi-Fi SSID can exploit a memory corruption issue and execute remote code on the device. This device acts as an Endoscope camera that allows its users to use it in various industrial systems and settings, car garages, and also in some cases in the medical clinics to get access to areas that are difficult for a human being to reach. Any breach of this system can allow an attacker to get access to video feed and pictures viewed by that user and might allow them to get a foot hold in air gapped networks especially in case of nation critical infrastructure/industries. The firmware contains binary uvc_stream that is the UDP daemon which is responsible for handling all the UDP requests that the device receives. The client application sends a UDP request to change the Wi-Fi name which contains the following format: "SETCMD0001+0002+[2 byte length of wifipassword]+[Wifipassword]. This request is handled by "control_Dev_thread" function which at address "0x00409AE4" compares the incoming request and determines if the 10th byte is 02 and if it is then it redirects to 0x0040A7D8, which calls the function "setwifipassword". The function "setwifipassword" uses a memcpy function but uses the length of the payload obtained by using strlen function as the third parameter which is the number of bytes to copy and this allows an attacker to overflow the function and control the $PC value.
CVE-2017-9391 (v3: 8.8) 17 Jun 2019
An issue was discovered on Vera VeraEdge 1.7.19 and Veralite 1.7.481 devices. The device provides UPnP services that are available on port 3480 and can also be accessed via port 80 using the url "/port_3480". It seems that the UPnP services provide "request_image" as one of the service actions for a normal user to retrieve an image from a camera that is controlled by the controller. It seems that the "URL" parameter passed in the query string is not sanitized and is stored on the stack which allows an attacker to overflow the buffer. The function "LU::Generic_IP_Camera_Manager::REQ_Image" is activated when the lu_request_image is passed as the "id" parameter in query string. This function then calls "LU::Generic_IP_Camera_Manager::GetUrlFromArguments" and passes a "pointer" to the function where it will be allowed to store the value from the URL parameter. This pointer is passed as the second parameter $a2 to the function "LU::Generic_IP_Camera_Manager::GetUrlFromArguments". However, neither the callee or the caller in this case performs a simple length check and as a result an attacker who is able to send more than 1336 characters can easily overflow the values stored on the stack including the $RA value and thus execute code on the device.
CVE-2017-9392 (v3: 8.8) 17 Jun 2019
An issue was discovered on Vera VeraEdge 1.7.19 and Veralite 1.7.481 devices. The device provides UPnP services that are available on port 3480 and can also be accessed via port 80 using the url "/port_3480". It seems that the UPnP services provide "request_image" as one of the service actions for a normal user to retrieve an image from a camera that is controlled by the controller. It seems that the "res" (resolution) parameter passed in the query string is not sanitized and is stored on the stack which allows an attacker to overflow the buffer. The function "LU::Generic_IP_Camera_Manager::REQ_Image" is activated when the lu_request_image is passed as the "id" parameter in the query string. This function then calls "LU::Generic_IP_Camera_Manager::GetUrlFromArguments". This function retrieves all the parameters passed in the query string including "res" and then uses the value passed in it to fill up buffer using the sprintf function. However, the function in this case lacks a simple length check and as a result an attacker who is able to send more than 184 characters can easily overflow the values stored on the stack including the $RA value and thus execute code on the device.
CVE-2017-14854 (v3: 9.8) 3 Jun 2019
A stack buffer overflow exists in one of the Orpak SiteOmat CGI components, allowing for remote code execution. The vulnerability affects all versions prior to 2017-09-25.
CVE-2017-7773 (v3: 8.8) 15 Apr 2019
Heap-based Buffer Overflow write in Graphite2 library in Firefox before 54 in lz4::decompress src/Decompressor.
CVE-2017-7777 (v3: 8.8) 15 Apr 2019
Use of uninitialized memory in Graphite2 library in Firefox before 54 in graphite2::GlyphCache::Loader::read_glyph function.
CVE-2017-7772 (v3: 8.8) 12 Apr 2019
Heap-based Buffer Overflow in Graphite2 library in Firefox before 54 in lz4::decompress function.
CVE-2017-14199 (v3: 9.8) 12 Apr 2019
A buffer overflow has been found in the Zephyr Project's getaddrinfo() implementation in 1.9.0 and 1.10.0.
CVE-2017-16253 (v3: 6.5) 21 Mar 2019
An exploitable buffer overflow vulnerability exists in the PubNub message handler Insteon Hub 2245-222 - Firmware version 1012 for the cc channel of Insteon Hub running firmware version 1012. Specially crafted commands sent through the PubNub service can cause a stack-based buffer overflow overwriting arbitrary data. An attacker can send an authenticated HTTP request At 0x9d014dd8 the value for the id key is copied using strcpy to the buffer at $sp+0x290. This buffer is 32 bytes large, sending anything longer will cause a buffer overflow.
CVE-2017-16254 (v3: 6.5) 21 Mar 2019
An exploitable buffer overflow vulnerability exists in the PubNub message handler Insteon Hub 2245-222 - Firmware version 1012. Specially crafted commands sent through the PubNub service can cause a stack-based buffer overflow overwriting arbitrary data. An attacker can send an authenticated HTTP request at 0x9d014e4c the value for the flg key is copied using strcpy to the buffer at $sp+0x270. This buffer is 16 bytes large, sending anything longer will cause a buffer overflow.
CVE-2017-16255 (v3: 6.5) 21 Mar 2019
An exploitable buffer overflow vulnerability exists in the PubNub message handler Insteon Hub 2245-222 - Firmware version 1012. Specially crafted commands sent through the PubNub service can cause a stack-based buffer overflow overwriting arbitrary data. An attacker can send an authenticated HTTP request at At 0x9d014e84 the value for the cmd1 key is copied using strcpy to the buffer at $sp+0x280. This buffer is 16 bytes large.
CVE-2017-16231 (v3: 5.5) 21 Mar 2019
** DISPUTED ** In PCRE 8.41, after compiling, a pcretest load test PoC produces a crash overflow in the function match() in pcre_exec.c because of a self-recursive call. NOTE: third parties dispute the relevance of this report, noting that there are options that can be used to limit the amount of stack that is used.
CVE-2017-12447 (v3: 7.8) 7 Mar 2019
GdkPixBuf (aka gdk-pixbuf), possibly 2.32.2, as used by GNOME Nautilus 3.14.3 on Ubuntu 16.04, allows attackers to cause a denial of service (stack corruption) or possibly have unspecified other impact via a crafted file folder.
CVE-2017-18329 (v3: 7.8) 3 Jan 2019
Possible Buffer overflow when transmitting an RTP packet in snapdragon automobile and snapdragon wear in versions MDM9615, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, MSM8996AU, SD 210/SD 212/SD 205, SD 425, SD 427, SD 430, SD 435, SD 450, SD 615/16/SD 415, SD 625, SD 636, SD 650/52, SD 712 / SD 710 / SD 670, SD 810, SD 820, SD 835, SD 845 / SD 850, SDA660, SDM630, SDM660, Snapdragon_High_Med_2016, SXR1130
CVE-2017-16909 (v3: 8.8) 7 Dec 2018
An error related to the "LibRaw::panasonic_load_raw()" function (dcraw_common.cpp) in LibRaw versions prior to 0.18.6 can be exploited to cause a heap-based buffer overflow and subsequently cause a crash via a specially crafted TIFF image.
CVE-2017-14888 (v3: 7.8) 7 Dec 2018
In all android releases(Android for MSM, Firefox OS for MSM, QRD Android) from CAF using the linux kernel, Userspace can pass IEs to the host driver and if multiple append commands are received, then the integer variable that stores the length can overflow and the subsequent copy of the IE data may potentially lead to a heap buffer overflow.
CVE-2017-18124 (v3: 7.8) 26 Oct 2018
During secure boot, addition is performed on uint8 ptrs which led to overflow issue in Small Cell SoC, Snapdragon Automobile, Snapdragon Mobile, Snapdragon Wear in version FSM9055, IPQ4019, MDM9206, MDM9607, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, MSM8996AU, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 810, SD 820, SD 820A, SD 835, SD 845, SD 850, SDA660, SDX20
CVE-2017-18171 (v3: 8.8) 23 Oct 2018
Improper input validation for GATT data packet received in Bluetooth Controller function can lead to possible memory corruption in Snapdragon Mobile in version QCA9379, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 427, SD 430, SD 435, SD 450, SD 615/16/SD 415, SD 625, SD 650/52, SD 820, SD 835, SD 845, SD 850, SDM630, SDM636, SDM660, SDM710, Snapdragon_High_Med_2016.
CVE-2017-18283 (v3: 6.5) 23 Oct 2018
Possible memory corruption when Read Val Blob Req is received with invalid parameters in Snapdragon Mobile in version QCA9379, SD 210/SD 212/SD 205, SD 625, SD 835, SD 845, SD 850, SDA660.
CVE-2017-18295 (v3: 7.8) 23 Oct 2018
Possible buffer overflow if input is not null terminated in DSP Service module in Snapdragon Automobile, Snapdragon Mobile, Snapdragon Wear in version MDM9206, MDM9607, MDM9650, MSM8909W, MSM8996AU, SD 210/SD 212/SD 205, SD 450, SD 615/16/SD 415, SD 625, SD 650/52, SD 820, SD 820A, SD 835, SDX20.
CVE-2017-18303 (v3: 7.8) 23 Oct 2018
While processing the sensors registry configuration file, if inputs are not validated a buffer overflow will occur in Snapdragon Automobile, Snapdragon Mobile, Snapdragon Wear in version MMDM9206, MDM9607, MDM9650, MSM8909W, MSM8996AU, SD 210/SD 212/SD 205, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 810, SD 820, SD 820A, SD 835, SDA660, SDX20.
CVE-2017-7908 (v3: 7.6) 2 Oct 2018
A heap-based buffer overflow exists in the third-party product Gigasoft, v5 and prior, included in GE Communicator 3.15 and prior. A malicious HTML file that loads the ActiveX controls can trigger the vulnerability via unchecked function calls.
CVE-2017-2876 (v3: 7.5) 19 Sep 2018
An exploitable buffer overflow vulnerability exists in the Multi-Camera interface used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. A specially crafted request on port 10000 can cause a buffer overflow resulting in overwriting arbitrary data.
CVE-2017-2855 (v3: 8.1) 19 Sep 2018
An exploitable buffer overflow vulnerability exists in the DDNS client used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. On devices with DDNS enabled, an attacker who is able to intercept HTTP connections will be able to fully compromise the device by creating a rogue HTTP server.
CVE-2017-2875 (v3: 7.5) 19 Sep 2018
An exploitable buffer overflow vulnerability exists in the Multi-Camera interface used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. A specially crafted request on port 10000 can cause a buffer overflow resulting in overwriting arbitrary data.
CVE-2017-2878 (v3: 7.5) 19 Sep 2018
An exploitable buffer overflow vulnerability exists in the web management interface used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. A specially crafted HTTP request can cause a buffer overflow resulting in overwriting arbitrary data. An attacker can simply send an HTTP request to the device to trigger this vulnerability.
CVE-2017-2879 (v3: 5.3) 19 Sep 2018
An exploitable buffer overflow vulnerability exists in the UPnP implementation used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. A specially crafted UPnP discovery response can cause a buffer overflow resulting in overwriting arbitrary data. An attacker needs to be in the same subnetwork and reply to a discovery message to trigger this vulnerability.
CVE-2017-2854 (v3: 8.1) 17 Sep 2018
An exploitable buffer overflow vulnerability exists in the DDNS client used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. On devices with DDNS enabled, an attacker who is able to intercept HTTP connections will be able to fully compromise the device by creating a rogue HTTP server.
CVE-2017-2856 (v3: 8.1) 17 Sep 2018
An exploitable buffer overflow vulnerability exists in the DDNS client used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. On devices with DDNS enabled, an attacker who is able to intercept HTTP connections will be able to fully compromise the device by creating a rogue HTTP server.
CVE-2017-2857 (v3: 8.1) 17 Sep 2018
An exploitable buffer overflow vulnerability exists in the DDNS client used by the Foscam C1 Indoor HD Camera running application firmware 2.52.2.43. On devices with DDNS enabled, an attacker who is able to intercept HTTP connections will be able to fully compromise the device by creating a rogue HTTP server.
CVE-2017-1083 (v3: 7.5) 12 Sep 2018
In FreeBSD before 11.2-RELEASE, a stack guard-page is available but is disabled by default. This results in the possibility a poorly written process could be cause a stack overflow.

2016

CVE-2016-11035 (v3: 5.5) 7 Apr 2020
An issue was discovered on Samsung mobile devices with software through 2016-05-27 (Exynos AP chipsets). A local graphics user can cause a Kernel Crash via the fb0(DECON) frame buffer interface. The Samsung ID is SVE-2016-7011 (October 2016).
CVE-2016-11038 (v3: 9.8) 7 Apr 2020
An issue was discovered on Samsung mobile devices with software through 2016-04-05 (incorporating the Samsung Professional Audio SDK). The Jack audio service doesn't implement access control for shared memory, leading to arbitrary code execution or privilege escalation. The Samsung ID is SVE-2016-5953 (July 2016).
CVE-2016-11045 (v3: 7.8) 7 Apr 2020
An issue was discovered on Samsung mobile devices with L(5.0/5.1) software. The Gallery library allow memory corruption via a malformed image. The Samsung ID is SVE-2016-5317 (May 2016).
CVE-2016-3182 (v3: 5.5) 20 Feb 2020
The color_esycc_to_rgb function in bin/common/color.c in OpenJPEG before 2.1.1 allows attackers to cause a denial of service (memory corruption) via a crafted jpeg 2000 file.
CVE-2016-10764 (v3: 9.8) 27 Jul 2019
In the Linux kernel before 4.9.6, there is an off by one in the drivers/mtd/spi-nor/cadence-quadspi.c cqspi_setup_flash() function. There are CQSPI_MAX_CHIPSELECT elements in the ->f_pdata array so the ">" should be ">=" instead.
CVE-2016-5800 (v3: 7.5) 21 Mar 2019
A malicious attacker can trigger a remote buffer overflow in the Communication Server in Fatek Automation PM Designer V3 Version 2.1.2.2, and Automation FV Designer Version 1.2.8.0.
CVE-2016-7576 (v3: 7.8) 11 Jan 2019
In iOS before 9.3.3, a memory corruption issue existed in the kernel. This issue was addressed through improved memory handling.
CVE-2016-2123 (v3: 8.8) 1 Nov 2018
A flaw was found in samba versions 4.0.0 to 4.5.2. The Samba routine ndr_pull_dnsp_name contains an integer wrap problem, leading to an attacker-controlled memory overwrite. ndr_pull_dnsp_name parses data from the Samba Active Directory ldb database. Any user who can write to the dnsRecord attribute over LDAP can trigger this memory corruption. By default, all authenticated LDAP users can write to the dnsRecord attribute on new DNS objects. This makes the defect a remote privilege escalation.
CVE-2016-4402 (v3: 9.8) 6 Aug 2018
A security vulnerability was identified in the Filter SDK component of HP KeyView earlier than v11.2. The vulnerability could be exploited remotely to allow code execution via buffer overflow.
CVE-2016-4403 (v3: 9.8) 6 Aug 2018
A security vulnerability was identified in the Filter SDK component of HP KeyView earlier than v11.2. The vulnerability could be exploited remotely to allow code execution via memory corruption.
CVE-2016-4404 (v3: 9.8) 6 Aug 2018
A security vulnerability was identified in the Filter SDK component of HP KeyView earlier than v11.2. The vulnerability could be exploited remotely to allow code execution via a memory allocation issue.
CVE-2016-8654 (v3: 7.8) 1 Aug 2018
A heap-buffer overflow vulnerability was found in QMFB code in JPC codec caused by buffer being allocated with too small size. jasper versions before 2.0.0 are affected.
CVE-2016-9581 (v3: 8.8) 1 Aug 2018
An infinite loop vulnerability in tiftoimage that results in heap buffer overflow in convert_32s_C1P1 was found in openjpeg 2.1.2.
CVE-2016-9597 (v3: 7.5) 30 Jul 2018
It was found that Red Hat JBoss Core Services erratum RHSA-2016:2957 for CVE-2016-3705 did not actually include the fix for the issue found in libxml2, making it vulnerable to a Denial of Service attack due to a Stack Overflow. This is a regression CVE for the same issue as CVE-2016-3705.
CVE-2016-9603 (v3: 9.9) 27 Jul 2018
A heap buffer overflow flaw was found in QEMU's Cirrus CLGD 54xx VGA emulator's VNC display driver support before 2.9; the issue could occur when a VNC client attempted to update its display after a VGA operation is performed by a guest. A privileged user/process inside a guest could use this flaw to crash the QEMU process or, potentially, execute arbitrary code on the host with privileges of the QEMU process.
CVE-2016-9577 (v3: 8.8) 27 Jul 2018
A vulnerability was discovered in SPICE before 0.13.90 in the server's protocol handling. An authenticated attacker could send crafted messages to the SPICE server causing a heap overflow leading to a crash or possible code execution.
CVE-2016-6559 (v3: 9.8) 13 Jul 2018
Improper bounds checking of the obuf variable in the link_ntoa() function in linkaddr.c of the BSD libc library may allow an attacker to read or write from memory. The full impact and severity depends on the method of exploit and how the library is used by applications. According to analysis by FreeBSD developers, it is very unlikely that applications exist that utilize link_ntoa() in an exploitable manner, and the CERT/CC is not aware of any proof of concept. A blog post describes the functionality of link_ntoa() and points out that none of the base utilities use this function in an exploitable manner. For more information, please see FreeBSD Security Advisory SA-16:37.
CVE-2016-6563 (v3: 9.8) 13 Jul 2018
Processing malformed SOAP messages when performing the HNAP Login action causes a buffer overflow in the stack in some D-Link DIR routers. The vulnerable XML fields within the SOAP body are: Action, Username, LoginPassword, and Captcha. The following products are affected: DIR-823, DIR-822, DIR-818L(W), DIR-895L, DIR-890L, DIR-885L, DIR-880L, DIR-868L, and DIR-850L.
CVE-2016-9893 (v3: 9.8) 11 Jun 2018
Memory safety bugs were reported in Thunderbird 45.5. Some of these bugs showed evidence of memory corruption and we presume that with enough effort that some of these could be exploited to run arbitrary code. This vulnerability affects Firefox < 50.1, Firefox ESR < 45.6, and Thunderbird < 45.6.
CVE-2016-9894 (v3: 7.5) 11 Jun 2018
A buffer overflow in SkiaGl caused when a GrGLBuffer is truncated during allocation. Later writers will overflow the buffer, resulting in a potentially exploitable crash. This vulnerability affects Firefox < 50.1.
CVE-2016-9897 (v3: 7.5) 11 Jun 2018
Memory corruption resulting in a potentially exploitable crash during WebGL functions using a vector constructor with a varying array within libGLES. This vulnerability affects Firefox < 50.1, Firefox ESR < 45.6, and Thunderbird < 45.6.
CVE-2016-9066 (v3: 7.5) 11 Jun 2018
A buffer overflow resulting in a potentially exploitable crash due to memory allocation issues when handling large amounts of incoming data. This vulnerability affects Thunderbird < 45.5, Firefox ESR < 45.5, and Firefox < 50.
CVE-2016-9080 (v3: 9.8) 11 Jun 2018
Memory safety bugs were reported in Firefox 50.0.2. Some of these bugs showed evidence of memory corruption and we presume that with enough effort that some of these could be exploited to run arbitrary code. This vulnerability affects Firefox < 50.1.
CVE-2016-5289 (v3: 9.8) 11 Jun 2018
Memory safety bugs were reported in Firefox 49. Some of these bugs showed evidence of memory corruption and we presume that with enough effort that some of these could be exploited to run arbitrary code. This vulnerability affects Firefox < 50.
CVE-2016-5290 (v3: 9.8) 11 Jun 2018
Memory safety bugs were reported in Firefox 49 and Firefox ESR 45.4. Some of these bugs showed evidence of memory corruption and we presume that with enough effort that some of these could be exploited to run arbitrary code. This vulnerability affects Thunderbird < 45.5, Firefox ESR < 45.5, and Firefox < 50.
CVE-2016-5296 (v3: 7.5) 11 Jun 2018
A heap-buffer-overflow in Cairo when processing SVG content caused by compiler optimization, resulting in a potentially exploitable crash. This vulnerability affects Thunderbird < 45.5, Firefox ESR < 45.5, and Firefox < 50.
CVE-2016-10518 (v3: 7.5) 31 May 2018
A vulnerability was found in the ping functionality of the ws module before 1.0.0 which allowed clients to allocate memory by sending a ping frame. The ping functionality by default responds with a pong frame and the previously given payload of the ping frame. This is exactly what you expect, but internally ws always transforms all data that we need to send to a Buffer instance and that is where the vulnerability existed. ws didn't do any checks for the type of data it was sending. With buffers in node when you allocate it when a number instead of a string it will allocate the amount of bytes.
CVE-2016-10523 (v3: 7.5) 31 May 2018
MQTT before 3.4.6 and 4.0.x before 4.0.5 allows specifically crafted MQTT packets to crash the application, making a DoS attack feasible with very little bandwidth.
CVE-2016-10721 (v3: 9.8) 3 May 2018
partclone.restore in Partclone 0.2.87 is prone to a heap-based buffer overflow vulnerability due to insufficient validation of the partclone image header. An attacker may be able to execute arbitrary code in the context of the user running the affected application.
CVE-2016-10722 (v3: 9.8) 3 May 2018
partclone.fat in Partclone before 0.2.88 is prone to a heap-based buffer overflow vulnerability due to insufficient validation of the FAT superblock, related to the mark_reserved_sectors function. An attacker may be able to execute arbitrary code in the context of the user running the affected application.
CVE-2016-8729 (v3: 7.8) 24 Apr 2018
An exploitable memory corruption vulnerability exists in the JBIG2 parser of Artifex MuPDF 1.9. A specially crafted PDF can cause a negative number to be passed to a memset resulting in memory corruption and potential code execution. An attacker can specially craft a PDF and send to the victim to trigger this vulnerability.
CVE-2016-8382 (v3: 8.8) 24 Apr 2018
An exploitable heap corruption vulnerability exists in the Doc_SetSummary functionality of AntennaHouse DMC HTMLFilter. A specially crafted doc file can cause a heap corruption resulting in arbitrary code execution. An attacker can send a malicious doc file to trigger this vulnerability.
CVE-2016-8383 (v3: 8.8) 24 Apr 2018
An exploitable heap corruption vulnerability exists in the Doc_GetFontTable functionality of AntennaHouse DMC HTMLFilter. A specially crafted doc file can cause a heap corruption resulting in arbitrary code execution. An attacker can send/provide malicious doc file to trigger this vulnerability.
CVE-2016-8384 (v3: 8.8) 24 Apr 2018
An exploitable heap corruption vulnerability exists in the DHFSummary functionality of AntennaHouse DMC HTMLFilter.
CVE-2016-9601 (v3: 5.5) 24 Apr 2018
ghostscript before version 9.21 is vulnerable to a heap based buffer overflow that was found in the ghostscript jbig2_decode_gray_scale_image function which is used to decode halftone segments in a JBIG2 image. A document (PostScript or PDF) with an embedded, specially crafted, jbig2 image could trigger a segmentation fault in ghostscript.
CVE-2016-9586 (v3: 8.1) 23 Apr 2018
curl before version 7.52.0 is vulnerable to a buffer overflow when doing a large floating point output in libcurl's implementation of the printf() functions. If there are any application that accepts a format string from the outside without necessary input filtering, it could allow remote attacks.
CVE-2016-10473 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 835, and SDX20, in a supplementary services function, a buffer overflow can occur.
CVE-2016-10474 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear IPQ4019, MDM9206, MDM9607, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, if the buffer length passed to the RIL interface is too large, the buffer size calculation may overflow, resulting in an undersize allocation for the buffer, and subsequently buffer overwrite.
CVE-2016-10476 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 835, and SDX20, missing array index checks on app index in function qcril_uim_clear_encrypted_pin results in accessing addresses outside the bounds of the buffer when app index is too large.
CVE-2016-10477 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile SD 210/SD 212/SD 205, SD 400, SD 430, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, and SD 820, while processing smart card requests, a buffer overflow can occur.
CVE-2016-10484 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear IPQ4019, MDM9206, MDM9607, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, and SDX20, if a RPMB listener is registered with a very small buffer size, the calculation of the maximum transfer size for read and write operations may underflow, resulting in buffer overflow.
CVE-2016-10485 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear IPQ4019, MDM9206, MDM9607, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 615/16/SD 415, SD 617, SD 650/52, SD 800, SD 808, SD 810, and SDX20, lack of proper bounds checking may lead to a buffer overflow.
CVE-2016-10486 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile and Snapdragon Mobile MDM9640, MDM9645, SD 210/SD 212/SD 205, SD 450, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, and SD 820A, PD failure reason string from user PD is used directly in root PD, so if the buffer parameter is non-NULL terminated in Diag F3 APIs, a buffer overread occurs.
CVE-2016-10493 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9607, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 425, SD 430, SD 450, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, SD 850, and SDX20, NPA routines on the rootPD that handle resource requests remoted over QDI may not validate pointers passed from user space which may result in guest OS memory corruption.
CVE-2016-10497 (v3: 7.5) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9615, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 835, SD 845, SD 850, and SDX20, improper CFG allocation can cause heap leak.
CVE-2016-10410 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9615, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 835, SD 845, SD 850, and SDX20, buffer overflow vulnerability in RTP during Volte call.
CVE-2016-10419 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile MDM9640, MDM9645, MDM9650, MDM9655, SD 450, SD 625, SD 650/52, SD 820, SD 835, SD 845, SD 850, and SDX20, when initializing scheduler object service request, an out of bounds access could occur due to uninitialized object number.
CVE-2016-10420 (v3: 5.5) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9650, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 835, and SDX20, while playing back a .flv clip which doesn't have an inbuilt seek table, a dynamic index table access is out of bounds and leads to crash.
CVE-2016-10425 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, and SD 835, if GPT listener response is passed a large buffer offset, a buffer overflow occurs.
CVE-2016-10426 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile and Snapdragon Mobile SD 410/12, SD 425, SD 430, SD 450, SD 617, SD 625, SD 650/52, SD 810, SD 820, and SD 820A, a buffer overflow can occur in SafeSwitch.

2015

CVE-2015-0565 (v3: 10) 25 Feb 2020
NaCl in 2015 allowed the CLFLUSH instruction, making rowhammer attacks possible.
CVE-2015-5290 (v3: 7.5) 26 Dec 2019
A Denial of Service vulnerability exists in ircd-ratbox 3.0.9 in the MONITOR Command Handler.
CVE-2015-3166 (v3: 9.8) 20 Nov 2019
The snprintf implementation in PostgreSQL before 9.0.20, 9.1.x before 9.1.16, 9.2.x before 9.2.11, 9.3.x before 9.3.7, and 9.4.x before 9.4.2 does not properly handle system-call errors, which allows attackers to obtain sensitive information or have other unspecified impact via unknown vectors, as demonstrated by an out-of-memory error.
CVE-2015-1006 (v3: 9.8) 10 May 2019
A vulnerable file in Opto 22 PAC Project Professional versions prior to R9.4006, PAC Project Basic versions prior to R9.4006, PAC Display Basic versions prior to R9.4f, PAC Display Professional versions prior to R9.4f, OptoOPCServer versions prior to R9.4c, and OptoDataLink version R9.4d and prior versions that were installed by PAC Project installer, versions prior to R9.4006, is susceptible to a heap-based buffer overflow condition that may allow remote code execution on the target system. Opto 22 suggests upgrading to the new product version as soon as possible.
CVE-2015-1007 (v3: 7.8) 25 Mar 2019
A specially crafted configuration file could be used to cause a stack-based buffer overflow condition in the OPCTest.exe, which may allow remote code execution on Opto 22 PAC Project Professional versions prior to R9.4008, PAC Project Basic versions prior to R9.4008, PAC Display Basic versions prior to R9.4g, PAC Display Professional versions prior to R9.4g, OptoOPCServer version R9.4c and prior that were installed by PAC Project installer, versions prior to R9.4008, and OptoDataLink version R9.4d and prior that were installed by PAC Project installer, versions prior to R9.4008. Opto 22 suggests upgrading to the new product version as soon as possible.
CVE-2015-6457 (v3: 8.8) 21 Mar 2019
Moxa SoftCMS 1.3 and prior is susceptible to a buffer overflow condition that may crash or allow remote code execution. Moxa released SoftCMS version 1.4 on June 1, 2015, to address the vulnerability.
CVE-2015-6458 (v3: 8.8) 21 Mar 2019
Moxa SoftCMS 1.3 and prior is susceptible to a buffer overflow condition that may crash or allow remote code execution. Moxa released SoftCMS version 1.4 on June 1, 2015, to address the vulnerability.
CVE-2015-9262 (v3: 9.8) 2 Aug 2018
_XcursorThemeInherits in library.c in libXcursor before 1.1.15 allows remote attackers to cause denial of service or potentially code execution via a one-byte heap overflow.
CVE-2015-9112 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile and Snapdragon Mobile MDM9625, SD 400, SD 800, SD 820, and SD 820A, lack of input validation in QSEE can cause potential buffer overflow.
CVE-2015-9118 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile MDM9615, MDM9625, MDM9635M, SD 400, SD 410/12, SD 615/16/SD 415, SD 800, SD 808, and SD 810, in ADSP's QDI Root-PD driver, untrusted arguments from User PD may cause integer overflow resulting in buffer overflow.
CVE-2015-9122 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9635M, MDM9640, MDM9645, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, and SD 835, possible buffer overflow if SIM card sends a response greater than 64KB of data for stream APDU command.
CVE-2015-9126 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9607, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 425, SD 430, SD 450, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 835, SD 845, SD 850, and SDX20, possible buffer overflow when processing 1X circuit service message.
CVE-2015-9128 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, and SD 835, lack of validation of the buffer size could lead to a buffer overread.
CVE-2015-9136 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9640, MDM9650, SD 210/SD 212/SD 205, SD 425, SD 430, SD 450, SD 600, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 835, and SDX20, in pre-auth request, Host driver uses FT IEs sent by the supplicant. A buffer overflow may occur if FT IEs sent by the supplicant are larger than the expected value.
CVE-2015-9138 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, Snapdragon Wear, and Small Cell SoC FSM9055, IPQ4019, MDM9206, MDM9607, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, SD 850, and SDX20, when an RSA encryption operation is called, the ce_util_to_unsigned_bin is invoked to convert the input buffer to unsigned binary. The ce_util_to_unsigned_bin function, instead of operating on the size of the unsigned character buffer that is passed, operates on the address - i.e. operates on "c" instead of "*c". Decrementing the address to check if it is less than zero means that the operation will always pass, since a pointer will never be less than zero, and may result in a buffer overflow.
CVE-2015-9141 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9635M, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 615/16/SD 415, SD 617, SD 800, SD 808, and SD 810, in HHO scenarios, during the ACQ procedure, there are possible instances where the search database is incorrectly updated resulting in memory corruption due to buffer overflow.
CVE-2015-9143 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear IPQ4019, MDM9206, MDM9607, MDM9615, MDM9625, MDM9640, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 600, SD 615/16/SD 415, and SDX20, when reading CDT from eMMC with a very large meta offset (>size of default CDT-array compiled in bootloader) for one of the CDBs, a buffer overflow occurs.
CVE-2015-9144 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9615, MDM9625, MDM9635M, MDM9640, MDM9645, MDM9650, MDM9655, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 835, SD 845, SD 850, and SDX20, while processing scheduling message information, a buffer overflow can occur.
CVE-2015-9153 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear IPQ4019, MDM9206, MDM9607, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in a DRM function, a buffer over-read can occur.
CVE-2015-9156 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MDM9206, MDM9607, MDM9635M, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 615/16/SD 415, SD 617, SD 800, SD 808, and SD 810, when making a high speed Dual Carrier Downlink Data call in a multicell environment, a buffer overflow may occur.
CVE-2015-9158 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in a QTEE crypto function, a buffer overflow can occur.
CVE-2015-9159 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, lack of input validation OEMCrypto_GetRandom can cause potential buffer overflow.
CVE-2015-9161 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 615/16/SD 415, SD 617, SD 650/52, SD 800, SD 808, and SD 810, TOCTOU condition could lead to a buffer overflow in function playready_reader_bind().
CVE-2015-9164 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, and SD 820A, a buffer overread in Playready may occur due to lack of input validation of the buffer size provided by HLOS.
CVE-2015-9170 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, incorrect offset check in wv_dash_core_refresh_keys() may lead to a buffer overread.
CVE-2015-9171 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, if OEMCrypto_Dash_InstallEncapKeybox() is called with keyBoxLength set to a value higher than TZ_WV_MAX_DATA_LEN (20k), a buffer over-read occurs.
CVE-2015-9172 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in a WideVine API function, a buffer over-read can occur.
CVE-2015-9173 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile SD 410/12, SD 617, SD 650/52, SD 800, SD 808, and SD 810, missing of return value check in memscpy can cause memory corruption in TQS App.
CVE-2015-9174 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile SD 410/12, SD 617, SD 650/52, SD 800, SD 808, and SD 810, lack of validation of the return value prior to using for buffer allocation in QSEE application, TQS, may result in memory overwrite.
CVE-2015-9177 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in a crypto API function, a buffer over-read can occur.
CVE-2015-9178 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, while processing the rmp secure command, memory corruption may result if the response buffer is smaller than the expected size.
CVE-2015-9179 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile MSM8974, lack of length checking in OEMCrypto_DeriveKeysFromSessionKey() could lead to a buffer overflow vulnerability.
CVE-2015-9180 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, the response pointer passed from user space to SDMX_process is not checked before it is used. If the given response buffer length is smaller than 16 bytes, the response values will be written to a memory outside the buffer, possibly in the secure memory area.
CVE-2015-9181 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, and SD 835, in a crypto API function, a buffer over-read can occur.
CVE-2015-9182 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, lack of input validation in OEMCrypto_GenerateSignature() can cause buffer over read.
CVE-2015-9184 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, and SD 835, lack of length checking in wv_dash_core_load_keys_v8() could lead to a buffer overflow vulnerability.
CVE-2015-9185 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in multiple Secure DEMUX functions (e.g., SDMX_open_session, SDMX_close_session, SDMX_set_session_cfg), when parameter validation fails, an error code is written into a response buffer, without checking that response buffer length (rsplen) passed from HLOS is large enough to hold the response. If the buffer is at the end of a non-secure page followed by secured memory page, this can cause a secure memory corruption.
CVE-2015-9186 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in a PlayReady API function, a buffer over-read can occur.
CVE-2015-9187 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, lack of buffer length validation in pvr_cmd_handler leads to unauthorized access to secure memory.
CVE-2015-9188 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 600, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, in Secure DEMUX command handler, when parameter validation fails, an error code is written into a response buffer without checking that response buffer length, passed from HLOS, which may result in memory corruption.
CVE-2015-9192 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, out of bounds memory access vulnerability may occur in the content protection manager due to improper validation of incoming messages.
CVE-2015-9195 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile MDM9625, MDM9635M, MDM9650, MDM9655, SD 400, SD 410/12, SD 615/16/SD 415, SD 617, SD 650/52, SD 808, SD 810, and SDX20, in a QTEE syscall handler, HLOS can cause a buffer overflow to occur.
CVE-2015-9199 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile and Snapdragon Mobile IPQ4019, MDM9625, MDM9635M, MDM9640, MDM9650, MDM9655, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 615/16/SD 415, SD 800, SD 808, SD 810, SD 820, and SD 820A, A non-secure region check is done while registering QSEE buffer address which is passed by HLOS but not while logging in the QSEE buffer, so corruption of dynamically protected secure region can occur if the non-secure buffer is changed between the time it's checked and when it's used.
CVE-2015-9202 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, while processing the content headers in the Playready module, a buffer overread may occur if the header count exceeds the expected value.
CVE-2015-9203 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, lack of input validation in playready_set_domainid could lead to a buffer overread.
CVE-2015-9204 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 410/12, SD 615/16/SD 415, SD 808, and SD 810, if cchFriendlyName is greater than TZ_PR_MAX_NAME_LEN in function playready_leavedomain_generate_challenge(), a buffer overread occurs.
CVE-2015-9205 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 410/12, SD 615/16/SD 415, SD 808, and SD 810, in a PlayReady API function, a buffer over-read can occur.
CVE-2015-9206 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 615/16/SD 415, SD 617, SD 650/52, SD 808, and SD 810, during XML encoding of a message in the Playready module, a buffer overread may occur if the message passed is large.
CVE-2015-9207 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Mobile and Snapdragon Wear MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 615/16/SD 415, SD 617, SD 650/52, SD 800, SD 808, and SD 810, lack of input validation in playready_getadditional_responsedata could lead to a buffer overread.
CVE-2015-9210 (v3: 9.8) 18 Apr 2018
In Android before 2018-04-05 or earlier security patch level on Qualcomm Snapdragon Automobile, Snapdragon Mobile, and Snapdragon Wear MDM9206, MDM9650, MSM8909W, SD 210/SD 212/SD 205, SD 400, SD 410/12, SD 425, SD 430, SD 450, SD 615/16/SD 415, SD 617, SD 625, SD 650/52, SD 800, SD 808, SD 810, SD 820, SD 820A, SD 835, SD 845, and SD 850, lack of input validation in playready_licacq_process_response() can lead to memory over read.